Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Plant Science 287 (2019) 110184

Contents lists available at ScienceDirect

Plant Science
journal homepage: www.elsevier.com/locate/plantsci

The β-amylase PbrBAM3 from pear (Pyrus betulaefolia) regulates soluble T


sugar accumulation and ROS homeostasis in response to cold stress
Liangyi Zhaoa,1, Tianyuan Yanga,b,1, Caihua Xinga,1, Huizheng Donga, Kaijie Qia, Junzhi Gaoa,
⁎ ⁎
Shutian Taoa, Juyou Wua, Jun Wua, Shaoling Zhanga, , Xiaosan Huanga,
a
College of Horticulture, State Key Laboratory of Crop Genetics and Germplasm Enhancement, Nanjing Agricultural University, Nanjing 210095, China
b
State Key Laboratory of Tea Plant Biology and Utilization, Anhui Agricultural University, Hefei 230036, China

A R T I C LE I N FO A B S T R A C T

Keywords: β-Amylase (BAM) is involved in sugar metabolism, but the role of BAM genes in cold tolerance remains poorly
Cold stress understood. Here, we report the identification and functional characterization of the chloroplast-localized BAM-
β-amylase (BAM) encoding gene PbrBAM3 isolated from Pyrus betulaefolia. The transcript levels of PbrBAM3 were up-regulated
Osmolyte under cold, dehydration and ABA, but repressed by maltose. Overexpression of PbrBAM3 in tobacco (Nicotiana
Starch degradation
tabacum) and pear (P. ussuriensis) conferred increased BAM activity, promoted starch degradation after chilling
Antioxidant enzyme
treatments and enhanced tolerance to cold. Under the chilling stress, the transgenic tobacco and P. ussuriensis
Reactive oxygen species (ROS)
exhibited lessened reactive oxygen species (ROS) generation, higher levels of antioxidant enzymes activity, and
greater accumulation of soluble sugars (specially maltose) than the corresponding wild type plants. Taken to-
gether, these results demonstrate that PbrBAM3 plays an important role in cold tolerance, at least in part, by
raising the levels of soluble sugars capable of acting as osmolytes or antioxidants.

1. Introduction scientific interest.


Plants cope with cold stress by regulating a large number of genes at
As sessile organisms, plants are inevitably exposed to environmental a transcriptional level [31]. These cold-responsive genes can be clas-
stresses. Of these, cold stress is one of the most devastating abiotic sified according to their response timing into early-response genes and
stresses: it impairs plant growth and development, reduces productivity delayed-response genes. The former group codes for regulatory pro-
and limits geographical distribution. Over a long evolutionary period, teins, such as transcription factors (TFs) and protein kinases, which
plants have evolved a set of sophisticated mechanisms to cope with and function in stress signal transduction and gene expression regulation.
adapt to cold stress. During the last decades, significant advancements The latter group consists of genes whose products protect plant cells
have been made in revealing the mechanisms underlying plant defen- against damage derived from stresses and sustain cell viability [10–12].
sive responses to abiotic stresses [1–6]. Such defensive responses of Under cold stress, ROS accumulate due to a disrupted balance be-
plants are considerably complex processes regulated through multiple tween ROS production and ROS scavenging by scavenging enzymes and
signaling pathways. A general stress signal transduction pathway starts non-enzymatic antioxidants [13,14]. They oxidize lipids, proteins and
with a signal perception, transduction to secondary messengers and nucleic acids, which in turn activates further stress (possibly leading to
signal amplification (Ca2+, reactive oxygen species (ROS)), which in cell membrane damage and osmotic potential unbalance [15–17], but
turn leads to an array of biochemical and physiological modifications also counter-acting defense pathways. Contributing to the latter,
allowing the plant to tolerate the stress [7–9]. In comparison to tradi- modulated metabolism-related genes lead to the production and accu-
tional breeding, genetic engineering has been shown as a more effective mulation of “compatible solutes”, i.e., osmolytes or osmoprotectants
and time-saving approach to generate new germplasms with improved such as proline and soluble sugars. It is becoming increasingly clear that
stress tolerance [17]. With this aim in mind, the exploitation of a large soluble sugars play a key role in ROS scavenging and mediating the cold
reservoir of genes with desirable functions in cold tolerance is of utmost tolerance [18–20].


Corresponding authors.
E-mail addresses: 2016104019@njau.edu.cn (L. Zhao), yangtianyuan2008@163.com (T. Yang), 2015104018@njau.edu.cn (C. Xing),
2017104017@njau.edu.cn (H. Dong), qikaijie@njau.edu.cn (K. Qi), 2017804152@njau.edu.cn (J. Gao), toast@njau.edu.cn (S. Tao), juyouwu@njau.edu.cn (J. Wu),
wujun@njau.edu.cn (J. Wu), slzhang@njau.edu.cn (S. Zhang), huangxs@njau.edu.cn (X. Huang).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.plantsci.2019.110184
Received 29 December 2018; Received in revised form 3 July 2019; Accepted 6 July 2019
Available online 09 July 2019
0168-9452/ © 2019 Elsevier B.V. All rights reserved.
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 1. Phylogenetic analysis of β-amylases from five different plants. A phylogenetic tree was constructed based on the deduced amino acid sequences of β-amylases
in Arabidopsis thaliana, Malus x domestica, Populus trichocarpa, Poncirus trifoliata, Oryza sativa and Pyrus betulaefolia.

A large number of genes encode enzymes involved in the synthesis ussuriensis confers cold tolerance, concurrent with an increased BAM
or degradation of soluble sugars in plants [21]. β-amylase (BAM) is an activity and higher accumulation of maltose and other soluble sugars.
exohydrolase hydrolyzing α-1,4-linked glucan chains and functions in Overall, our data suggest that PbrBAM3 enhances cold tolerance, at
converting starch into maltose [22]. The maltose can be exported from least in part, by enhancing sugar metabolism.
chloroplast into cytosol, and then it is converted into glucose through
disproportionating glucosyltransferases [23–26]. BAM is the mediator
of starch degradation into the downstream sugars and its role in cold 2. Materials and methods
stress has been characterized in a range of plant species. For example,
cold stress increased the BAM activity of potato more than 4-fold [27]. 2.1. Plant material, growth conditions and stress treatments
Nine BAM homologues have been identified in the Arabidopsis genome
[25]. Knock down of AtBAM3 by RNA inference (RNAi) led to the de- Pyrus betulaefolia seedlings were grown in the experimental base of
crease of soluble sugars levels [28] and silencing of AtBAM3 resulted in the National Center of Pear Engineering Technology Research, Nanjing
starch accumulation [29]. In contrast, silencing of AtBAM1 did not Agricultural University. To examine PbrBAM3 expression pattern, uni-
result in starch accumulation [30], and AtBAM4 did not code for an form and healthy shoots of Pyrus betulaefolia from grown 45-day-old
active BAM enzyme [26]. Two other BAM genes, AtBAM7 and AtBAM8, seedlings were subjected to various stress treatments. For cold treat-
have also been associated with plant growth and development, via a ment, the seedlings were kept in the growth chamber at 4 °C for 0, 6, 24,
crosstalk with brassinosteroid signaling [22], a pathway different from 72 and 144 h. For dehydration treatment, the seedlings were placed on
that of AtBAM3. Collectively, these findings suggest that AtBAM3 is the dry filter paper above a work bench and dried at 25 °C (in autumn, with
dominant hydrolase involved in starch breakdown. relative humidity of 60–70 %) for 0, 0.5, 1, 3, and 6 h. For the ABA
Having originated in Northern China, Pyrus betulaefolia is extremely treatment was performed based on a previous method [33,34]. The
cold-hardy [32], making it an ideal source for genes of agronomical seedlings were incubated with their roots dipped in solution containing
importance with potential use for genetic engineering of cold tolerance. 100 μM ABA for 0, 1, 3, 6, 12, 24, and 72 h. In addition, maltose
Although the BAMs of Arabidopsis have been characterized, the roles of treatment was performed by incubating the seedlings in 200 mM mal-
BAM3 orthologs in stress tolerance remain poorly characterized in non- tose solution for 0, 6, 12, 24, 48 and 72 h. For each treatment, at least
model plants, especially in perennial fruit trees. In this study, we report 40 seedlings shoots were harvested at the indicated time points, im-
the isolation and functional characterization of PbrBAM3 from Pyrus mediately frozen in liquid nitrogen and stored at -80 °C until further
betulaefolia. Transgenic overexpression of PbrBAM3 in tobacco and P. use.

2
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 2. Alignment of the core glucosyl hydro-


lase domains of PbrBAM3 (line 3) and its nine
Arabidopsis orthologs, AtBAM1–9. Identical
and similar amino acid residues among the
aligned sequences are indicated by black and
grey shading, respectively. Substrate binding
sites and two catalytic sites (Glu186 and
Glu380) are marked with large black arrow-
heads. The residues forming flexible loop and
inner loop are indicated with solid lines above
the sequences.

2.2. Quantitative real-time RT-PCR analysis of PbrBAM3 expression 2.4. Subcellar localization of PbrBAM3

Total RNA was extracted from the shoot samples using TRIZOL re- The complete ORF of PbrBAM3 was amplified by RT-PCR using
agent (Invitrogen, United States). Approximately 1 μg of total RNA was primers (GSP3; Table S1), then cloned into the pCAMBIA1302 vector
reversely transcribed into cDNA using the PrimeScript™ RT reagent kit and fused in-frame to the N-terminal of GFP (Green Fluorescent
(Takara, China) according to manufacturer’s instructions. QRT-PCR was Protein), under the control of CaMV 35S promoter. After validation by
performed on a BioRad CFX96 real-time system using an SYBR® Green sequencing, the fusion constructs 35S pro:PbrBAM3::GFP or 35S
Master Mix kit (Takara) according to the manufacturer’s instructions as pro:GFP (as a control) were used for transient expression in A. thaliana
described previously [35]. The primers used are listed in Supplemen- protoplast cells (about 1 μg plasmids of each) according to the poly-
tary Table S1. Each sample was analyzed in four replicates, and the ethyleneglycol method [35,39]. For the subcellular localization of
2−ΔΔCt method [36] was applied to calculate relative expression levels PbrBAM3 protein in A. thaliana protoplast cells, we used a confocal
of each gene. Expression of Ubiquitin (DQ830978) was used as an in- laser scanning microscope (LSM410; Carl Zeiss) to observe the green
ternal reference for tobacco, while Tubulin (AB239681) was used for GFP and RFP fluorescence (at an excitation (Ex) wavelength of 488 nm
Pyrus betulaefolia. The expression analysis at each time point was re- and emission (Em) wavelength of 525 nm) and the red autofluorescence
peated at least three times, and the data are shown as the mean va- of chloroplasts (at Ex = 488 and Em = 587–610 nm), as previously
lues ± SE. described [33].

2.5. Generation and identification of transgenic plants


2.3. Isolation and sequence analysis of PbrBAM3
To generate transgenic tobacco plants overexpressing PbrBAM3, the
Based on the sequence a pair of primers was designed for RT-PCR to whole ORF of PbrBAM3 was PCR-amplified with special primers GSP4
amplify the ORF of BAM3 gene in Pyrus betulaefolia. RT-PCR assays (Table S1) and inserted into binary vector pCMABIA1301 linearized
were performed as described earlier [37]. The multiple alignments of with BamH I and Xba I under the control of the CaMV 35S promoter.
the deduced amino acid sequence were performed using the ClustalW The overexpression vectors were mobilized into Agrobacterium tumefa-
program. A phylogenetic tree of PbrBAM3 and other plant species was ciens strain EHA105 and then used to transform tobacco or P. ussuriensis
constructed by the Maximum Likelihood (ML) method using MEGA according to previous reports [40,41]. The overexpression vectors were
(version 6.0) software [38]. The analysis of PbrBAM3 theoretical iso- mobilized into a positive transformants and identified by genomic PCR
electric point and molecular weight was performed using the online using primers specific for sequences of the hygromycin gene and
website ExPASy (http://www.align.genome.jp/). In addition, multiple PbrBAM3 in tobacco or P. ussuriensis. The PbrBAM3 mRNA levels were
alignments of the BAM sequences containing the core glucosyl hydro- examined by RT-PCR, as described [33], using Tubulin and Ubiquitin as
lase domains were identified by ClustalX program with default settings internal controls for P. ussuriensis and tobacco, respectively. T2 homo-
and presented by Jalview (http://www.jalview.org/). zygous plants of tobacco and propagated plants of P. ussuriensis [40,41]
were used for the following experiments.

3
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 3. Expression of PbrBAM3 in Pyrus betulaefolia in response to various treatments. Time-courses of expression levels in the shoot were analyzed by qPCR in
response to the indicated stresses. Error bars are S.E. Note the different durations of the stress treatments (Materials and Methods).

4
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 4. Subcellular localization of PbrBAM3-GFP. The fusion construct (35S:PbrBAM3-GFP, A–D) and a control vector with 35S:GFP (E–F) were transformed into
Arabidopsis protoplasts by PEG transformation. Images of fluorescence (A, B; see Materials and Methods) and transmitted-light (C and F), and merged images (D and
G). The horizontal scale bar = 20 μm.

Fig. 5. Morphological, structural and biochemical analyses of PbrBAM3 effect on starch metabolism in tobacco plants. (A, B) Starch iodine staining of tobacco leaves
from a wild-type plant (WT) and three lines (indicated by #s) of PbrBAM3-overexpressing plants, grown in the dark (A) and under light (B). Note that only in the WT
leaf did the illumination cause starch accumulation. (C) Ultra-structural partial view of the tobacco leaf cells of light-grown plants (WT and the three transgenic
lines). CW, cell wall; SG, starch grain. Note the relative scarcity of starch grains in the leaves of the trangenes. (D) β-Amylase activity, (E) maltose content and (F)
total soluble sugar content in the WT and the transgenic tobacco plants. ** and *** indicate significant differences between the WT and the three transgenic lines (at
P < 0.01 and P < 0.001, respectively).

2.6. Starch accumulation analysis and ultra-structure observation previously [26]. In addition, we observed the ultra-structure of the
leaves of WT and transgenic plants by an already described method
In order to analyse starch accumulation, tobacco and P. ussuriensis [43] with a minor modification. Briefly, the leaves were cut into ultra-
plants were kept at 25 °C in darkness for 12 h and then under constant thin sections and collected onto 200-mesh PELCO® copper grids (TED
light for 12 h (P. ussuriensis) or 48 h (tobacco). Iodine staining was PELLA, Inc., Redding, CA, USA). After staining with uranyl acetate and
performed to examine the in-situ starch accumulation in WT and lead citrate, the air-dried grids were visualized under the HT7700
transgenic plants, collected from dark and light treatments as described transmission electron microscope (Hitachi, Tokyo, Japan).

5
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 6. Morphological, structural and biochemical analyses of PbrBAM3 effect on starch metabolism in P. ussuriensis plants. (A, B) Starch iodine staining of P.
ussuriensis leaves from a wild-type plant (WT) and three (numbered) lines of PbrBAM3-overexpressing (OE) plants, grown in the dark (A) and under light (B). Note
that only in the WT leaf did the illumination cause starch accumulation. (C) Ultra-structural partial view of the P. ussuriensis leaf cells in light-grown plants (WT and
the three transgenic lines). CW, cell wall; SG, starch grain. Note the relative scarcity of starch grains in the leaves of the trangenes. (D) β-Amylase activity, (E) maltose
content and (F) total soluble sugar content in the WT and the transgenic P. ussuriensis plants. *, ** and *** indicate significant differences between the WT and the
three transgenic lines (at P < 0.05; P < 0.01 and P < 0.001, respectively).

2.7. Assessment of cold effects and cold tolerance hallmarks and nitro-blue tetrazolium chloride (NBT) was performed to detect the
in situ accumulation of H2O2 and O2− according to [47] and [48], re-
For the chilling-tolerance assay, 15-day-old tobacco plants from spectively. The determination in the leaves of β-amylase activity, mal-
wild-type (WT) and three transgenic lines (Fig. S1F, hereafter desig- tose content and other soluble sugars was as described previously [49]
nated as #3, #7 and #11) were chilled (at 0 °C) without cold accli- and [50].
mation, for 24 h followed by recovery at room temperature (25 °C) for
further growth. Survival rate was recorded after 10 d of recovery
2.8. Statistical analysis
growth. In addition, the 60-day-old WT and transgenic plants (#3, #7
and #11) were photographed before and after the chilling treatment
Each treatment was repeated at least three times with consistent
were taken at each time point. In parallel, the electrolyte leakage (EL),
results. Data are presented as means of three biological replicates ± SE
and malondialdehyde (MDA) and proline content in the leaves of to-
from one representative experiment. The data were analyzed by
bacco WT and transgenic (#3, #7 and #11) plants were measured as in
Duncan’s multiple range tests in the ANOVA program of SPSS (IBM
previous reports [40,42,44,45]. Cell death was assessed with trypan
SPSS 17), taking P < 0.05, P < 0.01, P < 0.001 as significantly dif-
blue staining of leaves in the WT and transgenic lines as described [46].
ferent.
In another experiment, P. ussuriensis WT plants and three transgenic
lines (OE2, OE3 and OE19; Fig.S2F), were chilled (at 0 °C) for 8 h; at the
end of the treatment, their leaves were collected for measurement of EL, 3. Results
MDA, and proline levels as well as for cell death assessment by trypan
blue staining. The chilling assay was repeated at least three times, with 3.1. PbrBAM3 cloning and sequence analysis
three biological replicates for each sample Antioxidant enzymes (SOD,
POD and CAT) activity, H2O2 level, and anti-superoxide anion activity In an earlier work, we obtained and analyzed a cold-induced β-
in WT and transgenic plants, were determined by the relevant kits amylase gene (Pbr014538.1) from a transcriptome of Pyrus betulaefolia
(Nanjing Jiancheng Bioengineering Institute) according to the manu- [51]. It contained a complete ORF of 1638 bp, flanked by an 80 bp 5’-
facturer’s instructions. Histochemical staining with 3,3′-diaminobenzi- untranslated region (UTR) and a 168 bp 3’-UTR. The cDNA, designated
dine (DAB) and nitro-blue tetrazolium chloride (NBT) was performed to as PbrBAM3, contains an intact ORF and encodes a polypeptide of 541
detect the in situ accumulation of H2O2 and O2− according to [44] and amino acids with a calculated molecular mass of 60.88 kDa and an
[45], respectively. The determination in the leaves of β-amylase ac- isoelectric point of 6.58. A phylogenetic tree was constructed using a
tivity, maltose content and other soluble sugars was as described pre- total of 43 BAM protein sequences, 8 from P. trifoliata, 9 from Arabi-
viously [46] and [47]. Antioxidant enzymes (SOD, POD and CAT) ac- dopsis, 12 from apple, 6 from rice and 9 from poplar. The BAMs from
tivity, H2O2 level, and anti-superoxide anion activity in WT and different plant species are classified into four major groups, I to IV.
transgenic plants, were determined by the relevant kits (Nanjing Jian- PbrBAM3 was grouped in the same cluster as BAM3 of Arabidopsis
cheng Bioengineering Institute) according to the manufacturer’s in- thaliana (Fig. 1). Multiple alignments between PbrBAM3 and nine other
structions. Histochemical staining with 3,3′-diaminobenzidine (DAB) BAM proteins revealed a highly conserved glucosyl hydrolase domain
(Fig. 2). In the domain, there are 18 substrate- (amylopectin-) binding

6
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 7. The effect of PbrBAM3 overexpression on cold tolerance of tobacco plants. (A, B and C), Phenotypes of 15-d-old WT and transgenic seedlings (lines #3, #7 and
#11) before (A) and after a chilling treatment (12 h at 0 °C) (B), followed by recovery at room temperature for 10 d (C). (D), Trypan blue staining of the seedlings of B.
(E) Survival rates of the transgenic and wild-type plants assessed after recovery from the chilling treatment. (F), Phenotypes of 60-d-old WT and transgenic plants
before and after the chilling treatment. (G, H and I), biochemical hallmarks of cold stress effects measured in WT and transgenic plants; electrolyte leakage (G),
malondialdehyde (MDA) (H) and proline content (I). ** and *** indicate that the values measured in the three transgenic lines were significantly different from those
in WT (at P < 0.01 and P < 0.001, respectively.).

residues (small black and blue arrow heads) and 2 catalytic residues that PbrBAM3 was localized in the chloroplast. In contrast, the GFP-
(large blue arrow heads). alone green fluorescence was found in the entire cytoplasm region
(Fig. 4E–G).
3.2. Time-course expression of PbrBAM3 in Pyrus betulaefolia leaves in
response to abiotic stresses 3.4. Analysis of amylase activity, starch and soluble sugars contents in the
transgenic plants
Upon exposure to cold, the transcript level of PbrBAM3 was up-
regulated gradually within 6 h, increased progressively until reaching The strong induction by cold of PbrBAM3 expression in Pyrus betu-
the highest level at 72 h (greater than 25-fold of the initial level), and laefolia (Fig. 3) suggested that a critical role for PbrBAM3 in the reg-
declined to half this value at 144 h (Fig. 3A). Upon exposure to dehy- ulation of cold response. To test this hypothesis, we analyzed three
dration, the transcript level of PbrBAM3 exhibited steady up-regulation independent overexpression lines of tobacco with the highest transcript
during a 6 h period, with > 12-fold increase relative to the onset of levels of PbrBAM3 (lines #3, #7 and #11; Fig. S1), along with the
treatment (Fig. 3B). ABA application resulted in a slow up-regulation of corresponding untransformed wild type (WT).
PbrBAM3 expression level, to less than 5-fold the initial value at 6 h, Neither WT nor transgenic lines of tobacco (60-day-old) grown in
peaking at over 25-fold the initial value at 24 h, followed by a decrease the dark for 12 h accumulated any starch, as revealed by the absence of
to 15-fold of the initial level at 72 h (Fig. 3C). In contrast to these, iodine staining (Fig. 5A), whereas only the leaves of transgenic plants
PbrBAM3 expression was repressed by maltose treatment to below 20% showed starch absence under constant light for 48 h in contrast to the
of initial value at 6 h and then it fluctuated remaining below 80% of WT leaf (Fig. 5B). To further confirm this result, starch accumulation of
initial value till 72 h (Fig. 3D). all tested plants after light exposure was examined by transmission
electron microscope. 60-day-old leaves were fixed in 2.5% (v/v) glu-
3.3. Subcellular localization of PbrBAM3 taraldehyde in 50 mM sodium cacodylate buffer adjusted to pH 7.3 for
16 h at 25 °C. Ultra-thin gold transverse sections of 90 nm were cut and
Confocal microscopy of Arabidopsis protoplast cells showed a per- collected onto 200 mesh copper grids that were pre-filmed with 4% (v/
fect overlap between the green fluorescence of the PbrBAM3-fused GFP v) pyroxylin in amyl acetate and carbon coated. The sections were
and the red auto-fluorescence of chloroplasts (Fig. 4A–D), indicating stained with 2% (w/v) uranyl acetate for 1 h and 1% (w/v) lead citrate

7
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 8. Cold tolerance assay of transgenic P. ussuriensis plants overexpressing PbrBAM3. (A, B) Plant phenotypes of WT and transgenic lines (OE2, OE3 and OE19)
before (A) and after (B) a chilling stress (Materials and methods). (C, D–F) Cell death (C), proline content (D), electrolyte leakage (E) and MDA (E) determined in WT
and transgenic lines after the chilling treatment (Materials and methods). ** and *** indicate values that differ significantly between the WT and the three transgenic
lines (at P < 0.01 and P < 0.001, respectively).

for 1 min. The grids with washed in water and dried in air were viewed transgenic lines stored lower levels of maltose (Fig. 6E) and total so-
in the HT7700 transmission electron microscope (Hitachi, Tokyo, luble sugars (Fig. 6F) as compared to WT.
Japan).
Consistent with the previous result, transgenic lines exhibited fewer
starch grains in comparison with the WT (Fig. 5C). As can be seen in 3.5. PbrBAM3 overexpression increased the cold tolerance of the transgenic
Fig. 5D, the BAM activity in three transgenic lines was 60%–160% plants
higher than in the WT. Consistent with the higher BAM activity, the
leaves of PbrBAM3-overexpressing tobacco plants contained more Under normal growth conditions (Materials and methods), there
maltose (Fig. 5E) and more total soluble sugars (Fig. 5F) in comparison was no apparent difference in plant morphology between the transgenic
with WT. plants and the wild type (Fig. 7A). When 15-d-old seedlings were sub-
PbrBAM3 was also transferred into P. ussuriensis to further char- jected to a freezing temperature (0 °C) for 12 h, the WT plants suffered
acterize its function (Fig.S2). Three transgenic P. ussuriensis lines (OE2, more serious cold damage compared with the transgenic plants
OE3 and OE19; 21-day-old) exhibiting a higher abundance of PbrBAM3 (Fig. 7B), and all of them were dead by the end of the recovery period,
mRNA, along with untransformed plants (WT), were subjected to dark whereas most of three transgenic plants recovered and grew well
and light treatments (Materials and methods). The results resembled (Fig. 7C). The same contrasting outcome in another experiment was
those of tobacco: no starch accumulation was seen in the leaves of all confirmed by trypan blue staining directly after the chilling treatment
the tested plants grown in the dark (Fig. 6A). However, much deeper (strikingly dark in the dead tissues and almost absent in the live seed-
staining was visualized in the WT after exposure to the light as com- lings; Fig. 7D). At the end of the recovery period, the transgenic seed-
pared with the transgenic plants (Fig. 6B).The SGs (starch granules) lings exhibited significantly higher survival rates (3# (50%), #7
were very abundant in the WT chloroplasts, whereas in the chloroplasts (41.6%) and #11 (66.7%)) plants than that of the WT seedlings (0%;
of the three transgenic plants their numbers were dramatically reduced Fig. 7E). An enhanced cold tolerance relative to the WT was also seen in
(Fig. 6C), again indicating that PbrBAM3 overexpression led to reduced 60-d-old transgenic tobacco plants (Fig. 7F).
starch accumulation. BAM activity in the three transgenic lines was After the chilling stress, EL and MDA levels in the transgenic lines
significantly higher than that in the WT (Fig. 6D). In parallel, the were significantly lower relative to the WT after chilling stress (Fig. 7G
and H). Proline (Pro) is considered as “a compatible solute” which helps

8
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 9. Oxidative stress status in WT and transgenic plants after a chilling treatment. Representative images of leaves stained in situ by the H2O2 and O2− respective
probes DAB and NBT in: (A), 15-d-old seedlings of tobacco WT and transgenics (lines #3, #7 and #11), (B), 60-d-old tobacco WT and transgenics (lines #3, #7 and
#11), (C), P. ussuriensis WT and transgenics (lines OE2, OE3 and OE19). (D, E), Levels of H2O2 (D) and O2− (E) quantified (Materials and methods) in chilled tobacco
seedlings; (F, G), Levels of H2O2 (F) and O2− (G) quantified in chilled P. ussuriensis. ** and *** indicate that the values measured in the three transgenic lines in
tobacco and P. ussuriensis were significantly different from those in their respective WT (at P < 0.01 and P < 0.001, respectively).

plants to overcome abiotic stress. Therefore, Pro contents in the tested in the leaves of WT plants (Fig. 9A–C). This result was consistent across
plants were measured. After the chilling stress, the Pro content on three all the types of the chilling assays: in the 15-day-old seedlings chilled
transgenic lines was significantly higher than in the WT (Fig. 7I). for 12 h (Fig. 9A), in the 60-d-old tobacco plants chilled for 24 h
The growth performance of WT pear and pear plants overexpressing (Fig. 9B), and in the chilled pear leaves (Fig. 9C). In confirmation of the
PbrBAM3 did not differ before the chilling treatment (Fig. 8A). How- eye-balled in-situ the histochemical staining results, the quantitation of
ever, a chilling treatment (Materials and methods) caused more pro- H2O2 and O2− (Materials and methods), also showed that following a
nounced leaf drooping in the WT than in the transgenic lines (Fig. 8B). chilling treatment, the levels of these ROS were significantly lower in
Trypan blue staining was considerably more intense in the WT leaves the transgenic plants, than in WT, both in the tobacco seedlings (Fig. 9D
relative to the transgenic lines, attesting to more sever chilling stress and E) and in P. ussuriensis (Fig. 9F and G).
due cell damage in the WT (Fig. 8C). After the chilling treatment, the
Pro content of the transgenic lines was significantly higher than that of
the WT (Fig. 8D) and the EL of OE2 (36.5%), OE3 (31.4%) and OE19 3.7. Overexpression of PbrBAM3 increased antioxidant enzyme activities in
(43.5%) plants was remarkably lower than that of the WT (87.7%) transgenic plants
(Fig. 8E). In addition, the transgenic lines displayed significantly lower
MDA content compared with the WT (Fig. 8F). Since, after the chilling treatment, the transgenic lines contained
lower ROS levels relative to the WT, we expected there higher activities
of the three significant antioxidant enzymes (SOD, CAT, and POD).
3.6. Analysis of ROS accumulation by histochemical staining Indeed, activities of all three enzymes were significantly higher in the
transgenic than in the WT plants, both in the tobacco (Fig. 10A–C) and
Under chilling treatments, the leaves of the WT plants accumulated in P. ussuriensis plants (Fig. 10D and E).
H2O2 and O2− to a larger extent than the leaves of the transgenic
plants, as revealed by in-situ histochemistry using the respective probes,
DAB and NBT (Materials and methods) showing more intense staining

9
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 10. Analysis of enzyme activities in chilled tobacco seedlings and in P. ussuriensis (Materials and methods). (A, B and C), Activity of CAT (A), POD (B), and SOD
(C) in tobacco WT and transgenic lines (#3, #7 and #11). (D, E and F), Activity of CAT (D), POD (E), and SOD (F) in P. ussuriensis WT and transgenic lines (OE2, OE3
and OE19). ** and *** indicate that the values measured in the three transgenic lines of tobacco and P. ussuriensis were significantly different from those of their
respective WTs (at P < 0.01 and P < 0.001, respectively).

3.8. Overexpression PbrBAM3 promoted the accumulation of soluble sugars starch degradation, the genes encoding enzymes leading to starch de-
gradation are crucial for sugar metabolism [54]. Previous studies sug-
To gain further insight into the physiological mechanism underlying gested that in the beginning the SGs is mainly catalysed by BAM en-
the enhanced cold tolerance of the PbrBAM3 overexpressors, the β- zymes [55], therefore, BAM genes might hold a great potential for
amylase activity and soluble sugar contents were determined in trans- modulating sugar homesostasis under abiotic stresses.
genic lines and WT plants. As shown in Fig. 11A, the β-amylase activ- According to their relationship and gene structure, PbrBAMs can be
ities in the three transgenic tobacco lines were much higher than those classified into Group IIII category, suggesting that BAMs are evolutio-
of WT after cold treatment. Maltose levels in the three transgenic to- narily conserved among different plants, the reasonable fact plants
bacco lines were significantly higher than that of WT (Fig. 11B) ranging share commonality on sugar metabolism for providing an array of su-
from 1.6 to 1.8 folds. Consistent with the accumulation of maltose, gars required for growth development and stress response. It is note-
overexressing PbrBAM3 induced accumulation of soluble sugar levels worthy that not all BAMs are responsive for starch degradation. In
under chilling treatment (Fig. 11C). Similarly, after chilling stress the β- Arabidopsis, only AtBAM3 deletion caused significantly diminished
amylase activity in the three P. ussuriensis transgenic lines (OE2, OE3 starch degradation whereas down-regulation of the other family did not
and OE19) were significantly higher than that of the WT (Fig. 11D). [56,57]. Multiple alignments revealed that PbrBAM3 shared the same
Meanwhile, the maltose and soluble sugar levels were higher in the set of amino acids related to SG binding and hydrolytic catalysis in the
transgenic lines than WT (Fig. 11E and F). These results suggest that core glucosyl hydrolase domain with AtBAM3, indicating that PbrBAM3
overexpression PbrBAM3 could enhance the accumulation of compa- is a putative BAM3 homologue of Pyrus betulaefolia. Therefore, we hy-
tibles solutes such as maltose and soluble sugar in transgenic plant. pothesized that PbrBAM3 functions in Pyrus betulaefolia as does AtBAM3
in Arabidopsis, i.e., degrades starch in SGs. Indeed, PbrBAM3-over-
expressing lines tobacco and P. ussuriensis plants exhibited a higher
4. Discussion BAM activity, a reduced accumulation of starch and elevated level of
maltose. These results confirmed the crucial role of PbrBAM3 in starch
Over a long evolutionary period, plants have evolved a range of degradation, allowing it as a desirable target of interest for modulation
complex mechanisms to tolerate harsh environmental stresses [52]. For of sugar levels.
example, sugar metabolism and carbon partitioning are two effective Given that the greatest induction of PbrBAM3 transcript level by
ways to sustain the balance required for growth and energy supply in cold treatment, indicating that PbrBAM3 plays an important role in
cold acclimation [53]. Because of carbohydrates produced mainly by

10
L. Zhao, et al. Plant Science 287 (2019) 110184

Fig. 11. Analysis of β-amylase activity and soluble sugars in transgenic tobacco and P. ussuriensis after chilling treatment. Activity of β-amylase activity (A), maltose
content (B), and soluble sugar content (C) in tobacco WT and transgenic lines (#3, #7 and #11) after chilling treatment. (D, E and F) Activity of β-amylase (D),
maltose content (E), and soluble sugar content (F) in P. ussuriensis WT and transgenic lines (OE2, OE3 and OE19) after chilling stress.** and *** indicate that the
values measured in the three transgenic lines of tobacco and P. ussuriensis were significantly different from those of their respective WTs (at P < 0.01 and
P < 0.001, respectively).

response to cold tolerance. To this end, transgenic tobacco (pear) plants the higher level of soluble sugars, implying that the production of ROS
were produced via Agrobacterium-mediated transformation of PbrBAM3 in these lines under cold stress may be scavenged in a more powerful
under the control of CaMV 35S promoter. The three selected transgenic manner. Although ROS could be generated in different cellular com-
lines (OE2, OE3 and OE19 lines of pear, and #3, #7 and #11 lines of partments, chloroplast is the major site of ROS generation [66–68].
tobacco) exhibited better phenotypic morphology, concomitant with Interestingly, PbrBAM3 is localized in the chloroplast. Therefore, we
higher levels of soluble sugars than wild type under chilling stress, speculate that the PbrBAM3-OE overexpressors contained more maltose
suggesting that the significance of sugars derived from PbrBAM3 and its metabolites in the chloroplast, which work alone or in concert
-mediated starch hydrolysis in the stress tolerance. with other antioxidants scavenging the ROS produced in this apparatus.
Cold-stress-inflicted damage – depending on its severity and dura- Moreover, the ROS produced in the cytosol can also be scavenged by
tion – consists of membrane disruption, metabolism dysfunction, loss of the soluble sugars [18]. Thus, the ROS spread from the production site
turgor or even cell death [57]. Since cold stress is frequently mediated to other places in cell may be significantly decreased, resulting in lower
by osmotic stress, their damaging effects may countered by accumula- ROS accumulation, which is consistent with the dramatic reduction of
tion of osmolytes, such as soluble sugars, restoring osmotic balance ROS level in these lines. Taken together, our results indicate that the
[28,58,59,61–65]. Sugar metabolism is a complex enzymatic process, in promotion of ROS scavenging due to greater accumulation of soluble
which several of soluble sugars may be involved, such as sucrose, glu- sugars is a feasible strategy for increasing cold tolerance of the trans-
cose, fructose etc [60]. The higher soluble sugar levels (especially for genic plants.
maltose) in the cold- stressed PbrBAM3-overexpressors in our study
suggested these transgenic plants are more capable of osmotic adjust- 5. Summary
ment than the WT plants. In parallel, the PbrBAM3 overexpressors (both
tobacco and P. ussuriensis) had lower EL and showed better re-growth In conclusion, our data demonstrate that PbrBAM3 is a stress re-
during the recovery period as compared with WT under cold stress sponsive β-amylase and plays a positive role in cold, oxidative stress
(Figs. 7 and 8), which correlated well with their increased capacity for tolerance by participating in starch degradation and the mobilization of
osmotic adjustments due to elevated soluble sugars. Therefore, over- soluble sugar capable of cell osmo-protection and ROS-counteraction.
expressing-PbrBAM3 plants produced higher soluble sugars functioning Additional work is needed to decipher the details of the links between
as osmolytes or compatible solutes, which is one of the adapted me- soluble sugars and cold responses, to allow deeper insight into the
chanisms contributing to the enhanced cold tolerance in transgenic molecular mechanisms underlying PbrBAM3 function in cold tolerance.
plants.
Soluble sugars have been shown to act as antioxidants ameliorating
Author contributions
ROS-derived oxidative stresses [19,20,63]. Here, we found that ROS
accumulation was obviously reduced in PbrBAM3 -overexpressing lines
HX designed and conceived the research. ZL, XC, HS, DH, GJ per-
compared with wild-type plants under chilling stress, consistent with
formed the experiments and YT wrote and revised the manuscript with

11
L. Zhao, et al. Plant Science 287 (2019) 110184

help of HX. HX, ZS, TS, WJ and QK contributed to proofreading and proline accumulation in cactus pear, Plant Physiol. Biochem. 46 (2008) 82–92.
critical review of this manuscript. All the authors reviewed the results [16] A. Theocharis, C. Clément, E.A. Barka, Physiological and molecular changes in
plants grown at low temperatures, Planta 235 (2012) 1091–1105.
and approved the final version of the manuscript. [17] J.V. Cabello, A.F. Lodeyro, M.D. Zurbriggen, Novel perspectives for the engineering
of abiotic stress tolerance in plants, Curr. Opin. Plant Biol. 26 (2014) 62–70.
Acknowledgments [18] I. Couée, C. Sulmon, G. Gouesbet, A. El Amrani, Involvement of soluble sugars in
reactive oxygen species balance and responses to oxidative stress in plants, J. Exp.
Bot. 57 (2006) 449–459.
We sincerely appreciate Professor Nava Moran from Hebrew [19] M.R. Bolouri Moghaddam, K. Le Roy, L. Xiang, F. Rolland, W. Van den Ende, Sugar
University of Jerusalem for giving the constructive comments and signalling and antioxidant network connections in plant cells, FEBS J. 277 (2010)
2022–2037.
language editing to this work. This work has been supported by the [20] E. Keunen, D. Peshev, J. Vangronsveld, W. Van Den Ende, A. Cuypers, Plant sugars
Excellent Youth Natural Science Foundation of Jiangsu Province (Grant are crucial players in the oxidative challenge during abiotic stress: extending the
No. BK20170086to HX), the partly supported by the open funds of the traditional concept, Plant Cell Environ. 36 (2013) 1242–1255.
[21] K. Maruyama, Y. Sakuma, M. Kasuga, Y. Ito, M. Seki, H. Goda, Y. Shimada,
State Key Laboratory of Crop Genetics and Germplasm Enhancement
S. Yoshida, K. Shinozaki, K. Yamaguchi Shinozaki, Identification of cold-inducible
(Grant No. ZW201908 to YT), the National Key Research and downstream genes of the Arabidopsis DREB1A/CBF3 transcriptional factor using two
Development Program of China (Grant No. 2018YFD1000300 to HX), microarray systems, Plant J. 38 (2004) 982–993.
the National Science Foundation of China (Grant No. 31872070 to HX), [22] H. Reinhold, S. Soyk, K. Šimková, C. Hostettler, J. Marafino, S. Mainiero,
C.K. Vaughan, J.D. Monroe, S.C. Zeeman, β-Amylase–like proteins function as
the Jiangsu Agriculture Science and Technology Innovation Fund transcription factors in Arabidopsis, controlling shoot growth and development,
(Grant No. CX(18)3065 to HX), the Fundamental Research Funds for Plant Cell (2011) 110–81950.
the Central Universities (Grant No. KYZ201607 to HX), and also sup- [23] Y. Lu, T.D. Sharkey, The importance of maltose in transitory starch breakdown,
Plant Cell Environ. 29 (2006) 353–366.
ported in part by the National Natural Science Foundation of China [24] T. Niittylä, G. Messerli, M. Trevisan, J. Chen, A.M. Smith, S.C. Zeeman, A previously
(Grant No. 31800584 to YT), Anhui Provincial Natural Science unknown maltose transporter essential for starch degradation in leaves, Science 303
Foundation (Grant No. 1808085QC75 to YT), Anhui Provincial (2004) 87–89.
[25] S.C. Zeeman, D. Thorneycroft, N. Schupp, A. Chapple, M. Weck, H. Dunstan,
Postdoctoral Science Foundation (Grant No. 2017B158 to YT), and P. Haldimann, N. Bechtold, A.M. Smith, S.M. Smith, Plastidial α-glucan phos-
Science Foundation for Anhui Province (Grant No. KJ2018A0129 to phorylase is not required for starch degradation in Arabidopsis leaves but has a role
YT). in the tolerance of abiotic stress, Plant Physiol. 135 (2004) 849–858.
[26] D.C. Fulton, M. Stettler, T. Mettler, C.K. Vaughan, J. Li, P. Francisco, M. Gil,
H. Reinhold, S. Eicke, G. Messerli, β-AMYLASE4, a noncatalytic protein required for
Appendix A. Supplementary data starch breakdown, acts upstream of three active β-amylases in Arabidopsis chlor-
oplasts, Plant Cell 20 (2008) 1040–1058.
[27] T.H. Nielsen, U. Deiting, M. Stitt, A β-amylase in potato tubers is induced by storage
Supplementary material related to this article can be found, in the
at low temperature, Plant Physiol. 113 (1997) 503–510.
online version, at doi:https://doi.org/10.1016/j.plantsci.2019.110184. [28] F. Kaplan, C.L. Guy, β-Amylase induction and the protective role of maltose during
temperature shock, Plant Physiol. 135 (2004) 1674–1684.
References [29] A. Scheidig, A. Fröhlich, S. Schulze, J.R. Lloyd, J. Kossmann, Downregulation of a
chloroplast-targeted β-amylase leads to a starch-excess phenotype in leaves, Plant J.
30 (2002) 581–591.
[1] W. Wang, B. Vinocur, A. Altman, Plant responses to drought, salinity and extreme [30] M. Zanella, G.L. Borghi, C. Pirone, M. Thalmann, D. Pazmino, A. Costa, D. Santelia,
temperatures: towards genetic engineering for stress tolerance, Planta 218 (2003) P. Trost, F. Sparla, β-amylase 1 (BAM1) degrades transitory starch to sustain proline
1–14. biosynthesis during drought stress, J. Exp. Bot. 67 (2016) 1819–1826.
[2] F. Kaplan, J. Kopka, D.Y. Sung, W. Zhao, M. Popp, R. Porat, C.L. Guy, Transcript and [31] J. Zhang, X. Li, H. Lin, K. Chong, Crop Improvement through temperature resi-
metabolite profiling during cold acclimation of Arabidopsis reveals an intricate lience, Annu. Rev. Plant Biol. 70 (2019) 753–780.
relationship of cold‐regulated gene expression with modifications in metabolite [32] L.H. Xian, P.P. Sun, S.S. Hu, J. Wu, J.H. Liu, Molecular cloning and characterization
content, Plant J. 50 (2007) 967–981. of CrNCED1, a gene encoding 9-cis-epoxycarotenoid dioxygenase in Citrus reshni,
[3] J. Zhu, C. Dong, J. Zhu, Interplay between cold-responsive gene regulation, meta- with functions in tolerance to multiple abiotic stresses, Planta 239 (2014) 61–77.
bolism and RNA processing during plant cold acclimation, Curr. Opin. Plant Biol. 10 [33] X.Q. Gong, J.Y. Zhang, J.B. Hu, W. Wang, H. Wu, Q.H. Zhang, J.H. Liu, FcWRKY70,
(2007) 290–295. a WRKY protein of Fortunella crassifolia, functions in drought tolerance and
[4] K. Nakashima, Y. Ito, K. Yamaguchi-Shinozaki, Transcriptional regulatory networks modulates putrescine synthesis by regulating arginine decarboxylase gene, Plant
in response to abiotic stresses in Arabidopsis and grasses, Plant Physiol. 149 (2009) Cell Environ. 38 (2015) 2248–2262.
88–95. [34] K. Li, X. Xu, X. Huang, Identification of differentially expressed genes related to
[5] F. Qin, K. Shinozaki, K. Yamaguchi-Shinozaki, Achievements and challenges in dehydration resistance in a highly drought-tolerant pear, Pyrus betulaefolia, as
understanding plant abiotic stress responses and tolerance, Plant Cell Physiol. 52 through RNA-Seq, PLoS One 11 (2016) e149352.
(2011) 1569–1582. [35] T. Yang, S. Zhang, Y. Hu, F. Wu, Q. Hu, G. Chen, J. Cai, T. Wu, N. Moran, L. Yu, The
[6] C. Jin, K. Li, X. Xu, H. Zhang, H. Chen, Y. Chen, J. Hao, Y. Wang, X. Huang, role of OsHAK5 in potassium acquisition and transport from roots to shoots in rice
S. Zhang, A novel NAC transcription factor, PbeNAC1, of Pyrus betulifolia confers at low potassium supply levels, Plant Physiol. 166 (2014) 945–959.
cold and drought tolerance via interacting with PbeDREBs and activating the ex- [36] K.J. Livak, T.D. Schmittgen, Analysis of relative gene expression data using real-
pression of stress-responsive genes, Front. Plant Sci. 8 (2017) 1049. time quantitative PCR and the 2-ΔΔCT method, Methods 25 (2001) 402–408.
[7] K. Nakashima, K. Yamaguchi-Shinozaki, Regulons involved in osmotic stress‐re- [37] C. Jin, X. Huang, K. Li, H. Yin, L. Li, Z. Yao, S. Zhang, Overexpression of a bHLH1
sponsive and cold stress-responsive gene expression in plants, Physiol. Plantarum. transcription factor of P. ussuriensis confers enhanced cold tolerance and increases
126 (2006) 62–71. expression of stress-responsive genes, Front. Plant Sci. 7 (2016) 441.
[8] P. Achard, F. Gong, S. Cheminant, M. Alioua, P. Hedden, P. Genschik, The cold- [38] A. Dereeper, V. Guignon, G. Blanc, S. Audic, S. Buffet, F. Chevenet, J. Dufayard,
inducible CBF1 factor-dependent signaling pathway modulates the accumulation of S. Guindon, V. Lefort, M. Lescot, Phylogeny. fr: robust phylogenetic analysis for the
the growth-repressing DELLA proteins via its effect on gibberellin metabolism, Plant non-specialist, Nucleic Acids Res. 36 (2008) 465–469.
Cell 20 (2008) 2117–2129. [39] Y. Zhang, J. Su, S. Duan, Y. Ao, J. Dai, J. Liu, P. Wang, Y. Li, B. Liu, D. Feng, A
[9] J. Liu, T. Peng, W. Dai, Critical cis-acting elements and interacting transcription highly efficient rice green tissue protoplast system for transient gene expression and
factors: key players associated with abiotic stress responses in plants, Plant Mol. studying light/chloroplast-related processes, Plant Methods 7 (2011) 30.
Biol. Rep. 32 (2014) 303–317. [40] X. Huang, J. Liu, X. Chen, Overexpression of PtrABF gene, a bZIP transcription
[10] J. Zhu, Salt and drought stress signal transduction in plants, Annu. Rev. Plant Biol. factor isolated from Poncirus trifoliata, enhances dehydration and drought tolerance
53 (2002) 247–273. in tobacco via scavenging ROS and modulating expression of stress-responsive
[11] K. Yamaguchi-Shinozaki, K. Shinozaki, Organization of cis-acting regulatory ele- genes, BMC Plant Biol. 10 (2010) 230.
ments in osmotic-and cold-stress-responsive promoters, Trends Plant Sci. 10 (2005) [41] Y.J. Yang, D.F. Wang, C.S. Wang, X.H. Wang, J.N. Li, R. Wang, Construction of high
88–94. efficiency regeneration and transformation systems of P. ussuriensis Maxim, Plant
[12] K. Shinozaki, K. Yamaguchi-Shinozaki, Gene networks involved in drought stress Cell Tissue Organ Cult. 131 (2017) 139–150.
response and tolerance, J. Exp. Bot. 58 (2007) 221–227. [42] K. Li, C. Xing, Z. Yao, X. Huang, PbrMYB21, a novel MYB protein of Pyrus betu-
[13] R. Mittler, Oxidative stress, antioxidants and stress tolerance, Trends Plant Sci. 7 laefolia, functions in drought tolerance and modulates polyamine levels by reg-
(2002) 405–410. ulating arginine decarboxylase gene, Plant Biotechnol. J. 15 (2017) 1186–1203.
[14] G. Miller, N. Suzuki, S. Ciftci Yilmaz, R. Mittler, Reactive oxygen species home- [43] P. Derbyshire, K. Findlay, M.C. McCann, K. Roberts, Cell elongation in Arabidopsis
ostasis and signalling during drought and salinity stresses, Plant Cell Environ. 33 hypocotyls involves dynamic changes in cell wall thickness, J. Exp. Bot. 58 (2007)
(2010) 453–467. 2079–2089.
[15] C.O. Silva-Ortega, A.E. Ochoa-Alfaro, J.A. Reyes-Agüero, G.A. Aguado-Santacruz, [44] L. Zhao, F. Liu, W. Xu, C. Di, S. Zhou, Y. Xue, J. Yu, Z. Su, Increased expression of
J.F. Jiménez-Bremont, Salt stress increases the expression of p5cs gene and induces OsSPX1 enhances cold/subfreezing tolerance in tobacco and Arabidopsis thaliana,

12
L. Zhao, et al. Plant Science 287 (2019) 110184

Plant Biotechnol. J. 7 (2009) 550–561. [57] J. Dreier, A. Bermpohl, R. Eichenlaub, Southern hybridization and PCR for specific
[45] X. Fu, E.U. Khan, S. Hu, Q. Fan, J. Liu, Overexpression of the betaine aldehyde detection of phytopathogenic Clavibacter michiganensis subsp. michiganensis,
trifoliate orange (Poncirus trifoliata L. Raf.), Environ. Exp. Bot. 74 (2011) 106–113. Phytopathology 85 (1995) 462–468.
[46] M. Pogány, J. Koehl, I. Heiser, E.F. Elstner, B. Barna, Juvenility of tobacco induced [58] R. Datta, M.T. Selvi, N. Seetharama, R. Sharma, Stress-mediated enhancement of β-
by cytokinin gene introduction decreases susceptibility to Tobacco necrosis virus amylase activity in pearl millet and maize leaves is dependent on light, J. Plant
and confers tolerance to oxidative stress, Physiol. Mol. Biol. Plants 65 (2004) 39–47. Physiol. 154 (1999) 657–664.
[47] M. Orozco-Cardenas, C.A. Ryan, Hydrogen peroxide is generated systemically in [59] T. Peng, X.F. Zhu, Q.J. Fan, P.P. Sun, J.H. Liu, Identification and characterization of
plant leaves by wounding and system in via the octadecanoid pathway, Proc. Acad. low temperature stress responsive genes in Poncirus trifoliata by suppression sub-
Sci. India Sect. B 96 (1999) 6553–6557. tractive hybridization, Gene 492 (2012) 220–228.
[48] H.L. Wong, R. Pinontoan, K. Hayashi, R. Tabata, T. Yaeno, K. Hasegawa, C. Kojima, [60] Z. Chen, T.A. Cuin, M. Zhou, A. Twomey, B.P. Naidu, S. Shabala, Compatible solute
H. Yoshioka, K. Iba, T. Kawasaki, Regulation of rice NADPH oxidase by binding of accumulation and stress-mitigating effects in barley genotypes contrasting in their
Rac GTPase to its N-terminal extension, Plant Cell 19 (2007) 4022–4034. salt tolerance, J. Exp. Bot. 58 (2007) 4245–4255.
[49] R.J. Laby, D. Kim, S.I. Gibson, The ram1 mutant of Arabidopsis exhibits severely [61] H.J. Bohnert, R.G. Jensen, Strategies for engineering water-stress tolerance in
decreased β-amylase activity, Plant Physiol. 127 (2001) 1798–1807. plants, Trends Biotechnol. 14 (1996) 89–97.
[50] P.S. Chow, S.M. Landhäusser, A method for routine measurements of total sugar and [62] J. Krasensky, C. Jonak, Drought, salt, and temperature stress-induced metabolic
starch content in woody plant tissues, Tree Physiol. 24 (2004) 1129–1136. rearrangements and regulatory networks, J. Exp. Bot. 63 (2012) 1593–1608.
[51] T. Yang, X. Huang, Deep sequencing-based characterization of transcriptome of P. [63] L. Xiong, J.K. Zhu, Molecular and genetic aspects of plant responses to osmotic
ussuriensis in response to cold stress, Gene 661 (2018) 109–118. stress, Plant Cell Environ. 25 (2002) 131–139.
[52] K. Shinozaki, K. Yamaguchi-Shinozaki, Molecular responses to dehydration and low [64] H. Chen, J. Jiang, Osmotic adjustment and plant adaptation to environmental
temperature: differences and cross-talk between two stress signaling pathways, changes related to drought and salinity, Environ. Rev. 18 (2010) 309–319.
Curr. Opin. Plant Biol. 3 (2000) 217–223. [65] R. Valluru, W. Van den Ende, Myo-inositol and beyond–emerging networks under
[53] V. Hurry, N. Druart, A. Cavaco, P. Gardeström, Å. Strand, Photosynthesis at low stress, Plant Sci. 181 (2011) 387–400.
temperatures A case study with Arabidopsis, Plant Cold Hardiness (2002) 161–179. [66] K. Asada, Production and scavenging of reactive oxygen species in chloroplasts and
[54] A.M. Smith, S.C. Zeeman, S.M. Smith, Starch degradation, Annu. Rev. Plant Biol. 56 their functions, Plant Physiol. 141 (2006) 391–396.
(2005) 73–98. [67] D. An, J. Yang, P. Zhang, Transcriptome profiling of low temperature-treated cas-
[55] S.C. Zeeman, S.M. Smith, A.M. Smith, The diurnal metabolism of leaf starch, sava apical shoots showed dynamic responses of tropical plant to cold stress, BMC
Biochem. J. 401 (2007) 13–28. Genomics 13 (2012) 64.
[56] F. Kaplan, C.L. Guy, RNA interference of Arabidopsis beta-amylase8 prevents [68] S.E. Weise, A.P. Weber, T.D. Sharkey, Maltose is the major form of carbon exported
maltose accumulation upon cold shock and increases sensitivity of PSII photo- from the chloroplast at night, Planta 218 (2004) 474–482.
chemical efficiency to freezing stress, Plant J. 44 (2005) 730–743.

13

You might also like