ArticlePDF Available

Novel and conserved functions of S-nitrosoglutathione reductase (GSNOR) in tomato

Authors:

Abstract and Figures

Nitric oxide (NO) is emerging as a key signalling molecule in plants. The chief mechanism for the transfer of NO bioactivity is thought to be S-nitrosylation, the addition of an NO moiety to a protein cysteine thiol to form an S-nitrosothiol (SNO). The enzyme S-nitrosoglutathione reductase (GSNOR) indirectly controls the total levels of cellular S-nitrosylation, by depleting S-nitrosoglutathione (GSNO), the major cellular NO donor. Here we show that depletion of GSNOR function impacts tomato (Solanum lycopersicum. L) fruit development. Thus, reduction of GSNOR expression through RNA interference (RNAi), modulated both fruit formation and yield, establishing a novel function for GSNOR. Further, depletion of S. lycopersicum GSNOR (SlGSNOR) additionally impacted a number of other developmental processes, including seed development, which also has not been previously linked with GSNOR activity. In contrast to Arabidopsis, depletion of GSNOR function did not influence root development. Further, reduction of GSNOR transcript abundance compromised plant immunity. Surprisingly, this was in contrast to previous data in Arabidopsis that reported reducing Arabidopsis thaliana GSNOR (AtGSNOR) expression by antisense technology increased disease resistance. We also show increased SlGSNOR expression enhanced pathogen protection, uncovering a potential strategy to enhance disease resistance in crop plants. Collectively, our findings reveal at the genetic level, some but not all GSNOR activities are conserved out with the Arabidopsis reference system. Thus, manipulating the extent of GSNOR expression may control important agricultural traits in tomato and possibly other crop plants.
Content may be subject to copyright.
Journal of Experimental Botany, Vol. 70, No. 18 pp. 4877–4886, 2019
doi:10.1093/jxb/erz234 Advance Access Publication 14 May, 2019
This paper is available online free of all access charges (see https://academic.oup.com/jxb/pages/openaccess for further details)
© The Author(s) 2019. Published by Oxford University Press on behalf of the Society for Experimental Biology.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/
by-nc/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the original work is properly cited. For commercial
re-use, please contact journals.permissions@oup.com
RESEARCH PAPER
Novel and conserved functions of S-nitrosoglutathione
reductase intomato
Adil Hussain1,5, Byung-Wook Yun2, JiHyun Kim3, KapugantiJagadis Gupta4,, Nam-In Hyung3 and
GaryJ. Loake5,*,
1 Department of Agriculture, Abdul Wali Khan University Mardan, Khyber-Pakhtunkhwa, Pakistan
2 School of Applied Biosciences, College of Agriculture and Life Sciences, Kyungpook National University, Republic of Korea
3 Department of Plant and Food Sciences, Sangmyung University, Cheonan, Republic of Korea
4 National Institute of Plant Genome Research, Aruna Asaf Ali Marg, 110067, Delhi, India
5 Institute of Molecular Plant Sciences, School of Biological Sciences, University of Edinburgh, King’s Buildings, Mayfield Road,
Edinburgh EH9 3BF, UK
* Correspondence: gloake@ed.ac.uk
Received 15 March 2019; Editorial decision 29 April 2019; Accepted 29 April 2019
Editor: Stanislav Kopriva, University of Cologne, Germany
Abstract
Nitric oxide (NO) is emerging as a key signalling molecule in plants. The chief mechanism for the transfer of NO
bioactivity is thought to be S-nitrosylation, the addition of an NO moiety to a protein cysteine thiol to form an
S-nitrosothiol (SNO). The enzyme S-nitrosoglutathione reductase (GSNOR) indirectly controls the total levels of cel-
lular S-nitrosylation, by depleting S-nitrosoglutathione (GSNO), the major cellular NO donor. Here we show that de-
pletion of GSNOR function impacts tomato (Solanum lycopersicum. L) fruit development. Thus, reduction of GSNOR
expression through RNAi modulated both fruit formation and yield, establishing a novel function for GSNOR. Further,
depletion of S. lycopersicum GSNOR (SlGSNOR) additionally impacted a number of other developmental processes,
including seed development, which also has not been previously linked with GSNOR activity. In contrast to Arabidopsis,
depletion of GSNOR function did not influence root development. Further, reduction of GSNOR transcript abundance
compromised plant immunity. Surprisingly, this was in contrast to previous data in Arabidopsis that reported that re-
ducing Arabidopsis thaliana GSNOR (AtGSNOR) expression by antisense technology increased disease resistance.
We also show that increased SlGSNOR expression enhanced pathogen protection, uncovering a potential strategy to
enhance disease resistance in crop plants. Collectively, our findings reveal, at the genetic level, that some but not all
GSNOR activities are conserved outside the Arabidopsis reference system. Thus, manipulating the extent of GSNOR
expression may control important agricultural traits in tomato and possibly other crop plants.
Keywords: Climacteric fruit, fruit development, GSNOR, MicroTom, nitric oxide, NO, S-nitrosation, S-nitrosylation, tomato,
tomato fruit.
Introduction
Nitric oxide (NO) underpins a plethora of cellular processes
integral to the biology of plants. The chief mechanism for the
transfer of NO bioactivity is thought to be S-nitrosylation, the
addition of an NO moiety to a peptide or protein cysteine thiol
to form an S-nitrosothiol (SNO) (Spadaro etal., 2010). This
redox-based post-translational modication controls a number
applyparastyle "g//caption/p[1]" parastyle "FigCapt"
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
4878 | Hussain etal.
of key activities related to growth, development, and environ-
mental interactions, including immune function. Typically, pro-
tein SNOs can be denitrosylated by the antioxidant tripeptide,
glutathione (GSH), resulting in the reconstitution of the pro-
tein Cys thiol and formation of S-nitrosogluthionine (GSNO)
(Airaki etal., 2011), which functions as a natural NO donor
and consequently a reservoir for NO bioactivity (Feechan
etal., 2005).
S-Nitrosoglutathione reductase (GSNOR), rst identied
in bacteria (Liu etal., 2001), is thought to be the major deter-
minant in the control of total cellular SNO levels in Arabidopsis
(Feechan etal., 2005; Lee etal., 2008; Leterrier etal., 2011). The
enzyme has a high anity for GSNO (Liu etal., 2001; Achkor
etal., 2003). Loss-of-function mutations in AtGSNOR1 com-
promise multiple modes of plant disease resistance, while
overexpression of this gene conveys increased disease resistance.
Further, AtGSNOR1 has been shown to regulate both salicylic
acid (SA) biosynthesis and associated signalling (Feechan etal.,
2005; Rustérucci etal., 2007; Tada etal., 2008).
In the context of SA signalling (Loake and Grant, 2007; Fu
and Dong, 2013) , S-nitrosylation of the A.thaliana SA-binding
protein 3 (AtSABP3) at Cys280 suppresses its binding to both
the immune activator, SA, and the carbonic anhydrase activity
of this protein, negatively regulating disease resistance (Y.J.
Wang et al., 2009). Further, NO via GSNO has been shown
to protect the TGA1 transcriptional regulator from oxygen-
mediated modications and enhance the DNA binding activity
of this protein to its cognate cis-element in the presence of
the transcriptional co-activator, NPR1. In addition, the trans-
location of NPR1 into the nucleus may be promoted by NO
(Lindermayr etal., 2010). In contrast, GSNO accumulation in
atgsnor1-3 plants has been reported to inhibit the translocation
of NPR1 from the cytoplasm to the nucleus, thereby curbing
SA signalling and associated plant immunity (Tada etal., 2008;
Yun et al., 2016). NO is also proposed to play a central role
in signalling activated by the fungal elicitor, cryptogein (Kulik
etal., 2015), and is required for disease resistance against Botrytis
cinerea triggered by oligogalacturonides (Rasul etal., 2012).
Both NO and reactive oxygen intermediates (ROIs) have
been implicated in the programmed cell death of pathogen-
challenged cells through the hypersensitive response (HR;
Delledonne et al., 1998, 2001; Torres et al., 2002). Aloss-of-
function allele of atgsnor1 [paraquat resistance 1-2 (par1-2)] con-
veyed protection against cell death mediated by the herbicide,
paraquat (Chen et al., 2009). Further, NO has been shown
to regulate the production of ROIs by the S-nitrosylation
of NADPH oxidase (AtRBOHD) at Cys890, reducing ROI
production at later stages of the plant defence, curbing devel-
opment of the HR (Yun et al., 2011). Interestingly, oxidative
post-translational modication of GSNOR inhibited the ac-
tivity of this enzyme, suggesting an additional mechanism of
direct crosstalk between ROI and NO signalling (Frungillo
etal., 2014; Guerra etal., 2016; Kovacs etal., 2016; Lindermayr,
2018).
AtGSNOR1 has also been shown to control some key
aspects of plant development (Lee etal., 2008; Leterrier etal.,
2011; Kwon etal., 2012). For example, atgsnor1-3 mutants show
loss of apical dominance, a subtle change in leaf shape, and
increased sensitivity to auxin (Kwon et al., 2012). This line is
also reduced in fertility, principally due to very short stamens
which do not function as eective self-pollinators (Kwon etal.,
2012; Xu etal., 2013). AtGSNOR1 has also been implicated in
responses to abiotic stress (Corpas etal., 2011; Fancy etal., 2017;
Begara-Morales etal., 2018). Missense alleles of hot5/atgsnor/
par2 cannot acclimate to heat as do dark-grown seedlings, but
grow normally and can heat-acclimate in the light. In contrast,
null alleles cannot heat-acclimate like light-grown plants (Lee
etal., 2008). Thus, AtGSNOR is required for heat acclimation.
In sunower seedlings exposed to high temperature (38°C for
4h), GSNOR activity, protein levels, and transcript abundance
have been found to be reduced in hypocotyls, with the sim-
ultaneous accumulation of SNOs (Leterrier et al., 2011). The
consequence was a rise in protein tyrosine nitration, which
is considered a marker of nitrosative stress. Collectively, these
ndings imply that AtGSNOR also has an important function
in plant development and abiotic stress.
While the genetics of tomato (Solanum lycopersicum) GSNOR
(SlGSNOR) have largely been unexplored, the crystal structure
for the corresponding enzyme has been solved to 1.9Å reso-
lution (Kubienová etal., 2013), being the rst plant GSNOR
to be structurally determined. Here, taking a genetic approach,
we uncover key roles for SlGSNOR in plant development,
fruit formation, and immunity in tomato. Collectively, these
data imply that the function of GSNOR is conserved across
plant species. Further, manipulating levels of GSNOR expres-
sion may provide novel mechanisms for the incorporation of
disease resistance and advantageous developmental traits into
crop plants.
Materials and methods
DNA constructs
Tomato GSNOR (Solyc09g064370, http://solgenomics.net/
locus/34669/view) was amplied using SlGSNOR-PstI-F and
SlGSNOR-NotI-R primers (Supplementar y Table S1 at JXB online) and
cloned behind the CaMV2x35S promoter in the pGreenI0029 binary
expression vector to make the GSNOR overexpression (OE) DNA con-
struct. For the SlGSNOR-RNAi DNA construct, sense and antisense
DNA fragments of 369bp were amplied using SlGSNORS-XhoI-F and
SlGSNORS-KpnI-R for the sense fragment, and SlGSNORA-ClaI-F
and SlGSNORA-XbaI-R for the antisense fragment (Table 1), and cloned
in the pHANNIBAL intermediate vector separated by a pDK intron to
make the CaMV35S:sense:intron:antisense:terminator RNAi cassette.
The cassette was then transferred to the pGreenI0029 binary vector.
RNA extraction was performed using an RNeasy Mini Kit (Qiagen)
according to the manufacturer’s instructions. Atotal of 1 µg of RNA
was reverse transcribed to synthesize cDNA using the Omniscript RT
kit (Qiagen) according to the manufacturer’s instructions. A1µl aliquot
of this cDNA was subsequently used in a semi-quantitative PCR for the
amplication of DNA fragments. All DNA fragments were amplied in
a 40μl reaction using Phusion® High Fidelity DNA polymerase (New
England Biolabs). Both the OE and RNAi SlGSNOR DNA constructs
were conrmed by colony PCR and sequenced before transformation in
tomato cultivar MicroTom.
Tomato transformation
The SlGSNOR-OE and SlGSNOR-RNAi DNA constructs were
transformed into tomato cv. MicroTom. Briey, seeds of tomato variety
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
GSNOR function in tomato | 4879
MicroTom were surface sterilized in 70% ethanol for 40s, rinsed with
sterile distilled water, kept in 40% sodium hypochlorite (NaOCl)+3
drops/100 ml Tween-20, for 15 min, and rinsed with sterile distilled
water ve times. The seeds were germinated on germination medium
[1/2 MS (Murashige and Skoog, 1962) salts and vitamins, 3% sucrose,
0.8% agar] in the dark at 25±2°C for 7 d.Single Agrobacterium tumefaciens
colonies carrying SlGSNOR-OE and SlGSNOR-RNAi constructs were
grown at 28°C with shaking at 250 rpm for 24h. The cultures were
centrifuged at 12 000rpm for 10 min and bacteria were washed and
re-suspended in liquid MS medium to OD600=0.7. MicroTom explants
were prepared by cutting both ends of the cotyledons and dipped in
A. tumefaciens SlGSNOR-OE and SlGSNOR-RNAi suspension cul-
tures for 5min. The explants were then incubated on shoot induction
medium [SIM: 1/2 MS salts and vitamins, 3% sucrose, 0.8% agar, 2mg
l–1 6-benzylaminopurine (BA), 0.01mg l–1 indole-3-butyric acid (IBA)]
for 2 d and rinsed with sterile distilled water. The explants were then
screened on SIM plates containing 25 mg l–1 kanamycin+500 mg l–1
cefotaxime (transgenic shoot selection medium: TSM). Transgenic shoots
were transferred to fresh TSM every 4 weeks to ensure stringent selection.
Transgenic shoots (~0.5cm) were then transferred to transgenic shoot
elongation medium (TEM: MS salts and vitamins, 3% sucrose, 0.8% agar,
30mg l–1 kanamycin+500mg l–1 cefotaxime) at 25±2 °C in the light.
Plants were transferred to fresh TEM every 3–4 weeks and then to sterile
soil in pots after rooting. Homozygous transgenic plants were obtained as
described previously (Harrison etal., 2006).
Growth of tomatoplants
Seeds from wild-type (WT) and transgenic plants were germinated ei-
ther on 1/2 MS medium (1.1g of MS salt and 5g of sucrose dissolved
in 300ml of water at pH 5.7–5.9 and volume adjusted to 500ml after
adding 4g of agar) or a special peat-based UC (University of California)
compost [100 litres of medium grade peat (Sinclair Horticulture), 25
litres of horticultural sand, 375g of garden limestone (J. Arthur Bowers),
150g of Osmocote Exact 3–4 months (Scotts) and 3 g of Intercept-
70WG (Scotts)]. Seedlings were transplanted to new pots 1 week after
germination and grown at 21°C under long days (16h light/8h dark)
at 800 µmol m−2 s−1 PAR (photosynthetically active radiation) light
intensity.
Growth and inoculation of PstDC3000
Pseudomonas syringae pv. tomato DC3000 (PstDC3000) was grown in LB
liquid medium [tryptone 10 g l–1, yeast extract (Oxoid) 5g l–1, NaCl
(VWR, UK) 10g l–1], with 50µg ml–1 rifampicin at 28°C overnight.
Cells were pelleted by centrifugation before re-suspension in 10 mM
MgCl2. Plants were inoculated as described by Mudgett and Staskawicz
(1999) with some modication. Three-week-old plants were sprayed with
a PstDC3000 virulent suspension with cell density adjusted to 2×108 cfu
ml–1 at OD600. Plants were kept covered inside plastic bags under high
humidity for 24h to allow the opening of stomata for successful bacterial
entrance/inoculation.
PstDC3000 colonycounts
PstDC3000 was inoculated to the tomato WT and transgenic plants as de-
scribed above. The plants were examined for disease symptoms at regular
intervals. Leaf samples were collected at 0, 2, and 4days post-infection
(DPI). Using 12 plants per line, three leaf discs (1cm2) were collected per
plant. Each leaf disc was ground in a microfuge tube in 500µl of 10mM
sterile MgCl2 using a tissue lyser (Qiagen/Retsch) for 2min at 30 shakes
per second. A200µl aliquot of the bacterial suspension was transferred to
a new microfuge tube, and serial dilutions were made to 10–2. After the
serial dilution, 10µl of each dilution was plated on NYG plates [Bacto
peptone 5g l–1, yeast extract (Oxoid) 3g l–1, glycerol (Fisher Scientic)
20ml l–1, Bacto agar 15g l–1] containing 50µg ml–1 rifampicin. The plates
were incubated for 2 d at 28°C and the number of bacterial colonies for
each sample counted and recorded in the best countable dilution. The ex-
periment was repeated three times.
PR1 gene expression
Leaf samples collected for colony counts, from PstDC3000-infected WT
and transgenic plants at 0, 2, and 4 DPI, were used to cut 1cm2 leaf discs
for colony count assay. The rest of the leaf samples were used to extract
RNA for PR gene expression analysis. RNA was isolated from plant sam-
ples as described earlier, and quantication of reverse transcription–PCR
(RT–PCR) was carried out to check PR1 expression in response to in-
fection. Tomato actin was used as a reference gene. Primers for tomato
PR1 and actin genes are given in Supplementary Table S1.
Salicylic acid measurement
Free and conjugated endogenous SA levels were determined using HPLC, as
described by Aboul-Soud etal. (2004) with minor modications. A200mg
aliquot of leaf tissue per sample was collected and promptly frozen in liquid
nitrogen. Samples were then ground in liquid nitrogen using a mortar and
pestle, and transferred to a 2ml microfuge tube, followed by the addition of
1ml of 90% methanol (Fisher Scientic), and vortexed for 1min, before the
sample thawed. The sample was then centrifuged at 15000 g for 5min and
the supernatant transferred to a new tube. The pellet was re-extracted in 1ml
of 100% methanol, centrifuged, and the two supernatants pooled together
and dried in a speed vacuum centrifuge (Speed Vac DNA110, Savant) at me-
dium temperature. The residue resulting from drying the supernatant was then
re-suspended in 1ml of 5% trichloroacetic acid, followed by the addition
of 1ml of ethyl acetate:cyclopentane (Fluka):isopropanol (Fisher Scientic)
(50:50:1) and vortexed for 1min. The organic phase was transferred to a new
tube. The aqueous phase was re-extracted with another 1ml of the organic
50:50:1 mix, and the two supernatants pooled together and evaporated under
heat in the vacuum centrifuge. The aqueous phase was then acidied to pH 1
by addition of 50µl of absolute HCl, boiled for half an hour to release SA from
any acid-labile conjugated forms, and extracted with the organic mix twice.
The two supernatants were pooled together and dried in the vacuum centri-
fuge. The residues were dissolved in 100μl of 100% methanol before 100μl of
H2O was added to give a nal 50% (v/v) methanol concentration. The sam-
ples were ltered through a 0.25μM lter (Millex-GP, Millipore Corporation,
Billerica, MA, USA) and subjected to HPLC analysis. Samples were taken at 4
DPI. SA samples of 1mM and 10mM were used as standard.
Results
SlGSNOR depletion negatively affects seed
development and germination but not root
development
After the transformation of SlGSNOR-OE and SlGSNOR-
RNAi constructs in tomato (cv. MicroTom), SlGSNOR-RNAi
and SlGSNOR-OE lines were generated. RT–PCR results
showed a signicant reduction in SlGSNOR expression in the
RNAi lines, whereas a signicant increase in SlGSNOR expres-
sion was observed in the OE lines (Fig. 1A). Asignicant impact
of SlGSNOR knockdown on the germination percentage was
observed in the RNAi lines, with >80% reduction in germin-
ation (Fig. 1B). This shows that the accumulation of SlGSNOR1
transcripts is tightly regulated in tomato MicroTom and a signi-
cant reduction in its expression leads to lethality, as SlGSNOR-
RNAi lines with greater than ~60% reduction in SlGSNOR1
expression were not viable (Fig. 1C). Both SlGSNOR-RNAi
and SlGSNOR-OE lines conveyed signicant eects on the
overall development of tomato plants, ranging from seed ger-
mination to fruiting and net yield per plant. Reduced GSNOR
expression in SlGSNOR-RNAi lines drastically aected seed
development and reduced the number of seeds produced in
the fruits of the resulting transgenic plants. In representative
RNAi lines that could not be maintained, the seeds were small,
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
4880 | Hussain etal.
misshapen, and lacked endosperm. Consequently, these seeds
failed to germinate. In a representative fertile SlGSNOR-RNAi
line (2-2-2), seed germination was reduced by 80% relative to
the WT. The SlGSNOR-OE plants, however, also showed an
average reduction of 20% in germination frequency (Fig. 1B),
although OE plants produced normal healthy seeds (Fig. 1C).
Counterintuitively, the SlGSNOR-RNAi plants also showed
faster germination as compared with WT and SlGSNOR-OE
plants on either MS medium or soil, and showed the appearance
of fresh green tissues at least 1 d before the WT and OE plants.
Suppression of SlGSNOR in tomato did not seem to have
a major impact on the root system of plants under optimal en-
vironmental conditions. The root length of 1- and 5-week-old
plants was analysed (Fig. 2A, B). Root length in SlGSNOR-
RNAi plants was not signicantly dierent from that of WT
plants. In a similar fashion, root length in SlGSNOR-OE plants
was also not visibly dierent from that of WT plants (Fig. 2A,
B). Quantitative analysis and associated statistical testing con-
rmed that the root length of SlGSNOR-RNAi plants and
SlGSNOR-OE plants is not statistically dierent from that of
the WT (Fig. 2C). Collectively, our data imply that reducing
SlGSNOR gene expression impacts tomato seed develop-
ment and germination, but not root development. In contrast,
increasing SlGSNOR expression does not impact either seed
development, seed germination, or root development.
SlGSNOR is required for leaf development
Multiple developmental phenotypes were found to be dif-
ferent in the transgenic plants as compared with WT plants.
Therefore, the average leaf area for WT, SlGSNOR-RNAi, and
SlGSNOR-OE tomato plants was calculated using 5-week-old
plants. Leaf area was measured as an average of multiple leaves of
dierent sizes, ranging from the smallest to the largest and from
the bottom to the top of the plants. The smallest average leaf area
of 468.49mm2 was recorded for SlGSNOR-RNAi plants, com-
pared with 734.35mm2 for the WT. Tomato transgenic plants
overexpressing SlGSNOR had the largest leaves, with an area of
783.10mm2 (Fig. 3A, B) as compared with leaves of WT plants,
although the dierence was not statistically signicant.
SlGSNOR regulates fruit production and flower
development
WT and transgenic SlGSNOR-RNAi and SlGSNOR-OE
tomato plants were analysed for their respective yields. Both
the SlGSNOR-RNAi and SlGSNOR-OE plants showed re-
duced yield per plant as compared with the WT. However,
SlGSNOR-RNAi plants produced an average of 73.56 g of
fruit per plant as compared with 63.96g for SlGSNOR-OE and
179.10g for WT plants (Fig. 4A, B). However, fruits produced
by the SlGSNOR-OE plants were signicantly larger in size
as compared with those of WT plants, whereas those produced
by the SlGSNOR-RNAi plants were smaller in size (Fig. 4B) .
Consistent with the atgsnor1 Arabidopsis mutant, the tomato
SlGSNOR-RNAi plants produced misshapen owers with car-
pels beyond the reach of stamens, which presumably negatively
impacted self-fertilization (Fig. 4C–H). In contrast, SlGSNOR
overexpression had no eect on the oral phenotype; these lines
resembled WT plants with respect to these traits (Fig. 4C–H).
Fig. 1. SlGSNOR suppression negatively affects seed development and germination. (A) Quantification of RT–PCR results showing SlGSNOR transcript
levels in representative tomato WT, RNAi, and OE plants. (B) Percentage germination of tomato WT, RNAi, and OE seeds. Ahighly significant reduction
in germination frequency was recorded for the seeds produced by RNAi plants as compared with WT plants. (C) Phenotype of the WT, RNAi, and
OE seeds. SlGSNOR-RNAi plants produced misshapen, small, and deformed seeds as compared with WT and OE plants. Reduction of SlGSNOR
expression by up to ~60% resulted in lethality and the seeds could not germinate, However, the OE plants produced seeds with a normal phenotype.
Statistical analyses were performed through one-way ANOVA test at a 95% level of confidence. Statistically significant differences are shown by an
asterisk (*). Error bars represent the SD. (This figure is available in colour at JXB online.)
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
GSNOR function in tomato | 4881
Overexpression of SlGSNOR promotes resistance to
bacterial pathogens
The tomato cv. MicroTom is not well characterized with respect
to microbial pathogens. The well-characterized leaf pathogen,
PstDC3000, is thought to be an opportunistic pathogen of
this tomato cultivar, with little increase in growth over time
(Takahashi et al., 2005). Three-week-old WT, SlGSNOR-
RNAi, and SlGSNOR-OE plants were spray inoculated with
a PstDC3000 suspension of 2×108 cfu ml–1 in 10mM MgCl2
and 0.02% Silwet L77. This ensured even application of in-
oculum and avoided potential injury. Development of disease
symptoms was monitored daily. Disease symptoms appeared in
SlGSNOR-RNAi plants after 7 d.Chlorosis, the appearance
of typical dark brown lesions surrounded by chlorotic areas,
was clearly visible, especially near leaf margins, on the leaves
of SlGSNOR-RNAi plants. These symptoms are typical of
PstDC3000 infection in susceptible tomato plants. In contrast,
WT and SlGSNOR-OE plants showed delayed and reduced
symptom development (Fig. 5A). Leaf samples were collected
from the inoculated plants at 2 and 4 DPI for the determin-
ation of bacterial titre. SlGSNOR-RNAi plants supported
an increased bacterial titre relative to WT plants at both time
Fig. 2. SlGSNOR suppression does not negatively impact root development. No significant differences were found between the root length of WT, RNAi,
and OE plants after 1 week (A) or 5 weeks of growth (B). Root length measurements of WT, RNAi, and OE plants were statistically analysed using a two-
way ANOVA test at a 95% level of significance. Error bars represent the SD. (This figure is available in colour at JXB online.)
Fig. 3. SlGSNOR suppression reduces average leaf area in tomato. (A) Scanned image showing the arrangement of WT, SlGSNOR-RNAi, and
SlGSNOR-OE plant leaves along with a UK 10 pence coin as a scale marker. (B) Significantly reduced average leaf area was recorded for the RNAi plants
compared with WT and OE plants. However, the increase in the leaf area for OE plants compared with the WT was not significant. Statistical analysis was
performed using one-way ANOVA with 95% confidence Statistical analyses were performed through one-way ANOVA test at a 95% level of confidence.
Statistically significant differences are shown by an asterisk (*). Error bars represent the SD.
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
4882 | Hussain etal.
points. Conversely, the number of bacteria in SlGSNOR-OE
plants was signicantly reduced relative to the WT (Fig. 5B).
Thus, reduction of SlGSNOR expression promotes enhanced
disease susceptibility. In contrast, increased SlGSNOR ex-
pression enhances disease resistance against PstDC3000. Thus,
modulation of SlGSNOR transcript abundance impacts the
level of basal disease resistance.
In Arabidopsis, GSNOR has been shown to be a positive
regulator of both SA synthesis and associated signalling (Feechan
etal., 2005; Tada etal., 2008). Therefore, to explore the molecular
mechanism underpinning the regulation of basal disease resist-
ance by SlGSNOR, we rst determined the levels of SA in WT,
SlGSNOR-RNAi, and SlGSNOR-OE plants. The basal and
pathogen-induced levels of SA in SlGSNOR-RNAi plants were
55% and 57.74%, respectively, of those present in WT plants.
The concentrations of SA found in the GSNOR-overexpressing
plants were 360% and 132.8% higher than those in the WT be-
fore and after infection, respectively (Fig. 5C).
Presumably, these changes in the levels of SA impact the ex-
pression of SA-dependent genes, including the well-established
SA marker gene, PR1 (Uknes etal., 1992). We therefore deter-
mined the level of PR1 gene expression in WT, SlGSNOR-
RNAi, and SlGSNOR-OE plants in response to pathogen
challenge. Reduced PR1 transcript accumulation was observed
in SlGSNOR-RNAi lines 2 and 4 DPI with respect to the
WT. Conversely, PR1 gene expression was signicantly in-
creased in SlGSNOR-OE plants at 2 and 4 DPI relative to the
WT (Fig. 5D).
Discussion
Our data show that depletion of SlGSNOR levels impacts
growth and development of tomato cv. MicroTom. Reduction
of SlGSNOR function results in the loss of apical domin-
ance, changes in leaf shape, perturbations in seed development
and germination, and, signicantly, a reduction in the yield of
Fig. 4. Manipulation of SlGSNOR levels impacts fruit production and flower development. (A) Average yield (g per plant) of WT, SlGSNOR-RNAi, and
SlGSNOR-OE tomato plants. Both SlGSNOR-RNAi and SlGSNOR-OE transgenic lines showed a highly significant reduction in yield. (B) OE plants
produced large, healthy fruits with a good number of healthy viable seeds, while RNAi plants produced small fruits typically <2.5cm in diameter with few
and mostly non-viable seeds. (C–H) SlGSNOR-RNAi plants produced misshapen flowers with long carpels extending beyond the reach of the stamens
(D, G) compared with the WT (C, F) and OE plants (E, H). Statistical analyses were performed through one-way ANOVA test at a 95% confidence level.
Statistically significant differences are shown by an asterisk (*). Error bars represent the SD. (This figure is available in colour at JXB online.)
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
GSNOR function in tomato | 4883
tomato fruit. In contrast, overexpression of SlGSNOR has no
impact on the growth of tomato cv. MicroTom. Our ndings
also highlight a key role for SlGSNOR in disease resistance.
Thus, depletion of SlGSNOR levels resulted in enhanced
disease susceptibility to the bacterial pathogen, PstDC3000.
Conversely, overexpression of SlGSNOR promoted disease re-
sistance due to increased SA accumulation and associated ex-
pression of SA-dependentgenes.
While AtGSNOR has been extensively studied in
Arabidopsis, there been no previous information on the
genetics of GSNOR in crop plants. In contrast, the eect
of NO on seed germination, root architecture, and fruit
ripening has been studied employing NO donors in crop
plants (Zandonadi etal., 2010; Semchuk etal., 2011). Analysis
of GSNO in the main organs of pepper plants established
that this metabolite was most abundant in roots, followed by
leaves and stems. These ndings directly correlated with the
content of NO in each organ and inversely correlated with
GSNOR activity (Airaki etal., 2011). Subcellular localization
of GSNO in pea leaves established the presence of GSNO
in the cytosol, chloroplasts, mitochondria, and peroxisomes
(Barroso etal., 2013). While these studies have provided excel-
lent and compelling circumstantial evidence for a key role for
GSNO and, by extension, GSNOR in plant developmental
processes in crop plants, direct genetic evidence has not been
established. Our data show that depletion of SlGSNOR does
indeed impact a number of development processes outside
the model plant, Arabidopsis. Thus, reduction of SlGSNOR
transcript accumulation decreases individual leaf size and con-
sequently total leaf area. Seed size is also reduced and seed
Fig. 5. Overexpression of SlGSNOR promotes disease resistance. (A) Development of disease symptoms in the stated lines after 1 week of infection
with PstDC3000. Yellow chlorotic areas surrounding dark brown lesions can be seen on the leaves of RNAi plants. (B) Graph showing bacterial growth
after 2 d and 4 d of infection. Significantly higher bacterial growth was observed in RNAi plants, whereas the OE plants supported significantly lower
bacterial growth after 2 d and 4 d of infection. (C) Total salicylic acid (SA) levels were measured in unchallenged leaves, in addition to plants infected with
PstDC3000 after 4 d of infection. SlGSNOR-OE plants showed a significantly higher level of basal and induced SA compared with WT plants. On the
other hand, the RNAi plants produced lower quantities of SA both before and after infection. (D) Expression of the SA marker gene, PR1, was found to
be significantly higher in the OE plants after 2 d and 4 d of infection by PstDC3000 as compared with that in WT plants. RNAi plants showed significantly
lower PR1 expression compared with WT plants. Statistical analyses were performed using two-way ANOVA test at a 95% level of confidence.
Statistically significant differences are shown by an asterisk (*). Error bars represent the SD. (This figure is available in colour at JXB online.)
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
4884 | Hussain etal.
from SlGSNOR-RNAi plants exhibit a lower germination
frequency. Signicantly, fruit size and total fruit yield are also
reduced.
NO bioactivity has been strongly linked to plant repro-
ductive biology (Bright etal., 2009; Zafra etal., 2010). Thus, NO
can act as a negative regulator of pollen tube growth in plants
such as Lilium longiorum, Arabidopsis, and Paulownia tomentosa
(Prado etal., 2004, 2008; He etal., 2007). Conversely, NO has
been reported as a positive stimulus of pollen tube growth in
Pinus bangeana, functioning in a dose-dependent manner (Y.
Wang etal., 2009). In SlGSNOR-RNAi plants, our data show
that the structure of the reproductive organs was impacted;
these lines developed long carpels, resulting in the stigma being
spatially removed from the surrounding anthers, decreasing
pollen transfer and, by extension, self-fertility. These pheno-
types parallel those observed in Arabidopsis plants possessing
null mutations in AtGSNOR (Kwon etal., 2012), implicating
conservation of GSNOR function from Arabidopsis to to-
mato across a number of developmental processes with vis-
ible outcomes. Interestingly, null mutations in Arabidopsis
AtGSNOR also perturb root development, resulting in shorter
roots. However, in contrast, depletion of SlGSNOR transcripts
did not visibly impact root development. Perhaps sucient
SlGSNOR activity was still present in the relevant root cells
of these plants to enable the completion of key growth and/or
developmental processes in this organ. Alternatively, a role for
GSNOR function in root development may not be conserved
between Arabidopsis and tomato. GSNOR is a single-copy
gene in both tomato and Arabidopsis. Our ndings suggest that
strong reduction of SlGSNOR expression resulted in the for-
mation of non-viable seeds. Thus, null mutants of SlGSNOR
might not be maintained in tomato cv. MicroTom and perhaps
other tomato cultivars.
Ripening of both climacteric (e.g. tomato) and non-
climacteric (e.g. pepper) fruits is another area where NO func-
tion and associated S-nitrosylation have been explored (Corpas
etal., 2018). Climacteric fruits continue ripening after being
picked, a process accelerated by ethylene. Non-climacteric
fruits can ripen only when still attached to their respective
plant. These fruits have a short shelf-life if harvested when ripe.
The application of NO gas or NO donors to a number of dif-
ferent climatic fruits has been shown to delay fruit ripening.
In this context, NO can repress both ethylene metabolism
and signalling, while simultaneously inducing antioxidative
enzymes, which are thought to prevent oxidative damage.
Intriguingly, NO gas has also been shown to delay fruit
ripening in non-climacteric fruits and, in addition, increase the
amount of ascorbate (vitamin C) (Rodríguez-Ruiz etal., 2017;
Corpas etal., 2018). Thus, NO treatment of non-climatic fruits
may convey dual advantages: extending both fruit shelf-life and
quality. This exciting research has clearly uncovered a potential
biotechnological application of NO (Manjunatha etal., 2010;
Corpas etal., 2018). Our ndings suggest that continuous de-
pletion of SlGSNOR function and, by extension, increasing
GSNO and associated global S-nitrosylation, both decreases
the size of individual fruits and reduces the overall fruit yield.
It will now be interesting for future studies to explore the bio-
chemical composition and shelf-life of these fruits. Therefore,
to maximize the potential utility of NO to augment both fruit
shelf-life and quality, insights into the molecular mechanisms
whereby NO and cognate S-nitrosylation control these pro-
cesses will be important, because our data suggest that too
much NO/GSNO can have adverse eects on both individual
fruit size and totalyield.
The role of GSNO and NO during immunity in crop plants
remains relatively unclear, because a genetic analysis has not
complemented the biochemical studies to date. Two sunower
(Helianthus annuus L.) cultivars either resistant or susceptible to
infection by the downy mildew pathogen, Plasmopara halstedii,
were employed to investigate the role of GSNO and related
reactive nitrogen intermediates (RNIs) in the immune re-
sponse of this plant. In the susceptible cultivar, an increase in
both protein tyrosine nitration and SNOs was detected, inde-
pendent of NO generation, suggesting that microbial patho-
gens induce nitrosative stress in susceptible sunower cultivars.
Conversely, in the resistant cultivar, there was no increase in
either protein tyrosine nitration or SNOs, implying an absence
of nitrosative stress. Therefore, protein tyrosine nitration might
mark nitrosative stress in plants during microbial infection
(Chaki etal., 2009). In potato, after challenge with an avirulent
Phytophthora infestans isolate, relatively high levels of GSNO and
SNOs concentrated in the main vein of potato leaves, implying
a possible mobile function of these compounds in the transfer
of NO bioactivity. In contrast, during a virulent P. infestans in-
fection, low-level production of NO and ROIs occurred; it was
proposed that this might result in the delayed up-regulation of
PR genes and the subsequent compromised resistance towards
this pathogen (ArasimowiczJelonek etal., 2016). Moreover, in
lettuce (Lactuca sativa), a GSNOR-mediated decrease of SNOs
was found to be a general feature of lettuce responses to both
downy and powdery mildew infection, while resistance to
Bremia lactucae, the causal agent of lettuce downy mildew, was
found to parallel an increase of GSNOR activity. Thus, modu-
lation of GSNOR activity appears to play a key role in lettuce–
mildew interactions (Tichá etal., 2018).
In Arabidopsis, loss-of-function mutations in AtGSNOR
result in an increase in total S-nitrosylation and a reduction
in SA biosynthesis and associated signalling, compromising
disease resistance (Feechan etal., 2005; Tada etal., 2008; Yun
et al., 2016). In complete contrast, depletion of AtGSNOR
transcripts has been reported to result in disease resistance
(Rustérucci etal., 2007). This may reect the complex role of
(S)NO in plant immunity. Thus, depleting AtGSNOR levels
may increase SNO levels less relative to a null mutation in
AtGSNOR and this dierence in relative SNO concentrations
may result in dierent immune ouputs (i.e. resistance versus
susceptibility, respectively). If this posit is correct, our deple-
tion of SlGSNOR transcripts to a similar extent in tomato
might be predicted to lead to increased disease resistance. Our
ndings show that depletion of GSNOR function in tomato
results in decreased SA biosynthesis and signalling, leading to
compromised basal resistance. This surprisingly contrasts with
previous data showing that depletion of AtGSNOR transcripts
results in disease resistance (Rustérucci etal., 2007); however,
it is similar to other data proposing that null mutations in
AtGSNOR compromise plant immunity (Feechan etal., 2005;
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
GSNOR function in tomato | 4885
Tada etal., 2008; Yun et al., 2011). Therefore, it appears un-
likely that these previous ndings (Rustérucci et al., 2007)
can be explained by dierences in relative SNO concentra-
tions between AtGSNOR-depleted lines and those possessing
null mutations in AtGSNOR (Feechan etal., 2005; Tada etal.,
2008; Yun etal., 2011).
Importantly, our ndings show that the function of GSNOR
in disease resistance appears to be conserved from Arabidopsis
to tomato. Signicantly, the overexpression of AtGSNOR
in Arabidopsis conveyed broad-spectrum disease resistance,
without constitutive SA accumulation or associated signalling
(Feechan etal., 2005). Rather, AtGSNOR overexpression sup-
ported a potentiation of SA-dependent gene expression fol-
lowing attempted pathogen infection. Moreover, this resistance
was not associated with a negative impact on growth or a yield
penalty under laboratory conditions (Feechan etal., 2005), sug-
gesting that manipulation of GSNOR activity might provide
a novel mechanism to convey broad-spectrum disease resist-
ance in crop plants. Our data suggest that the overexpression
of SlGSNOR also does not negatively impact growth, but it
does decrease total fruit yield. However, on the positive side,
overexpression of SlGSNOR signicantly increased the size of
individual tomato fruits, which might be attractive for some
markets. Therefore, manipulating SlGSNOR expression via
traditional crop breeding or gene editing approaches might
provide novel strategies to convey disease resistance and per-
haps also modulate the properties of tomato fruits.
Supplementarydata
Supplementary data are available at JXB online.
Table S1. List of primers used in RT–PCR.
Acknowledgements
AH was supported by Higher Education Commission (HEC) Pakistan
for PhD Studentship. Research in the laboratory of GJL has been sup-
ported by BBSRC grant BB/DO11809/1.
References
Aboul-SoudM, Cook K, LoakeG. 2004. Measurement of salicylic acid
by a high-performance liquid chromatography procedure based on ion-
exchange. Chromatographia 59, 129–133.
AchkorH, DíazM, FernándezMR, BioscaJA, ParésX, MartínezMC.
2003. Enhanced formaldehyde detoxification by overexpression of
glutathione-dependent formaldehyde dehydrogenase from Arabidopsis.
Plant Physiology 132, 2248–2255.
Airaki M, Sánchez-Moreno L, Leterrier M, Barroso JB, Palma JM,
Corpas FJ. 2011. Detection and quantification of S-nitrosoglutathione
(GSNO) in pepper (Capsicum annuum L.) plant organs by LC-ES/MS. Plant
& Cell Physiology 52, 2006–2015.
ArasimowiczJelonekM, FloryszakWieczorekJ, IzbiańskaK, GzylJ,
JelonekT. 2016. Implication of peroxynitrite in defence responses of potato
to Phytophthora infestans. Plant Pathology 65, 754–766.
Barroso JB, Valderrama R, Corpas FJ. 2013. Immunolocalization of
S-nitrosoglutathione, S-nitrosoglutathione reductase and tyrosine nitration
in pea leaf organelles. Acta Physiologiae Plantarum 35, 2635–2640.
Begara-Morales JC, Chaki M, Valderrama R, Sánchez-Calvo B,
Mata-PérezC, PadillaMN, CorpasFJ, BarrosoJB. 2018. Nitric oxide
buffering and conditional nitric oxide release in stress response. Journal of
Experimental Botany 69, 3425–3438.
BrightJ, HiscockSJ, JamesPE, HancockJT. 2009. Pollen generates
nitric oxide and nitrite: a possible link to pollen-induced allergic responses.
Plant Physiology and Biochemistry 47, 49–55.
Chaki M, ValderramaR, Fernández-Ocaña AM, et al. 2009. Protein
targets of tyrosine nitration in sunflower (Helianthus annuus L.) hypocotyls.
Journal of Experimental Botany 60, 4221–4234.
Chen R, Sun S, Wang C, et al. 2009. The Arabidopsis PARAQUAT
RESISTANT2 gene encodes an S-nitrosoglutathione reductase that is a key
regulator of cell death. Cell Research 19, 1377–1387.
Corpas FJ, Freschi L, Rodríguez-Ruiz M, Mioto PT, González-
GordoS, PalmaJM. 2018. Nitro-oxidative metabolism during fruit ripening.
Journal of Experimental Botany 69, 3449–3463.
CorpasFJ, LeterrierM, ValderramaR, AirakiM, ChakiM, PalmaJM,
BarrosoJB. 2011. Nitric oxide imbalance provokes a nitrosative response
in plants under abiotic stress. Plant Science 181, 604–611.
DelledonneM, XiaY, DixonRA, LambC. 1998. Nitric oxide functions as
a signal in plant disease resistance. Nature 394, 585–588.
Delledonne M, Zeier J, Marocco A, Lamb C. 2001. Signal inter-
actions between nitric oxide and reactive oxygen intermediates in the plant
hypersensitive disease resistance response. Proceedings of the National
Academy of Sciences, USA 98, 13454–13459.
FancyNN, BahlmannAK, LoakeGJ. 2017. Nitric oxide function in plant
abiotic stress. Plant, Cell & Environment 40, 462–472.
FeechanA, KwonE, YunBW, WangYQ, PallasJA, LoakeGJ. 2005.
A central role for S-nitrosothiols in plant disease resistance. Proceedings of
the National Academy of Sciences, USA 102, 8054–8059.
Frungillo L, Skelly MJ, Loake GJ, Spoel SH, Salgado I. 2014.
S-Nitrosothiols regulate nitric oxide production and storage in plants through
the nitrogen assimilation pathway. Nature Communications 5, 5401.
FuZQ, DongX. 2013. Systemic acquired resistance: turning local infection
into global defense. Annual Review of Plant Biology 64, 839–863.
Guerra D, Ballard K, Truebridge I, Vierling E. 2016. S-Nitrosation of
conserved cysteines modulates activity and stability of S-nitrosoglutathione
reductase (GSNOR). Biochemistry 55, 2452–2464.
Harrison SJ, Mott EK, Parsley K, Aspinall S, Gray JC, Cottage A.
2006. A rapid and robust method of identifying transformed Arabidopsis
thaliana seedlings following floral dip transformation. Plant Methods 2, 19.
HeJM, Bai XL, Wang RB, Cao B, She XP. 2007. The involvement of
nitric oxide in ultraviolet-B-inhibited pollen germination and tube growth of
Paulownia tomentosa in vitro. Physiologia Plantarum 131, 273–282.
Kovacs I, Holzmeister C, WirtzM, et al. 2016. ROS-mediated inhib-
ition of S-nitrosoglutathione reductase contributes to the activation of anti-
oxidative mechanisms. Frontiers in Plant Science 7, 1669.
Kubienová L, Kopečný D, Tylichová M, et al. 2013. Structural and
functional characterization of a plant S-nitrosoglutathione reductase from
Solanum lycopersicum. Biochimie 95, 889–902.
KulikA, NoirotE, GrandperretV, BourqueS, FromentinJ, SalloignonP,
TruntzerC, Dobrowolska G, Simon-Plas F, Wendehenne D. 2015.
Interplays between nitric oxide and reactive oxygen species in cryptogein
signalling. Plant, Cell & Environment 38, 331–348.
Kwon E, Feechan A, Yun BW, Hwang BH, Pallas JA, Kang JG,
Loake GJ. 2012. AtGSNOR1 function is required for multiple develop-
mental programs in Arabidopsis. Planta 236, 887–900.
LeeU, WieC, FernandezBO, FeelischM, VierlingE. 2008. Modulation
of nitrosative stress by S-nitrosoglutathione reductase is critical for
thermotolerance and plant growth in Arabidopsis. The Plant Cell 20,
786–802.
LeterrierM, ChakiM, AirakiM, ValderramaR, PalmaJM, BarrosoJB,
Corpas FJ. 2011. Function of S-nitrosoglutathione reductase (GSNOR)
in plant development and under biotic/abiotic stress. Plant Signaling &
Behavior 6, 789–793.
LindermayrC. 2018. Crosstalk between reactive oxygen species and ni-
tric oxide in plants: key role of S-nitrosoglutathione reductase. Free Radical
Biology & Medicine 122, 110–115.
LindermayrC, SellS, MüllerB, LeisterD, DurnerJ. 2010. Redox regu-
lation of the NPR1–TGA1 system of Arabidopsis thaliana by nitric oxide. The
Plant Cell 22, 2894–2907.
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
4886 | Hussain etal.
LiuL, HausladenA, ZengM, QueL, HeitmanJ, StamlerJS. 2001. A
metabolic enzyme for S-nitrosothiol conserved from bacteria to humans.
Nature 410, 490–494.
LoakeG, GrantM. 2007. Salicylic acid in plant defence—the players and
protagonists. Current Opinion in Plant Biology 10, 466–472.
Manjunatha G, Lokesh V, Neelwarne B. 2010. Nitric oxide in fruit
ripening: trends and opportunities. Biotechnology Advances 28, 489–499.
Mudgett MB, Staskawicz BJ. 1999. Characterization of the
Pseudomonas syringae pv. tomato AvrRpt2 protein: demonstration
of secretion and processing during bacterial pathogenesis. Molecular
Microbiology 32, 927–941.
MurashigeT, SkoogF. 1962. A revised medium for rapid growth and bio
assays with tobacco tissue cultures. Physiologia Plantarum 15, 473–497.
PradoAM, ColaçoR, MorenoN, SilvaAC, FeijóJA. 2008. Targeting of
pollen tubes to ovules is dependent on nitric oxide (NO) signaling. Molecular
Plant 1, 703–714.
Prado AM, Porterfield DM, Feijó JA. 2004. Nitric oxide is involved in
growth regulation and re-orientation of pollen tubes. Development 131,
2707–2714.
Rasul S, Dubreuil-Maurizi C, Lamotte O, Koen E, Poinssot B,
AlcarazG, WendehenneD, JeandrozS. 2012. Nitric oxide production
mediates oligogalacturonide-triggered immunity and resistance to Botrytis
cinerea in Arabidopsis thaliana. Plant, Cell & Environment 35, 1483–1499.
Rodríguez-RuizM, Mateos RM, CodesidoV, CorpasFJ, PalmaJM.
2017. Characterization of the galactono-1,4-lactone dehydrogenase from
pepper fruits and its modulation in the ascorbate biosynthesis. Role of nitric
oxide. Redox Biology 12, 171–181.
Rustérucci C, Espunya MC, Díaz M, Chabannes M, Martínez MC.
2007. S-Nitrosoglutathione reductase affords protection against patho-
gens in Arabidopsis, both locally and systemically. Plant Physiology 143,
1282–1292.
SemchukNM, VasylykIuV, KubrakOI, Lushchak VI. 2011. Effect of
sodium nitroprusside and S-nitrosoglutathione on pigment content and
antioxidant system of tocopherol-deficient plants of Arabidopsis thaliana.
Ukrains’kyi Biokhimichnyi Zhurnal (1999) 83, 69–79.
SpadaroD, YunBW, SpoelSH, ChuC, WangYQ, LoakeGJ. 2010. The
redox switch: dynamic regulation of protein function by cysteine modifica-
tions. Physiologia Plantarum 138, 360–371.
TadaY, SpoelSH, Pajerowska-MukhtarK, MouZ, SongJ, WangC,
Zuo J, Dong X. 2008. Plant immunity requires conformational changes
[corrected] of NPR1 via S-nitrosylation and thioredoxins. Science 321,
952–956.
Takahashi H, Shimizu A, Arie T, Rosmalawati S, Fukushima S,
KikuchiM, HikichiY, KandaA, TakahashiA, KibaA. 2005. Catalog of
Micro-Tom tomato responses to common fungal, bacterial, and viral patho-
gens. Journal of General Plant Pathology 71, 8–22.
Tichá T, Sedlářová M, Činčalová L, Trojanová ZD, Mieslerová B,
LebedaA, LuhováL, PetřivalskýM. 2018. Involvement of S-nitrosothiols
modulation by S-nitrosoglutathione reductase in defence responses of let-
tuce and wild Lactuca spp. to biotrophic mildews. Planta 247, 1203–1215.
TorresMA, Dangl JL, Jones JD. 2002. Arabidopsis gp91phox homo-
logues AtrbohD and AtrbohF are required for accumulation of reactive
oxygen intermediates in the plant defense response. Proceedings of the
National Academy of Sciences, USA 99, 517–522.
UknesS, Mauch-ManiB, MoyerM, PotterS, WilliamsS, DincherS,
ChandlerD, SlusarenkoA, WardE, RyalsJ. 1992. Acquired resistance
in Arabidopsis. The Plant Cell 4, 645–656.
WangY, ChenT, ZhangC, HaoH, LiuP, ZhengM, BaluškaF, ŠamajJ,
LinJ. 2009. Nitric oxide modulates the influx of extracellular Ca2+ and actin
filament organization during cell wall construction in Pinus bungeana pollen
tubes. New Phytologist 182, 851–862.
WangYQ, FeechanA, YunBW, etal. 2009. S-Nitrosylation of AtSABP3
antagonizes the expression of plant immunity. Journal of Biological
Chemistry 284, 2131–2137.
XuS, GuerraD, LeeU, VierlingE. 2013. S-Nitrosoglutathione reductases
are low-copy number, cysteine-rich proteins in plants that control multiple
developmental and defense responses in Arabidopsis. Frontiers in Plant
Science 4, 430.
YunBW, FeechanA, YinM, et al. 2011. S-Nitrosylation of NADPH oxi-
dase regulates cell death in plant immunity. Nature 478, 264–268.
Yun BW, Skelly MJ, Yin M, Yu M, Mun BG, Lee SU, Hussain A,
SpoelSH, LoakeGJ. 2016. Nitric oxide and S-nitrosoglutathione function
additively during plant immunity. New Phytologist 211, 516–526.
ZafraA, Rodríguez-GarcíaMI, AlchéJdeD. 2010. Cellular localization
of ROS and NO in olive reproductive tissues during flower development.
BMC Plant Biology 10, 36.
Zandonadi DB, Santos MP, DobbssLB, Olivares FL, Canellas LP,
Binzel ML, Okorokova-Façanha AL, Façanha AR. 2010. Nitric oxide
mediates humic acids-induced root development and plasma membrane
H+-ATPase activation. Planta 231, 1025–1036.
Downloaded from https://academic.oup.com/jxb/article-abstract/70/18/4877/5489421 by guest on 19 March 2020
... Durant cette interaction, un mutant de AtGSNOR1 présente une accumulation de S-nitrosothiol et une sensibilité accrue au pathogène (Feechan et al., 2005). Chez la tomate, une surexpression de la GSNOR (SIGSNOR) entraine une augmentation de la résistance à la bactérie P. syringae (Hussain et al., 2019). En parallèle, d'autres études ont montré une implication différente de la GSNOR dans la défense contre certains pathogènes. ...
Thesis
Le monoxyde d’azote (NO) est une petite molécule gazeuse extrêmement réactive intervenant dans de nombreux processus biologiques. Dans les interactions hôte-pathogènes, le NO peut être produit par les deux partenaires et fait partie de l'arsenal de défense de l'hôte tout comme il fait aussi partie des armes d'attaque du pathogène. Les pathogènes se sont aussi adaptés en mettant en place des systèmes de réponse au NO produit par l'hôte. Dans les interactions symbiotiques plante-microorganismes, du NO a été également détecté. C’est le cas durant la symbiose fixatrice d’azote entre la bactérie Ensifer meliloti et la légumineuse Medicago truncatula. Dans cette interaction il a été montré que non seulement le NO est important lors de l'infection par le symbionte, mais qu'il peut aussi avoir un rôle à des étapes plus tardives comme lors de la fixation d'azote ou de la senescence nodulaire. Dans les nodules, le NO est produit par les deux partenaires avec une contribution d’environ 30% pour la bactérie. Chez la plante, des travaux récents ont montré que la synthèse de NO reposait en grande partie sur des nitrate réductases couplées à la chaine mitochondriale de transfert d'électrons. Coté bactérien, seule la voie de dénitrification était connue comme voie de synthèse possible du NO. Ce travail de thèse pose deux questions centrales : Quelles sont les voies de synthèse du NO chez E. meliloti ? et quel est le rôle du NO produit par les bactéries dans l'interaction symbiotique ? Nous avons montré que E. meliloti ne possède pas de NO synthase capable de produire du NO comme il en existe chez certaines bactéries pathogènes. Seule la voie de dénitrification est responsable de la synthèse du NO chez E. meliloti en vie libre et durant la symbiose. Nous avons aussi montré que la bactérie possède une voie assimilatrice du nitrate fonctionnelle, permettant de contribuer à la production de NO en augmentant la quantité de nitrite disponible pour alimenter la voie de dénitrification. La voie assimilatrice est active aussi bien en aérobie qu'en conditions micro-aérobiques mais ne contribuerait pas à la production de NO dans les nodules symbiotiques. Des résultats préliminaires montrent que cette voie serait régulée par un système à deux composants mais aussi, de façon originale, par un ARN non codant. Enfin nous avons montré que le NO produit par la bactérie ne présente pas de rôle essentiel, ni durant les étapes précoces, ni dans les étapes plus tardives de la symbiose. L'ensemble des résultats nous permet de proposer un modèle des voies de production de NO chez E. meliloti et suggère que dans les interactions symbiotiques le NO bactérien ne joue pas un rôle aussi déterminant que dans les interactions hôtes-pathogènes.
... Nitric oxide (NO) is a signaling molecule distributing throughout all living organisms, and it is involved in multiple plant processes, including growth, development, and biotic and abiotic stress responses [9,10]. The accumulating data indicate that NO is executed through S-nitrosylation, which is the addition of an NO moiety to a protein cysteine thiol to form an S-nitrosothiol (SNO) [10]. ...
Article
Full-text available
Background Basic leucine zipper (bZIP) transcription factors are crucial in plant development, and response to environmental stress, etc. With the development of sequencing technology and bioinformatics analysis, the bZIP family genes has been screened and identified in many plant species, but bZIP family genes has not been systematically characterized and identified their function in Betula platyphylla . Methods B. platyphylla reference genome was used to characterize bZIP family genes. The physicochemical properties, chromosome distribution, gene structure, and syntenic relationships were analyzed by bioinformatics methods. The effect of BpbZIP26 on triterpenoid production was investigated using Agrobacterium -mediated transient transformation under N6022 treatment. Results 51 bZIP family genes were identified in B. platyphylla , and named BpbZIP1 – BpbZIP51 sequentially according to their positions on chromosomes. All BpbZIP genes were unevenly distributed on 14 chromosomes, and divided into 13 subgroups according to the classification of Arabidopsis thaliana bZIP proteins. 12 duplication events were detected in the B. platyphylla genome, and 28 orthologs existed between B. platyphylla and A. thaliana, 83 orthologs existed between B. platyphylla and Glycine max , and 73 orthologs existed between B. platyphylla and Populus trichocarpa . N6022 treatment changed gene expression levels of most BpbZIPs in seedlings of B. platyphylla . Among of them, N6022 treatment significantly enhanced gene expression levels of BpbZIP26 in leaves, stems and roots of B. platyphylla . BpbZIP26 mediated triterpenoid production, and N6022 treatment further enhanced triterpenoid production in BpbZIP26 overexpression calli of B. platyphylla using Agrobacterium -mediated transient transformation. Conclusion This work highlights potential BpbZIP family genes responding to S-nitrosothiol and provides candidate genes for triterpenoid production in B. platyphylla . Graphical Abstract
... It was found that the disease resistance of OE fruit was higher than that of WT and RNAi lines (Fig. 6A-C), which validates the results that the high activity and gene expression of GSNOR were corresponding to the strong disease resistance of the fruit (Fig. 5). The present result is consistent with the finding that overexpression of GSNOR could significantly improve the resistance of tomato to Pseudomonas syringae (Hussain et al., 2019). Early results also reported that GSNOR actively improves immunity of plants to Blumeria graminis and Hyaloperonospora parasitica (Feechan et al., 2005). ...
Article
Gamma-aminobutyric acid (GABA), a widely distributed metabolite in prokaryotes and eukaryotes, has many functions for plants in stress responses. In this study, hypotonic treatment with 10 mmol L⁻¹ GABA in cherry tomato induced resistance to Botrytis cinerea with markedly lower disease incidence and lesion diameter, led to endogenous nitric oxide (NO) tansient accumulation before inoculation the pathogen then decrease after inoculation, and enhanced the content of arginine (Arg) and glutamic acid (Glu). The resistance of fruit treated with a NO scavenger, carboxy-2-phenyl-4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide (cPTIO), was significantly reduced. Moreover, the enzyme activity and gene expression of S-nitrosoglutathione reductase (GSNOR) were enhanced following endogenous NO increased. The endogenous NO level was excessively high after treatment with a GSNOR scavenger, N6022, making the fruit more susceptible to pathogen. Similarly, after break down of SlGSNOR, fruit had much higher endogenous NO and lower disease resistance. However, overexpression of SlGSNOR exhibited opposite consequences. These results suggest that a suitable level of NO is beneficial for enhancing disease resistance, and GABA can help tomatoes maintain NO equilibrium by regulating GSNOR.
... In the context of metabolism, Arabidopsis GSNOR, via its ability to regulate Snitrosylation, has been shown to control the biosynthesis of the immune activator, SA (Feechan et al., 2005). Similar phenotypes to those of Arabidopsis have also been described in tomato GSNOR RNAi lines (Hussain et al., 2019), suggesting that the function of this enzyme is conserved across dicotyledonous species. ...
Article
Nitric oxide (NO) has emerged as an important signal molecule in plants, having myriad roles in plant development. In addition, NO also orchestrates both biotic and abiotic stress responses, during which intensive cellular metabolic reprogramming occurs. Integral to these response is the location of NO biosynthetic and scavenging pathways in diverse cellular compartments, enabling plants to effectively organize signal transduction pathways. NO regulates plant metabolism and in turn, metabolic pathways reciprocally regulate NO accumulation and function. Thus, these diverse cellular processes are inextricably linked. This review addresses the numerous redox pathways, located in the various subcellular compartments, which produce NO, in addition to the mechanisms underpinning NO scavenging. We focus on how this molecular dance is integrated into the metabolic state of the cell. Within this context, a reciprocal relationship between NO accumulation and metabolite production is often apparent. We additionally showcase cellular pathways including those associated with nitrate reduction that provide evidence for this integration of NO function and metabolism. Finally, we discuss the potential importance of the biochemical reactions governing NO levels in determining plant responses to a changing environment.
Article
Following pathogen recognition, plant cells produce a nitrosative burst resulting in a striking increase in nitric oxide (NO), altering the redox state of the cell, which subsequently helps orchestrate a plethora of immune responses. NO is a potent redox cue, efficiently relayed between proteins through its co-valent attachment to highly specific, powerfully reactive protein cysteine (Cys) thiols, resulting in formation of protein S-nitrosothiols (SNOs). This process, known as S-nitrosylation, can modulate the function of target proteins, enabling responsiveness to cellular redox changes. Key targets of S-nitrosylation control the production of reactive oxygen species (ROS), the transcription of immune-response genes, the triggering of the hypersensitive response (HR) and the establishment of systemic acquired resistance (SAR). Here, we bring together recent advances in the control of plant immunity by S-nitrosylation, furthering our appreciation of how changes in cellular redox status reprogramme plant immune function.
Chapter
Nitrogen (N 2 ) is an essential macronutrient and plays an essential role in maintaining growth and yield of plants. This being the case, it is imperative to recognize the key traits involved in improving nitrogen use efficiency (NUE), leading to sustainable agriculture development. Owing to their immobile nature, plants growing in adverse environmental conditions face several stresses. As a result, they have evolved several types of adaptive or responsive strategies to cope with ecological conditions. This chapter elaborates on physiological modifications in plants that regulate N 2 uptake and utilization efficiency to enhance plant growth. Additionally, we discuss the molecular basis of NUE to increase plant growth and the contribution of N 2 in establishing sustainable agriculture to fulfill global food issues. Nitric oxide (NO) is a signaling molecule and regulates several developmental events in plants. The levels of NO can be influenced directly by the availability and assimilation of N, thereby altering the major cellular pathways. For sustainable agriculture generating nitrogen‐efficient crops and subsequently exploiting NO could be promising strategies.
Article
Nitric oxide (NO), as a redox molecule, played important role in plant response to environmental stress. Here, we found that loss of tomato S-Nitrosoglutathione reductase (SlGSNOR), a critical regulator of NO balance, led to global sensitivity to heat, salt, bicarbonate, and paraquat stresses in tomato, suggesting that SlGSNOR was a positive regulator in tomato against abiotic stresses. In addition, under these stresses, loss of SlGSNOR induced excessive NO accumulation in the roots of the transgenic plant. Bicarbonate accumulation has been a limiting factor of tomato production in the north of China, and there were less reports on the regulation of bicarbonate stress compared with other abiotic stress, therefore, in the following experiment, the mechanism of SlGSNOR function focused on investigating bicarbonate tolerance. Proteomes data indicated that loss of SlGSNOR triggered the expression of proteins in MAPK and ethylene signaling pathway under bicarbonate condition. Importantly, loss of SlGSNOR increased ethylene emission under bicarbonate condition, meaning that ethylene signaling participated in SlGSNOR-mediated tomato bicarbonate stress tolerance. Interestingly, under bicarbonate stress, silencing SlMAPK3 or SlACO1 in SlGSNOR-RNAi tomato partially compromised the sensitivity and ethylene emission. Mechanistically, under bicarbonate stress, silencing SlMAPK3 in SlGSNOR-RNAi tomato plants suppressed the transcription of SlACO1, whereas silencing SlACO1 in SlGSNOR-RNAi tomato activated the transcription of SlMAPK3. Collectively, these data demonstrated that SlGSNOR-mediated tomato abiotic stress tolerance was depended on SlMAPK3-SlACO1 cascade signaling, largely. In addition, the incorporation of global abiotic stress tolerance traits into crop plants may succeed by manipulating GSNOR expression.
Article
Full-text available
The liaison between Nitric oxide (NO) and phytohormones regulates a myriad of physiological processes at the cellular level. The interaction between NO and phytohormones is mainly influenced by NO-mediated post-translational modifications (PTMs) under basal as well as induced conditions. Protein S-nitrosylation is the most prominent and widely studied PTM among others. It is the selective but reversible redox-based covalent addition of a NO moiety to the sulfhydryl group of cysteine (Cys) molecule(s) on a target protein to form S-nitrosothiols. This process may involve either direct S-nitrosylation or indirect S-nitrosylation followed by transfer of NO group from one thiol to another (transnitrosylation). During S-nitrosylation, NO can directly target Cys residue (s) of key genes involved in hormone signaling thereby regulating their function. The phytohormones regulated by NO in this manner includes abscisic acid, auxin, gibberellic acid, cytokinin, ethylene, salicylic acid, jasmonic acid, brassinosteroid, and strigolactone during various metabolic and physiological conditions and environmental stress responses. S-nitrosylation of key proteins involved in the phytohormonal network occurs during their synthesis, degradation, or signaling roles depending upon the response required to maintain cellular homeostasis. This review presents the interaction between NO and phytohormones and the role of the canonical NO-mediated post-translational modification particularly, S-nitrosylation of key proteins involved in the phytohormonal networks under biotic and abiotic stresses.
Article
Full-text available
In plants senescence is the final stage of plant growth and development that ultimately leads to death. Plants experience age-related as well as stress-induced developmental ageing. Senescence involves significant changes at the transcriptional, post-translational and metabolomic levels. Furthermore, phytohormones also play a critical role in the programmed senescence of plants. Nitric oxide (NO) is a gaseous signalling molecule that regulates a plethora of physiological processes in plants. Its role in the control of ageing and senescence has just started to be elucidated. Here, we review the role of NO in the regulation of programmed cell death, seed ageing, fruit ripening and senescence. We also discuss the role of NO in the modulation of phytohormones during senescence and the significance of NO-ROS cross-talk during programmed cell death and senescence.
Article
Full-text available
The liaison between Nitric oxide (NO) and phytohormones regulates a myriad of physiological processes at the cellular level. The interaction between NO and phytohormones is mainly influenced by NO-mediated post-translational modifications (PTMs) under basal as well as induced conditions. Protein S-nitrosylation is the most prominent and widely studied PTM among others. It is the selective but reversible redox-based covalent addition of a NO moiety to the sulfhydryl group of cysteine (Cys) molecule(s) on a target protein to form S-nitrosothiols. This process may involve either direct S-nitrosylation or indirect S-nitrosylation followed by transfer of NO group from one thiol to another (transnitrosylation). During S-nitrosylation, NO can directly target Cys residue (s) of key genes involved in hormone signaling thereby regulating their function. The phytohormones regulated by NO in this manner includes abscisic acid, auxin, gibberellic acid, cytokinin, ethylene, salicylic acid, jasmonic acid, brassinosteroid, and strigolactone during various metabolic and physiological conditions and environmental stress responses. S-nitrosylation of key proteins involved in the phytohormonal network occurs during their synthesis, degradation, or signaling roles depending upon the response required to maintain cellular homeostasis. This review presents the interaction between NO and phytohormones and the role of the canonical NO-mediated post-translational modification particularly, S-nitrosylation of key proteins involved in the phytohormonal networks under biotic and abiotic stresses.
Article
Full-text available
Main conclusion Resistant Lactuca spp. genotypes can efficiently modulate levels of S-nitrosothiols as reactive nitrogen species derived from nitric oxide in their defence mechanism against invading biotrophic pathogens including lettuce downy mildew. Abstract S-Nitrosylation belongs to principal signalling pathways of nitric oxide in plant development and stress responses. Protein S-nitrosylation is regulated by S-nitrosoglutathione reductase (GSNOR) as a key catabolic enzyme of S-nitrosoglutathione (GSNO), the major intracellular S-nitrosothiol. GSNOR expression, level and activity were studied in leaves of selected genotypes of lettuce (Lactuca sativa) and wild Lactuca spp. during interactions with biotrophic mildews, Bremia lactucae (lettuce downy mildew), Golovinomyces cichoracearum (lettuce powdery mildew) and non-pathogen Pseudoidium neolycopersici (tomato powdery mildew) during 168 h post inoculation (hpi). GSNOR expression was increased in all genotypes both in the early phase at 6 hpi and later phase at 72 hpi, with a high increase observed in L. sativa UCDM2 responses to all three pathogens. GSNOR protein also showed two-phase increase, with highest changes in L. virosa–B. lactucae and L. sativa cv. UCDM2–G. cichoracearum pathosystems, whereas P. neolycopersici induced GSNOR protein at 72 hpi in all genotypes. Similarly, a general pattern of modulated GSNOR activities in response to biotrophic mildews involves a two-phase increase at 6 and 72 hpi. Lettuce downy mildew infection caused GSNOR activity slightly increased only in resistant L. saligna and L. virosa genotypes; however, all genotypes showed increased GSNOR activity both at 6 and 72 hpi by lettuce powdery mildew. We observed GSNOR-mediated decrease of S-nitrosothiols as a general feature of Lactuca spp. response to mildew infection, which was also confirmed by immunohistochemical detection of GSNOR and GSNO in infected plant tissues. Our results demonstrate that GSNOR is differentially modulated in interactions of susceptible and resistant Lactuca spp. genotypes with fungal mildews and uncover the role of S-nitrosylation in molecular mechanisms of plant responses to biotrophic pathogens.
Article
Full-text available
Pepper fruit is one of the highest vitamin C sources of plant origin for our diet. In plants, ascorbic acid is mainly synthesized through the L-galactose pathway, being the L-galactono-1,4-lactone dehydrogenase (GalLDH) the last step. Using pepper fruits, the full GalLDH gene was cloned and the protein molecular characterization accomplished. GalLDH protein sequence (586 residues) showed a 37 amino acids signal peptide at the N-terminus, characteristic of mitochondria. The hydrophobic analysis of the mature protein displayed one transmembrane helix comprising 20 amino acids at the N-terminus. By using a polyclonal antibody raised against a GalLDH internal sequence and immunoblotting analysis, a 56 kDa polypeptide cross-reacted with pepper fruit samples. Using leaves, flowers, stems and fruits, the expression of GalLDH by qRT-PCR and the enzyme activity were analyzed, and results indicate that GalLDH is a key player in the physiology of pepper plants, being possibly involved in the processes which undertake the transport of ascorbate among different organs. We also report that an NO (nitric oxide)-enriched atmosphere enhanced ascorbate content in pepper fruits about 40% parallel to increased GalLDH gene expression and enzyme activity. This is the first report on the stimulating effect of NO treatment on the vitamin C concentration in plants. Accordingly, the modulation by NO of GalLDH was addressed. In vitro enzymatic assays of GalLDH were performed in the presence of SIN-1 (peroxynitrite donor) and S-nitrosoglutahione (NO donor). Combined results of in vivo NO treatment and in vitro assays showed that NO provoked the regulation of GalLDH at transcriptional and post-transcriptional levels, but not post-translational modifications through nitration or S-nitrosylation events promoted by reactive nitrogen species (RNS) took place. These results suggest that this modulation point of the ascorbate biosynthesis could be potentially used for biotechnological purposes to increase the vitamin C levels in pepper fruits.
Article
Full-text available
Nitric oxide (NO) has emerged as a signaling molecule in plants being involved in diverse physiological processes like germination, root growth, stomata closing and response to biotic and abiotic stress. S-nitrosoglutathione (GSNO) as a biological NO donor has a very important function in NO signaling since it can transfer its NO moiety to other proteins (trans-nitrosylation). Such trans-nitrosylation reactions are equilibrium reactions and depend on GSNO level. The breakdown of GSNO and thus the level of S-nitrosylated proteins are regulated by GSNO-reductase (GSNOR). In this way, this enzyme controls S-nitrosothiol levels and regulates NO signaling. Here we report that Arabidopsis thaliana GSNOR activity is reversibly inhibited by H2O2 in vitro and by paraquat-induced oxidative stress in vivo. Light scattering analyses of reduced and oxidized recombinant GSNOR demonstrated that GSNOR proteins form dimers under both reducing and oxidizing conditions. Moreover, mass spectrometric analyses revealed that H2O2-treatment increased the amount of oxidative modifications on Zn2+-coordinating Cys47 and Cys177. Inhibition of GSNOR results in enhanced levels of S-nitrosothiols followed by accumulation of glutathione. Moreover, transcript levels of redox-regulated genes and activities of glutathione-dependent enzymes are increased in gsnor-ko plants, which may contribute to the enhanced resistance against oxidative stress. In sum, our results demonstrate that reactive oxygen species (ROS)-dependent inhibition of GSNOR is playing an important role in activation of anti-oxidative mechanisms to damping oxidative damage and imply a direct crosstalk between ROS- and NO-signaling.
Article
Nitric oxide (NO) has emerged as an essential biological messenger in plant biology that usually transmits its bioactivity by post-translational modifications such as S-nitrosylation, the reversible addition of a NO group to a protein cysteine residue leading to S-nitrosothiols (SNOs). In last year's, SNOs have emerged as key signalling molecules mainly involved in plant response to stress. Chief among SNOs is S-nitrosoglutathione (GSNO), generated by S-nitrosylation of the key antioxidant glutathione (GSH). GSNO is considered the major NO reservoir and a phloem mobile signal that confers to NO the capacity of a long-distance signalling molecule. GSNO is able to regulate protein function and gene expression resulting in a key role of GSNO in fundamental processes in plant such as development and response to a wide range of environmental stresses. In addition, GSNO is also able to regulate total SNO pool and consequently, it could be considered the storage of NO in cells that may control NO signalling under basal and stress-related responses. Thus, GSNO function could be crucial during plant response to environmental stresses. Besides this importance of GSNO in plant biology, its mode of action has not been widely discussed in the literature. In this review, we will firstly discuss the GSNO turnover in cells and secondly the role of GSNO as mediator of physiological and stress-related processes in plants, highlighting aspects in which there is still some controversy.
Article
Pepper (Capsicum annuum L.) and tomato (Solanum lycopersicum L.), which belong to the Solanaceae family, are among the most cultivated and consumed fleshy fruits worldwide and constitute excellent sources of many essential nutrients, such as vitamins A, C, and E, calcium, and carotenoids. While fruit ripening is a highly regulated and complex process, tomato and pepper have been classified as climacteric and non-climacteric fruits, respectively. These fruits differ greatly in shape, color composition, flavor, and several other features which undergo drastic changes during the ripening process. Such ripening-related metabolic and developmental changes require extensive alterations in many cellular and biochemical processes, which ultimately leads to fully ripe fruits with nutritional and organoleptic features that are attractive to both natural dispersers and human consumers. Recent data show that reactive oxygen and nitrogen species (ROS/RNS) are involved in fruit ripening, during which molecules, such as hydrogen peroxide (H2O2), NADPH, nitric oxide (NO), peroxynitrite (ONOO-), and S-nitrosothiols (SNOs), interact to regulate protein functions through post-translational modifications. In light of these recent discoveries, this review provides an update on the nitro-oxidative metabolism during the ripening of two of the most economically important fruits, discusses the signaling roles played by ROS/RNS in controlling this complex physiological process, and highlights the potential biotechnological applications of these substances to promote further improvements in fruit ripening regulation and nutritional quality. In addition, we suggest that the term 'nitro-oxidative eustress' with regard to fruit ripening would be more appropriate than nitro-oxidative stress, which ultimately favors the consolidation of the plant species.
Article
Nitric oxide (.NO) acts as signaling molecule in plants being involved in diverse physiological processes such as germination, root growth, stomata closing and response to biotic and abiotic stress. S-Nitrosoglutathione (GSNO) is the storage and transport form of.NO and has a very important function in.NO signaling since it can transfer its.NO moiety to other proteins (trans-nitrosylation). The level of GSNO and thus the level of S-nitrosylated proteins are regulated by GSNO-reductase (GSNOR). In this way, this enzyme regulates the S-nitrosothiol levels and plays a balancing role in fine-tuning.NO signaling. Interestingly, oxidative post-translationally modification of GSNOR inhibited the activity of this enzyme suggesting a direct crosstalk between ROS- and RNS-signaling. In this review article the regulatory effects of ROS on GSNOR are highlighted and their physiological function in context of crosstalk between ROS and.NO species in plants are discussed.
Article
The free radical nitric oxide (NO●) regulates diverse physiological processes from vasodilation in humans to gas exchange in plants. S-nitrosoglutathione (GSNO) is considered a principal nitroso reservoir due to its chemical stability. GSNO accumulation is attenuated by GSNO reductase (GSNOR), a cysteine-rich cytosolic enzyme. Regulation of protein nitrosation is not well understood since NO●-dependent events proceed without discernible changes in GSNOR expression. Because GSNORs contain evolutionarily-conserved cysteines that could serve as nitrosation sites, we examined the effects of treating plant (Arabidopsis thaliana), mammalian (human), and yeast (Saccharomyces cerevisiae) GSNORs with nitrosating agents in vitro. Enzyme activity was sensitive to nitroso donors, while the reducing agent dithiothreitol (DTT) restored activity, suggesting catalytic impairment was due to S-nitrosation. Protein nitrosation was confirmed by mass spectrometry, by which mono-, di-, and tri-nitrosation were observed, and these signals were sensitive to DTT. GSNOR mutants in specific non-zinc coordinating cysteines were less sensitive to catalytic inhibition by nitroso donors and exhibited reduced nitrosation signals by mass spectrometry. Nitrosation also coincided with decreased tryptophan fluorescence, increased thermal aggregation propensity, and increased polydispersity-properties reflected by differential solvent accessibility of amino acids important for dimerization and the shape of the substrate and coenzyme binding pockets as assessed by hydrogen-deuterium exchange mass spectrometry. Collectively, these data suggest a mechanism for NO● signal transduction in which GSNOR nitrosation and inhibition transiently permit GSNO accumulation.
Article
Nitric oxide (NO) is emerging as a key regulator of diverse plant cellular processes. A major route for the transfer of NO bioactivity is S-nitrosylation, the addition of an NO moiety to a protein cysteine thiol forming an S-nitrosothiol (SNO). Total cellular levels of protein S-nitrosylation are controlled predominantly by S-nitrosoglutathione reductase 1 (GSNOR1) which turns over the natural NO donor, S-nitrosoglutathione (GSNO). In the absence of GSNOR1 function, GSNO accumulates, leading to dysregulation of total cellular S-nitrosylation. Here we show that endogenous NO accumulation in Arabidopsis, resulting from loss-of-function mutations in NO Overexpression 1 (NOX1), led to disabled Resistance (R) gene-mediated protection, basal resistance and defence against nonadapted pathogens. In nox1 plants both salicylic acid (SA) synthesis and signalling were suppressed, reducing SA-dependent defence gene expression. Significantly, expression of a GSNOR1 transgene complemented the SNO-dependent phenotypes of paraquat resistant 2-1 (par2-1) plants but not the NO-related characters of the nox1-1 line. Furthermore, atgsnor1-3 nox1-1 double mutants supported greater bacterial titres than either of the corresponding single mutants. Our findings imply that GSNO and NO, two pivotal redox signalling molecules, exhibit additive functions and, by extension, may have distinct or overlapping molecular targets during both immunity and development.
Article
Abiotic stress is one of the main threats affecting crop growth and production. An understanding of the molecular mechanisms that underpin plant responses against environmental insults will be crucial to help guide the rational design of crop plants to counter these challenges. A key feature during abiotic stress is the production of nitric oxide (NO), an important concentration dependent, redox-related signalling molecule. NO can directly or indirectly interact with a wide range of targets leading to the modulation of protein function and the reprogramming of gene expression. The transfer of NO bioactivity can occur through a variety of potential mechanisms but chief among these is S-nitrosylation, a prototypic, redox-based, post-translational modification. However, little is known about this pivotal molecular amendment in the regulation of abiotic stress signalling. Here, we describe the emerging knowledge concerning the function of NO and S-nitrosylation during plant responses to abiotic stress.