1 Title: Aluminum induces distinct changes in the metabolism of

Title: 1. Aluminum induces distinct changes in the metabolism of reactive oxygen and. 2 nitrogen species in the roots of two wheat genotypes with diff...
1 downloads 0 Views 2MB Size
Subscriber access provided by LAURENTIAN UNIV

Article

Aluminum induces distinct changes in the metabolism of reactive oxygen and nitrogen species in the roots of two wheat genotypes with different aluminum resistance Chengliang Sun, lijuan liu, Weiwei Zhou, lingli lu, chongwei jin, and Xianyong Lin J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.7b03386 • Publication Date (Web): 10 Oct 2017 Downloaded from http://pubs.acs.org on October 11, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 44

Journal of Agricultural and Food Chemistry

Title:

1 2

Aluminum induces distinct changes in the metabolism of reactive oxygen and

3

nitrogen species in the roots of two wheat genotypes with different aluminum

4

resistance

5

Running Title:

6

ROS and RNS metabolism in wheat under Al stress

7

Authors:

8

Chengliang Sun1,2a, Lijuan Liu1a, Weiwei Zhou1, Lingli Lu1,3, Chongwei Jin1,3,

9

Xianyong Lin1,3*

10

Institutions:

11

1

12

of Natural Resource & Environmental Sciences, Zhejiang University, Hangzhou

13

310058, China;

14

2

15

92521, USA

16

3

17

College of Environmental and Resource Sciences, Zhejiang University, Hangzhou

18

310058, PR China.

19

a

20

*Corresponding author

21

Prof. Dr. Xianyong Lin

22

College of Natural Resource & Environmental Sciences,

MOE Key Laboratory of Environment Remediation and Ecological Health, College

Department of Environmental Science, University of California, Riverside, CA

Key Laboratory of Subtropical Soil Science and Plant Nutrition of Zhejiang Province,

Both the authors contributed equally to this work and paper.

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

23

Zhejiang University, Hangzhou 310058, China

24

Tel.: +86 571-88982476; fax: +86 571-86971395

25

E-mail address: [email protected]

2 ACS Paragon Plus Environment

Page 2 of 44

Page 3 of 44

Journal of Agricultural and Food Chemistry

26

TOC Graphical

27

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

28

Abstract

29

Aluminum (Al) toxicity in acid soils is a primary factor limiting plant growth and

30

crop yield worldwide. Considerable genotypic variation in resistance to Al toxicity

31

has been observed in many crop species. In wheat (Triticum aestivum L.), Al

32

phytotoxicity is a complex phenomenon involving multiple physiological mechanisms

33

which are yet to be fully characterized. To elucidate the physiological and molecular

34

basis of Al toxicity in wheat, we performed a detailed analysis of reactive oxygen

35

species (ROS) and reactive nitrogen species (RNS) under Al stress in one Al-tolerant

36

(Jian-864) and one Al-sensitive (Yang-5) genotype. We found Al induced a significant

37

reduction in root growth with the magnitude of reduction always being greater in

38

Yang-5 than in Jian-864. These reductions were accompanied by significant

39

differences in changes in antioxidant enzymes and the nitric oxide (NO) metabolism

40

in these two genotypes. In the Al-sensitive genotype Yang-5, Al induced a significant

41

increase in ROS, NO, peroxynitrite (ONOO-) and activities of NADPH oxidase,

42

peroxidase and S-nitrosoglutathione reductase (GSNOR). A concomitant reduction in

43

glutathione and increase in S-nitrosoglutathione contents was also observed in Yang-5.

44

In contrast, the Al-tolerant genotype Jian-864 showed lower levels of lipid

45

peroxidation, ROS and RNS accumulation, which was likely achieved through the

46

adjustment of its antioxidant defense system to maintain redox state of the cell. These

47

results indicate that Al stress affected redox state and NO metabolism and caused

48

nitro-oxidative stress in wheat. Our findings suggest that these molecules could be

49

useful parameters for evaluating physiological conditions in wheat and other crop

4 ACS Paragon Plus Environment

Page 4 of 44

Page 5 of 44

Journal of Agricultural and Food Chemistry

50

species under adverse conditions.

51

Key words: antioxidants, nitric oxide, nitrotyrosine, RNS, ROS, wheat.

52

5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

53

Introduction

54

Aluminum (Al) is generally phytotoxic in acid soils (pH< 5.5) and is one of the major

55

factors limiting crop growth and yield worldwide.1 It has been demonstrated that Al

56

stress can have a devastating effect on plant metabolism by disrupting cellular

57

homeostasis and uncoupling major physiological processes,2,3 consequently resulting

58

in alterations in water and nutrient uptake and distribution. The underlying

59

mechanisms of Al toxicity and tolerance in plants have been broadly focused on stress

60

phytophysiology and several possible mechanisms have been proposed to explain Al

61

tolerance.4,5 However, the precise mechanisms of Al toxicity still remain largely

62

obscure and a topic of debate.

63

A common response in plants to environmental constraints is the accumulation of

64

reactive oxygen species (ROS; superoxide radicals, hydrogen peroxide, and hydroxyl

65

radicals), which are can cause irreversible damage to DNA, proteins and lipids.6 It has

66

been reported that NADPH oxidase plays a key role in ROS production under stress

67

conditions,7 when NADPH oxidase reduces molecular oxygen by using cytosolic

68

NADPH to produce O2•−, which is rapidly dismutated to H2O2 by the superoxidase

69

dismutase isoenzymes.8 For example, NADPH oxidase-generated ROS seemed to be

70

involved in nickel-induced oxidative stress in wheat roots.9 To avoid oxidative

71

stresses, plants are well equipped with ROS scavenging such as superoxide dismutase,

72

catalase, peroxidase and glutathione reductase, as well as antioxidants such as

73

glutathione and ascorbate.6 In addition, maintenance of redox homeostasis appears to

74

be critical to improving tolerance to Al toxicity in various plants, although the

6 ACS Paragon Plus Environment

Page 6 of 44

Page 7 of 44

Journal of Agricultural and Food Chemistry

75

possible role of oxidative stress in Al toxicity is not clear.

76

Recent discovery of plant cells that can generate free radical nitric oxide (NO)

77

has opened new venues of research since NO and other NO-derived molecules like

78

S-nitrosoglutamine (GSNO), nitrogen dioxide (·NO2) and peroxynitrite (ONOO-),

79

collectively known as reactive nitrogen species (RNS), also play a role in response to

80

environmental stress.10-13 Nitric oxide has been shown to be involved in a plethora of

81

abiotic stresses such as low temperature,14 arsenic,15 and salinity.16 Recently, new data

82

show that characteristics of NO metabolism is its interaction with O2•− to generate

83

ONOO- or with protein and non-protein thiols to form nitrosothiols (SNOs), which

84

can mediate post-translational modifications of different bio-molecules, mainly

85

nitration and S-nitrosylation of proteins, respectively.10,11,15,16 Analysis of NO, ONOO-,

86

and RSNO revealed that salinity,16 arsenic15 and water stress17 caused a general rise of

87

these RNS at the cellular level, consequently inducing nitrosative stress in the plants

88

under stress.

89

In plant cells, metabolism of ROS and RNS must be integrated because these two

90

families of molecules are characterized by rigorous metabolic interplay in higher

91

plants.11,18 As a result, the antioxidant metabolism and homeostasis of ROS and RNS

92

is critical for plant performance under stress conditions. Under Al stress, the concept

93

of oxidative stress is commonly used to explain cellular damage caused by the

94

imbalance between ROS production and scavenging antioxidant protective

95

mechanisms.19-21 To our knowledge, however, little information is available on the

96

endogenous metabolism of NO under Al stress conditions, although the physiological

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

97

relevance of NO in plants under Al stress conditions has been reported.3,22 The main

98

objective of our study is to elucidate the mechanisms by analyzing occurrence,

99

distribution and metabolism of key components of both ROS and RNS that take place

100

in roots of an Al-tolerant and an Al-sensitive genotype of wheat under Al stress.

101

Materials and methods

102

Plant materials and growth conditions

103

Two winter wheat (Triticum aestivum L.) genotypes, Yang-5 (Al-sensitive) and

104

Jian-864 (Al-resistant), were chosen for this study from a collection of 24 leading

105

genotypes with good agronomic traits when grown in Zhejiang Province, China,

106

where soils are strongly acidic in nature as reported in our early studies.23 Seeds were

107

surface sterilized with 1% (v/v) sodium hypochlorite and placed in a glass beaker

108

containing a small amount of deionized water overnight. After germination, seedlings

109

were transplanted to plastic screens floating on a container filled with 2.5 l of 0.5 mM

110

CaCl2 solution (pH=4.3). The solution was renewed daily. All experiments were

111

conducted in a growth chamber with a 12-h/25°C day and a 12-h/22°C night regime, a

112

light intensity of 300 µmol m-2 s-1, and a relative humidity of 70%. After 3 d of culture,

113

uniform seedlings were transferred to 1.25 l of 0.5 mM CaCl2 (pH=4.3) containing

114

AlCl3 for another 24 h.

115

Evaluation of root elongation and Al accumulation

116

Relative root elongation was calculated as the percentage of root elongated by various

117

treatments compared to the Al-free control. Total Al content in root tips (0-1 cm) was

118

analyzed by Agilent 7500A ICP-MS (Agilent, Palo Alto, CA, USA) after being 8 ACS Paragon Plus Environment

Page 8 of 44

Page 9 of 44

Journal of Agricultural and Food Chemistry

119

digested with 10 ml of 2 M HCl as in Sun et al.22 Al accumulation in root apexes was

120

detected by hematoxylin staining as described by Yamamoto et al.24

121

H2O2 and MDA determination

122

H2O2 was visually detected in root tips by using 3,3′-diaminobenzidine (DAB;

123

Sigma-Aldrich) as substrate.25 Briefly, root tips were immersed in 1 mg ml-1 solution

124

of DAB in 50 mM Tris-HCl (pH 3.8) for 4 h at 25°C. The roots stained with DAB

125

were imaged under microscope. Membrane lipid peroxidation was estimated by

126

measuring the concentration of malondialdehyde (MDA), according to the reaction

127

with thiobarbituric acid, and historical staining was conducted with Schiff’s reagent

128

(Sigma-Aldrich) as Yamamoto et al described.24

129

Assessment of the loss of integrity of the plasma membrane

130

Plasma membrane integrity was detected by staining root tips with Evans blue

131

solution (0.25%, w/v) and imaged under microscope. The trapped Evans blue was

132

evaluated by a spectrophotometric assay as describe in Yamamoto et al with minor

133

modifications.24 After staining, the retained Evans blue was released by shaking the

134

root tips in 5 ml of N, N-dimethylformamide. The absorbance was determined

135

spectrophotometrically at 600 nm.

136

Detection of O2•− , NO and ONOO- by epifluorescence microscope

137

Superoxide (O2•−) content was detected according to Rodríguez-Serrano et al.26

138

Briefly, root tips were incubated with 10 µM dihydroethidium (DHE) in 10 mM

139

Tris-HCl buffer (pH 7.4) for 30 min at 37°C, and observed under a Nikon Eclipse

140

E600 epifluorescence microscope equipped with a Nikon U-2A filter block (380-420

9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

141

nm excitation filter, 430 nm dichroic mirror, 450 nm barrier filter). Nitric oxide (NO)

142

was visualized using the fluorescent probe diaminofluorescein-FM diacetate

143

(DAF-FM DA) according to the method described by Sun et al.3 Peroxynitrite

144

(ONOO-) was detected with 10 µM 3′-(p-aminophenyl) fluorescein (APF; Cayman

145

Chemical, Ann Arbor, MI, USA) prepared in 10 mM Tris-HCl at pH 7.4 for 1 h as

146

described by Saito et al.27

147

Determination GSH, GSSG and GSNO

148

The contents of reduced and oxidized glutathione were determined fluorimetrically

149

according to a previous method.19 Briefly, root tips were homogenized in 200 µl of 25%

150

H3PO3 and 1.5 ml of sodium phosphate-EDTA buffer (pH 8.0). For reduced

151

glutathione (GSH), 100 µl supernatant was mixed with 1.8 ml of PBS-EDTA (pH 8.0)

152

and 100 µl of o-phthalaldehyde (1 mg ml-1). After thorough mixing and incubation for

153

15 min, the samples were measured on a fluorescence spectrophotometer (F 4600;

154

Hitachi Ltd, Tokyo, Japan) at excitation 420 nm, emission 350 nm and slit width 10

155

nm. For oxidized glutathione (GSSG), 0.5 ml of the supernatant was incubated with

156

0.4 ml of 0.1 M NaOH and 100 µl of 0.04 M N-ethylmaleimide for 40 min before

157

GSSG was measured using the procedure outlined for GSH assay.

158

S-nitrosoglutathione (GSNO) content was determined by the Saville method

159

according to Tanou et al.28 Briefly, root tips were frozen in liquid nitrogen and lysed

160

in 50 mM Tris-HCl, pH 8.0 containing 150 mM NaCl, 1 mM phenylmethanesufonyl

161

fluoride (PMSF). After centrifugation, the supernatant was incubated with 1 ml of 1%

162

sulfanilamide in 3 M HCl plus 1 ml of 0.02% N-(1-naphthyl) ethylenediamine in 0.2

10 ACS Paragon Plus Environment

Page 10 of 44

Page 11 of 44

Journal of Agricultural and Food Chemistry

163

M HCl with or without HgCl2 in the dark. The absorbance at 540 nm was measured

164

with a spectrophotometer (Lambda 35; PerkinElmer, Waltham, MA, USA).

165

Determination of NADPH oxidase activity of plasma membrane

166

Plasma membrane was isolated according to Zhu et al with minor modification.29

167

Briefly, root tips (about 5 g) were extracted in 10 ml of 25 mM

168

Tris-2-morpholino-ethane-sulfonic acid (MES), pH 7.6, containing 0.25 M sucrose,

169

10%

170

ethylenediaminetetraacetic acid (EDTA), 1 mM dithiothreitol (DTT), 1 mM

171

phenylmethylsulphonyl fluoride (PMSF), and 15 mM mercaptoethanol. The

172

homogenate was filtered through two layers of cotton gauze and centrifuged at 13,000

173

g for 20 min. Microsomal membranes were pelleted from the supernatant by

174

re-centrifugation at 80,000 g for 45 min. The pellet was re-suspended in a buffer

175

containing 1 mM DTT, 1 mM PMSF and 5 mM Tris-Mes (pH 6.5). The plasma

176

membranes were further isolated and purified by two-phase partitioning method. The

177

microsomal suspension was added to a phase system consisting of 6.1% (w:w)

178

dextran T-500, 6.1% (w:w) PEG3350, 0.25 M sucrose, 3 mM KCl, 25 mM

179

MES-NaOH buffer (pH 7.6). The final upper phases obtained after three successive

180

rounds of partitioning were diluted with MES-NaOH buffer (pH 7.6), and centrifuged

181

120,000 g for 30 min. The pellets were then resuspended in MES-NaOH buffer and

182

used immediately for further analysis.

(w/v)

glycerol,

0.5%

(w/v)

polyvinylpyrrolidone,

3

mM

183

The purity of the isolated plasma membrane was determined by examining the

184

activity of H+-ATPase with or without inhibitors specific for different subcellular

11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

185

H+-ATPase activities.30 Compared to the H+-ATPases activity of the control treatment,

186

H+-ATPases activity was mostly inhibited by Na3VO4 (inhibitor of P-type

187

H+-ATPases), and seldom by NaN3 (inhibitor of F-type H+-ATPases) and NaNO3

188

(inhibitors of V-type H+-ATPases) associated with plasma membrane, mitochondria

189

and tonoplast, respectively (Table S1). These results were indicative of minor

190

contaminations with other membrane fractions.

191

Assay of enzyme activities

192

For the enzyme assays, 0.15 g of root were ground with 2 ml ice-cold 50 mM

193

potassium phosphate buffer (pH 7.0) containing 1 mM EDTA and 1%

194

polyvinylpyrrolidone (PVP). The homogenate was centrifuged at 4°C for 20 min at

195

15,000 g, and the resulting supernatant was used for the determination of enzymatic

196

activity as described by Jiang & Zhang.31 An aliquot of the extract was used to

197

determine the protein content by Coomassie brilliant blue G-250, using bovine serum

198

albumin as a standard.

199

Total superoxide dismutase (SOD; EC 1.15.1.1) activity was assayed by

200

measuring the ability to inhibit the photochemical reduction of nitro blue tetrazolium

201

(NBT). Catalase (CAT; EC 1.11.1.6) activity was determined spectrophotometrically

202

by measuring the rate of decomposition of H2O2 at 240 nm. Glutathione reductase

203

(GR; EC 1.6.4.2) activity was determined by using oxidized glutathione, and the

204

oxidation rate of NADH (ε = 6.22 mM-1 cm-1) was followed at 340 nm. Peroxidase

205

(POD; EC 1.11.1.7) activity was measured by monitoring the change of absorption at

206

470 nm due to guaiacol oxidation by hydrogen peroxide. The enzyme activity was

12 ACS Paragon Plus Environment

Page 12 of 44

Page 13 of 44

Journal of Agricultural and Food Chemistry

207

calculated using the extinction coefficient of 26.6 mM-1 cm-1 for tetraguaiacol.

208

S-nitrosoglutathione reductase (GSNOR; EC 1.2.1.1) was assayed by detecting the

209

oxidation of NADH (ε = 6.22 mM-1 cm-1) at 340 nm after addition of

210

S-nitrosoglutathione to the reaction mixture at a final concentration of 400 µM.15

211

Total RNA isolation and semi-quantitative RT-PCR

212

Total RNA was isolated from wheat root tips (0-10 mm) using 1 ml Trizol reagent

213

(Life Technologies, Rockville, MD, USA) according to the manufacture’s

214

recommendation. Total RNA (3 µg) was reverse transcribed using a PrimeScript™ II

215

1st Strand cDNA Synthesis Kit (Takara, Dalian, Liaoning, China), following the

216

manufacturer’s instructions. PCR reactions were performed using 1 µl of the cDNA,

217

10 µM of each oligonucleotide primer and 1.25 U of Taq polymerase (Takara) in a 25

218

µl reaction volume. Primers used were as follows: Mn-SOD (accession number

219

AF092524)

220

5’-GCTCCCAGACATCAATTCCCAACAAA-3’ (amplifying a 417 bp fragment);

221

Cu/Zn-SOD

222

5’-CCTCTCTTCCAGGCTCCTGCC-3’

223

5’-ATGAACAACAAACGCTCTCCC-3’ (amplifying a 465 bp fragment); CAT

224

(accession number D86327), forward 5’-CTTCTCCTACTCCGACACGC-3’) and

225

reverse 5’-TGTTGATGAATCGCTCTTGC-3’ (amplifying a 311 bp fragment); POD

226

(accession number X85228) forward 5’-ACTTCCACGACTGCTTTGT-3’ and reverse

227

5’-ACTGGGCCTTCCCGATG-3’ (amplifying a 643-bp fragment); for NADPH

228

oxidase

forward

5’-ACCAGAAGCACCACGCCACCTAC-3’

(accession

(accession

number

U69632) and

number

AY561153),

13 ACS Paragon Plus Environment

and

reverse

forward reverse

forward

Journal of Agricultural and Food Chemistry

and

Page 14 of 44

229

5’-TAAGATGGAGGAAGAGGAGG-3’

230

5’-TTGAGGCAATTTTGGCTAGG-3’ (amplifying a 404-bp fragment); for 18S rRNA

231

(accession number AJ272181), forward 5’-CAAGCCATCGCTCTGGATACATT-3’

232

and reverse 5’-CCTGTTATTGCCTCAAACTTCC-3’ (amplifying a 658 bp fragment).

233

After 26 PCR cycles for 18S rRNA and 25-32 cycles for others, aliquots from the PCR

234

reaction products were loaded on 2% agarose gels and stained with ethidium bromide.

235

Statistical analysis

236

Data are the means ± SD from at least three independent experiments. All data were

237

statistically analyzed using the SPSS package (version 11.0; SPSS Inc., Chicago, IL,

238

USA), analysis of variance (ANOVA) was performed on the data sets and LSD (P