Response of biofilms-leaves of two submerged macrophytes to high ammonium

Response of biofilms-leaves of two submerged macrophytes to high ammonium

Accepted Manuscript Response of biofilms-leaves of two submerged macrophytes to high ammonium Lixue Gong, Songhe Zhang, Deqiang Chen, Kaihui Liu, Jian...

8MB Sizes 2 Downloads 18 Views

Accepted Manuscript Response of biofilms-leaves of two submerged macrophytes to high ammonium Lixue Gong, Songhe Zhang, Deqiang Chen, Kaihui Liu, Jian Lv PII:

S0045-6535(17)31569-2

DOI:

10.1016/j.chemosphere.2017.09.147

Reference:

CHEM 20020

To appear in:

ECSN

Received Date: 3 July 2017 Revised Date:

20 September 2017

Accepted Date: 30 September 2017

Please cite this article as: Gong, L., Zhang, S., Chen, D., Liu, K., Lv, J., Response of biofilmsleaves of two submerged macrophytes to high ammonium, Chemosphere (2017), doi: 10.1016/ j.chemosphere.2017.09.147. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT Response of biofilms-leaves of two submerged macrophytes to high

1

ammonium

2 3

RI PT

Lixue Gong1, Songhe Zhang1, *, Deqiang Chen1, Kaihui Liu1, Jian Lv2

4 5

1

6

Development on Shallow Lakes, College of Environment, Hohai University, Nanjing

7

210098, China.

8

2

9

Yantai Institute of Coastal Zone Research, Chinese Academy of Sciences, Yantai,

SC

Ministry of Education Key Laboratory of Integrated Regulation and Resource

10

Shandong 264003, PR China.

11

*

12

Email: [email protected];

Corresponding author

AC C

EP

TE D

13

M AN U

Key Laboratory of Coastal Environmental Processes and Ecological Remediation,

1

ACCEPTED MANUSCRIPT Abstract: Submerged macrophytes can provide attached surface for biofilms (known

15

as periphyton) growth. In the present study, the alterations in biofilms formation, and

16

chemical compositions and physiological responses were investigated on leaves of

17

Vallisneria asiatica and Hydrilla verticillata exposed to 0.1 mg L-1 (control) or with

18

10 mg L-1 NH4+-N for 13 days. Results from physiological and biochemical indices

19

(content of H2O2, malondialdehyde, total chlorophyll and activity of superoxide

20

dismutase, catalase and peroxidase) showed that high ammonium caused oxidative

21

damage to leaves of two species of plant. Multifractal analysis (based on scanning

22

electron microscope images) showed that for the same plant, the values of width △α

23

(△α=αmax-αmin) of the f(α) and ∆f (∆f = f(αmin)−f(αmax)) were smaller on leaves surface

24

of two species of plant treated with 10 mg L-1 NH4+-N for 13 days than their controls,

25

suggesting high ammonium treatments reduced morphological heterogeneity of leaf

26

surface and enhanced area of the colony-like biofilms. X-ray photoelectron

27

spectroscopy analysis showed that C, O, N and P were dominant elements on leaves

28

surface of two species of plant and ammonium application increased the percentage of

29

C but decreased that of O. High ammonium increased C1 (C-C or C-H) percentage

30

but decreased C2 (C-O) and C3 (O-C-O or C=O) percentage on leaves surface of two

31

species of plant, indicating that ammonium stress changed the surface chemical states

32

and thus might reduce the capacity of leaves to adsorb nutrients from water column.

33

Our results provided useful information to understand ammonium induced toxicity to

34

submerged macrophytes.

SC

M AN U

TE D

EP

AC C

35

RI PT

14

36

Keywords: multifractal analysis, X-ray photoelectron spectroscopy, functional

37

components, ammonia nitrogen

38

2

ACCEPTED MANUSCRIPT 39

1. Introduction Submerged macrophytes growing under water surface are conspicuous

41

components of shallow water environments (Xiao, 2014; Phillips et al., 2016), but

42

they have declined extensively in eutrophic lakes due to the high nutrient (Wersal and

43

Madsen, 2011), toxic substances (Nuttens et al., 2016), sediments organic enrichment

44

(Raun et al., 2010), periphyton (Peterson et al., 2007) and low light (Lyche-Solheim et

45

al., 2013) in recent decades. Especially, high nitrogen concentrations in water column

46

have been considered as key factors causing the decline of species richness of

47

submerged macrophytes in shallow lakes (James et al., 2005; Barker et al., 2008).

48

Among dissolved nitrogen species, ammonium has been linked to the decrease of

49

submerged macrophytes in water column (Wang et al., 2008; Yuan et al., 2013).

50

Ammonium concentration above 5 mg L-1 NH4+-N that can can be considered as

51

excess/high ammonium, because it can induce toxicity to submersed macrophyte

52

(Wang et al., 2008 and 2010; Meng et al., 2010; Gao et al., 2015). High ammonium

53

can increase free amino acids content and photorespiration, induce oxidative stress,

54

affect the carbon and nitrogen metabolisms and inhibit the photosynthetic system and

55

plant growth of submerged macrophytes (Wang et al., 2008; Cao et al., 2011). Though

56

abundant data have been documented about the effects of ammonium on submerged

57

macrophytes, the ammonium-toxicity remains far from clear.

EP

TE D

M AN U

SC

RI PT

40

Aquatic plants can provide attached surface for bacteria, algae and other

59

microorganisms settlement forming microbial communities (known as biofilms or

60

periphyton) (Michael et al., 2008). Biofilms and submerged macrophytes compete for

61

dissolved inorganic carbon in water column (Jones et al., 2002). In addition, excess

62

boifilms have negative effects on their hosts through creating physical barriers to

63

nutrient uptake and gas exchange, or a combination of these factors (Drake et al.,

64

2003). Furthermore, the growth of epiphytic algae directly induces oxidative stress on

65

V. natans and epiphytic algal biomass, which are positively correlated with nutrient

66

(nitrogen and phosphorus) available in the water column (Song et al., 2015).

67

Moreover, algae in biofilms can form high-oxygen, high pH, and low-carbon micro

AC C

58

3

ACCEPTED MANUSCRIPT environment at the interface between biofilms and surface of submerged macrophytes

69

(Jones and Hardwick, 2000; Song et al., 2015). These data provided us useful

70

information in understanding the biofilm-induced stresses on submerged macrophytes.

71

However, little is known about the biofilms distribution and its potential impact on

72

leaf surface properties of submerged macrophytes under high ammonium.

RI PT

68

Fractal analysis is a powerful tool to describe the self-similarity between the tiny

74

part and the whole and to explain the complex nonlinear phenomena in nature

75

(Mandelbrot, 1967; Bialowiec et al., 2010). Therefore, it has been used as a statistical

76

method and mathematical model to find all possible sub-divisions of the object

77

(Peitgen et al., 2004), while multifractal describes a measurement that is defined in a

78

certain area or volume and can decompose the defined domain into a series of

79

sub-domains in space, and according to the singularity of this measurement every

80

sub-domain constitutes a single fractal (Zhang et al., 2010b). Multifractal analysis has

81

been used to investigate the structural and fractal characteristics of biofilms attached

82

to the surface of plants and gravels (Liang et al., 2013).

TE D

M AN U

SC

73

X-ray photoelectron spectroscopy (XPS) is one of the most effective tools for

84

surface analysis and is able to obtain a wealth of chemical information (Cai et al.,

85

2014), because XPS survey spectra could provide the inner electron binding energy of

86

almost all elements except H and He within a few surface nanometers (10 nm ) of

87

materials (Matuana et al., 2001; Wepasnick et al., 2010). XPS has been used to

88

describe the leaf surface properties of terrestrial plants, which are directly related with

89

leaf material absorption, disease resistance and other effects (Stevens and Baker,

90

1987).

AC C

91

EP

83

In the present study, an experiment was conducted to examine the effects of high

92

ammonium on leaf-biofilms of Vallisneria asiatica (V. asiatica) and Hydrilla

93

verticillata (H. verticillata). We analyzed the biofilms distribution, morphological and

94

physiological alterations on plant surface and antioxidant defense system in leaves of

95

two species of plants under 10 mg L-1 NH4+-N. The study was to test the following

96

hypotheses: high ammonium may (1) enhance the growth of biofilms on leaves of two 4

ACCEPTED MANUSCRIPT 97

species of plant; (2) alter the morphological characteristics of leaves surface of two

98

species of plant; and (3) change the oxidative properties of plant leaves.

99

2. Materials and methods 2.1. Experimental setup

RI PT

100

Healthy plants of submerged macrophytes, V. asiatica and H. verticillata, were

102

used in this experiment. After rinsed with tap water, these plants were acclimated in

103

140 L (750*550*430 mm) plastic container with 70 mm sediments and 260 mm water

104

depth for two weeks in greenhouse. V. asiatica and H. verticillata (initial optical

105

density of 1:1) were treated with 0.1 mg L-1 (control) and 10 mg L-1 NH4+-N (high

106

ammonium) for 13 days according to previous reports (Wang et al., 2008; Gao et al.,

107

2015; Cao et al., 2011). To maintain the NH4+-N concentration, the ammonium

108

concentrations were determined by continuous flow Analyzer Auto Analyzer3 (SEAL,

109

Germany) every day. The water was replaced every two days to maintain the content

110

of NH4+-N using static methods. In the present study, the plant samples were

111

harvested at 13 days post treatments. Four replicates were used for each treatment of

112

each species of plant.

113

2.2. Scanning electron microscope (SEM) and multifractal analysis

TE D

M AN U

SC

101

SEM was used to visualize the distribution of bacteria and algae on the biofilms

115

surface. For SEM analysis, plant leaves were cut into 0.6 × 0.6 cm squares, then,

116

further fixed in 2.5% glutaraldehyde solution. After a double rinse with 0.1 M sodium

117

phosphate buffer (PBS, pH 7.4), leaf samples were dehydrated with a serial of ethanol

118

concentration (30, 50, 70, 80 and 90%) for 15 min and with three times of 100%

119

ethanol for 15 min. Dried samples were analyzed by SEM (Hitachi, Japan, S-3400N

120

Ⅱ).

AC C

EP

114

121

The combination of scanning electron microscopy and fractal theory is an

122

effective tool to characterize the microstructure of biofilms, because it can reveal the

123

discipline and essential connection which hide behind the complex phenomenon, part

124

and entirety (Liang et al., 2013). In this study, box counting method was applied 5

ACCEPTED MANUSCRIPT 125

according to zhang et al (Zhang et al., 2010b). Briefly, the black-white SEM images

126

were divided into many boxes of size ε×ε, where ε=1/L (L=256, 128, 64, 32, 4, 2 or 1).

127

To calculate the multifractal spectrum, the biofilms value distribution probability in

128

the box (i, j) was expressed as

RI PT

(1)

129

Where is the biofilms value of the box (i, j) of size ε.

130

Pij(ε) can be described as multifractal as

SC

131

(2)

132

(3)

M AN U

133

where the exponent α depending upon the box (i, j) is the singularity of the

135

subset of probabilities, the number of boxes of size with the same biofilms value

136

distribution probability, and the fraction dimension of the α subset. The dependence of

137

on α is the multifractal spectrum. Generally, the fractal dimension can be obtained

138

from the partition function expressed as a power law of ε with an exponent applied in

139

statistical physics as following:

TE D

134

140

(4)

EP

where q is the moment order (−∞< q <∞), which

141

can be obtained from the

slope of ln(χq(ε)) ~ ln(ε) curve. And the generalized fractal dimension Dq is defined

143

as:

144 145 146 147

AC C

142

(5)

f(α) can be obtained by performing Legendre transformation as followings: (6)

Dq=

and -

148

(7)

149

The value of q cannot be infinite in real calculation in practical calculations

150

(Raoufi et al., 2008). Based on suggestions from (Wickens et al., 2014), we take the 6

ACCEPTED MANUSCRIPT maximum |q| as 10 if |dαmax|/∆α and |dαmin|/∆α were all less than 0.1%. The width of

152

the multifractal spectrum is ∆α and the difference of the fractal dimensions of the

153

maximum probability (α=αmin) and the minimum one (α=αmax) is ∆f (∆f =

154

f(αmin)−f(αmax)).

155

2.3. X-ray photoelectron spectroscopy (XPS) analysis

RI PT

151

XPS was employed to analyze the nutrient element compositions and oxidation

157

state of leaf surface under high ammonia nitrogen. For XPS analysis, fresh healthy

158

leaves were dried in the oven at 45℃ for 12 h after washing with distilled water. The

159

charge

160

electron-neutralizing gun. The XPS analyses were carried out with a PHI (5000

161

VersaProbe, UK) spectrometer with a hemispherical energy analyser and using a

162

monochromatic Al K a source (1486.6 eV). As the delay-line detector allows a high

163

count rate the power applied to the X-ray anode was reduced to 25 W so that the

164

possible X-ray induced degradation of the sample was minimized. High-resolution

165

spectra were collected using an analysis area of ≈300 nm × 700 nm and either a 58.7

166

eV pass energy. The takeoff angle was 90° in all measurements. The binding energies

167

of the peaks were determined using the C1s peak at 284.6 eV (Fang and Wan, 2008).

168

XPS peaks corresponding to C and O can be analyzed by deconvoluting using

169

Gussian -Lorentzian mixed curve fitting (Kontturi, 2005).

170

2.4. Aquatic parameters determinations and cells counting

of

dried

sample

was

eliminated

with

an

EP

TE D

M AN U

effect

AC C

171

neutralization

SC

156

Parameters including dissolved oxygen

(DO), pH, oxidation-reduction

172

potential (ORP), electrical conductivity (EC) was examined in water by Portable

173

Instruments (YSI, WEISS instrument,American) every day. NH4+-N content in the

174

water was determined by AtuoAnalyzer 3 system (SEAL, Germany). The values of

175

DO, pH, EC and ORP were provided in Figure S1 (See supplementary materials).

176

For cells counting, approximately 1 g leaf samples from healthy plants growing

177

at 0–20 cm below the water surface were harvested to prevent contamination from

178

sediments and was transferred into a sterile 100 mL polyethylene bottle containing 40 7

ACCEPTED MANUSCRIPT mL of 50 mM PBS solution. After treated with ultra-sonication for 3 min, mixtures

180

were shook for 30 min at 225 r min-1 and exposed to ultra-sonication for 3 min. Three

181

extracted suspensions from the same sample were combined and fixed with 2%

182

formaldehyde. The 100 µL suspension samples was stained with 10 µg mL−1 4,

183

6-diamino-2-phenyl indole (EDTA, 700 µL) and incubated in the dark for 30 min. The

184

samples were filtered through a 0.22 µm incubated black filter. The numbers of

185

bacteria on the black membrane were counted under a fluorescence microscope

186

(ZEISS, Germany). According to our previous reports, biofilms can be formed rapidly

187

and stabled within 10 dyas (Zhang et al., 2016). Therefore, the cell counting were

188

conducted for biofilms samples at 1 day (initial microbe densities) and 13day (stable

189

microbe densities).

190

2.5. Physiological parameters determinations

M AN U

SC

RI PT

179

About 0.2 g sample of fresh leaf samples was ground in liquid nitrogen for

192

chlorophyll. After extraction with 10 ml of 96% ethanol for 24 h in darkness at 4◦C,

193

the samples were centrifuged at 10,000×g for 10 min. The absorbance of supernatant

194

was determined at 649 and 665 nm. The contents of total chlorophyll (a+b) were

195

calculated according to the previous described method (Wang et al., 2008).

TE D

191

About 0.5 g fresh leaf samples was used for H2O2 determination (Wang et al.,

197

2008). After ground in liquid nitrogen, leaf sample was homogenized in 3 mL of 50

198

mM sodium PBS (pH 6.5) and then was centrifuged at 8000×g for 20 min. For the

199

determination of H2O2, a mixture was prepared with suitable of supernatant and 2.5

200

ml of 0.1% titanium sulfate in 20% (v/v) H2SO4 and centrifuged at 10000×g for 5 min.

201

The absorbance of supernatant was measured at 410 nm, and the absorbance values

202

were calibrated against a standard curve generated with known concentrations of

203

H2O2.

AC C

EP

196

204

Malondialdehyde (MDA) was measured to indicate polyunsaturated fatty acid

205

oxidation as a secondary end product and lipid peroxidation intensity and the injury

206

degree of membrane system (Gunes et al., 2007; Li and Ni, 2009). About 0.5 g 8

ACCEPTED MANUSCRIPT powdered leaf samples was used for MDA content determination according to the

208

method described in previous report (Wang et al., 2010). Briefly, samples were mixed

209

with 5ml of trichloroacetic acid (1%, TCA). The mixtures were centrifuged at

210

15,000×g for 10 min, and 0.5 ml of supernatant was mixed with 2 ml of 20% TCA

211

containing 0.5% 2-Thiobarbituric acid and heated at 96 ºC for 30 min. After cooled at

212

0ºC, the mixtures were centrifuged at 10,000×g for 5 min at 4 ºC and the absorbance

213

of supernatant was at 532 and 600 nm. The MDA concentration was calculated from

214

the difference in absorbance at 532 and 600 nm using an extinction coefficient of 155

215

mM-1cm-1.

216

2.6. Assay of Superoxide dismutase (SOD), Catalase (CAT) and Peroxidase (POD)

217

activity

M AN U

SC

RI PT

207

For enzyme extraction, 1 g of plant leaf samples was ground in liquid nitrogen

219

and mixed with 10 ml of 50 mM potassium phosphate buffer (pH 7.0) containing 1

220

mM EDTA and 1% polyvinylpyrrolidone. The mixture was centrifuged at 15,000×g

221

for 20 min at 4◦C, and the supernatant was used for the enzyme assays and protein

222

determination. The determinations of protein content, SOD, CAT and POD activity

223

were based on the previous described method (Wang et al., 2011).

224

2.7. Data processing

EP

TE D

218

One-way analysis of variance was used to compare differences between

226

treatments using SPSS 18.0 (IBM, USA) (p<0.05). Origin8.0 (OriginLab Corp, USA)

227

was used to process XPS data analysis as well as fractal analysis.

AC C

225

9

ACCEPTED MANUSCRIPT 228 229

3. Results and discussion

230

3.1.Biofilms community and growth analysis Figure 1

232

Nutrient loading is currently the most documented factor influencing epiphyte

233

communities (Michael et al., 2008). As revealed by SEM images, bacteria, algae and

234

microbial aggregates were observed on leaves surface of V. asiatica and H.

235

verticillata in all treatments (Fig. 1). As compared with control plants (Fig. 1a and 1c)

236

and plants at 1d (Fig. S2a and S2b), biofilms almost covered all the leaf surface,

237

especially on leaves of H. verticillata exposed to 10 mg L-1 NH4+-N for 13 days

238

(Fig.1d). Microbes including algae, cocci and Bacillus were also detected on the

239

surface of submerged macrophytes under high nitrate (Zhang et al., 2016).

M AN U

SC

RI PT

231

In control groups, the microbe densities increased from 9.73 × 105 to 4.46 ×

241

107cells g-1 Fresh Weight on leaves of H. verticillata and from 3.02 × 106 to 5.17 ×

242

107cells g-1 Fresh weight on leaves of V. asiatica after 13 days (Fig. 1e and 1f). After

243

treatments with 10 mg L-1 NH4+-N for 13 days, microbe densities increased

244

significantly in biofilms from two species of plant and reached to 1.63 × 108 cells g-1

245

Fresh weight and 3.79 × 108 cells g-1 Fresh weight, respectively, on leaves of V.

246

asiatica and H. verticillata (Fig. 1e and 1f). Previous report has shown that high

247

nutrimental (combinations of nitrogen and phosphorus) concentrations contributed to

248

the overgrowth of the epiphytic algae on aquatic macrophytes in shallow lakes (Song

249

et al., 2015), while our results demonstrated that ammonium application alone could

250

also stimulate the biofilms growth.

251

3.2. Multifractal analysis of leaf biofilms

AC C

EP

TE D

240

252

Table 1

253

There are complex interactions between biofilms and its hosts. Surface

254

topography and chemistry properties of substrates have an effect on cell dispersion, 10

ACCEPTED MANUSCRIPT density and clustering of microbes in biofilms (Wickens et al., 2014). Our previous

256

report showed that the microbes in biofilms can be affected by substrates,

257

environmental parameters and the distributions of biofilms on leaves of submerged

258

macrophytes (Zhang et al., 2016). However, little is known the impacts of ammonium

259

application on the distribution of biofilms. In this study, multifractal analysis was

260

employed to investigate whether ammonium application can alter the morphological

261

characterizations of leaf surface based on SEM images.

RI PT

255

In multifractal analysis, the linear regions in the graph of the curve cluster are

263

known as scaling regions, and curve plot for all moments of q should be strict

264

linearity that can tend to zero infinitely (ln ε→−∞). A random multifractal must have

265

preferable linearity (Raoufi et al., 2008). With the different biofilms as examples,

266

representative plots of ln(χq(ε)) ~ ln(ε) for values of q (ranged from -10 to 10) were

267

generated (Fig. S2c and S2d). For images from treatments at 13th days, the favorable

268

linear correlation were observed in the ln(χq(ε)) ~ ln(ε) plot (Fig. S2d), while linearity

269

was not observed for images from two species of plant at first day (Fig.S2c). These

270

results indicated that obvious multifractal characteristics can be observed on surface

271

structures of leaf biofilms of two species of plant at 13th day.

M AN U

TE D

272

SC

262

Figure 2

Multifractal spectra f(α) of the samples from two plants at 13th day and their

274

important parameters were shown in Fig. 2 and Table 1, respectively. Hook-like

275

multifractal spectrums f(α) of leaf biofilms at 13th day were obtained and showed that

276

the shape and the width △α (△α=αmax-αmin) of the f(α) of plants treated with

277

ammonium were different from their controls (Fig. 2). The αmin and αmax are the

278

variables of singularity indices of the smallest and biggest biofilms area distribution

279

rules with the changes of ε, respectively. The smaller αmin is, the bigger the maximum

280

probability will be; and the bigger αmax is, the smaller the minimum probability will

281

be. Thus the span of singularity indexes △α can quantitatively describe the degree of

282

heterogeneity of biofilms distribution probability in biofilms. The width

283

△α(△α=αmax-αmin) of the f(α) for leaf biofilms of V. asiatica and H. verticillata plants

AC C

EP

273

11

ACCEPTED MANUSCRIPT treated with 10 mg L-1 NH4+-N were smaller than their controls (Table 1). The larger

285

the width ∆α (∆α=∆αmax−∆αmin) of the multifractal spectra f(α) (Zhang et al., 2010b),

286

in other words, the more the roughness will be. For the same plants, the ∆α (∆α=∆αmax

287

−∆αmin) of control samples were higher than that of samples exposed to high

288

ammonium (Table 1), suggesting that biofilm growth reduced the roughness of leaf

289

surface.

RI PT

284

The ∆f of control was higher than that of treatments with ammonium for 13th day

291

for the same plants (Table 1). The f(α) can be used to depict the distribution of

292

microbes (from single clone to blocks) on plant leaves. The f(αmin) and f(αmax) showed

293

the minimum and maximum subsets of biofilms distribution probability, respectively.

294

The ∆f (∆f = f(αmin)−f(αmax) is the number difference between the maximum and

295

minimum probability subsets. When big probability subsets predominate, ∆f > 0;

296

when small probability subsets predominate, ∆f < 0. All ∆f

297

0 (Table 1), indicating that the microbes in subset of maximum biofilms distribution

298

probability is larger than that in the minimum subset. In this study, the △f value of

299

control was bigger than that of samples exposed to high ammonia nitrogen, indicating

300

the distribution probability of single colony were higher on control than that of

301

samples exposed to 10 mg L-1 NH4+-N. Namely, high ammonium application

302

enhanced the growth of single colony-like biofilms.

M AN U

SC

290

303

3.3.Response of chemical properties on leaf surface to ammonium

304

AC C

EP

TE D

were actually larger than

305

Figure 3

Leaf surface properties are closely related to their absorption of nutrients, gas

306

exchange, regulation of transpiration and secretion (Zhang et al., 2010b). XPS

307

technology has been widely used to characterize surface morphology and chemistry of

308

plant materials (Cai et al., 2014). In the present study, XPS was employed to

309

determine the chemical composition on leaves of two species of plant. Elements C, O,

310

N and P were detected on leaves surface of two species of plant, while Cl and K were

311

only detected in V. asiatica (Table S1 and Fig. 3). In general, the relative percentages 12

ACCEPTED MANUSCRIPT of C increased from 70.42% to 74.71% and from 78.45% to 83.66% on leaves surface

313

of V. asiatica and H. verticillata, respectively, but decreased that of O, from 22.92%

314

to 16.11% and from 15.53% to 11.9%, on leaf surfaces of V. asiatica and H.

315

verticillata, respectively. Especially, the relative percentages of N increased from

316

5.64% to 7.99% on leaf surfaces of V. asiatica, but decreased from 5.52% to 4.11% on

317

leaf surfaces of H. verticillata. While no significant alterations in percentages of P

318

were observed. These data demonstrated that ammonium application altered chemical

319

composition of leaf surface.

SC

RI PT

312

There were significant differences in the ratios of oxygen to carbon and nitrogen

321

to carbon on different plant species. The ratio values of O/C and N/C were 0.33 and

322

0.08, respectively in the control groups of V. asiatica and were 0.22 and 0.11 in the

323

treatment groups for V. asiatica, respectively. Meanwhile, the values of the O/C and

324

N/C were 0.20 and 0.07, respectively, in the control groups of V. asiatica, and the

325

corresponding values were 0.14 and 0.05 in the treatment groups for H. verticillata

326

(Table S1), respectively. Since ratio values of O/C can be used as an indicator of

327

smooth or roughness of wood surface (Sinn et al., 2001), in this study, the decreased

328

O/C ratio indicated that high ammonium reduced roughness of leaf-biofilms surface.

329

That was in accord with the results from multifractal analysis. The alterations in N/C

330

ratio may be ascribed to the instability of hydroxyl group in Cn(H2O)m that the amino

331

group can bind to Cn(H2O)m by replacing the hydroxyl group.

333

TE D

EP

AC C

332

M AN U

320

Figure 4

Fig. 4 shows the high-resolution C 1s signal following curve-fitting and three

334

components of C were detected in all samples. The major component C1 (at ~284.4

335

eV) is due to hydrocarbon (C-C or C-H). Peak value of C2 at ~285.8 eV consisted of

336

carbon connect with a non-carbonyl oxygen (C-O), while C3 (at 287.1-287.8 eV)

337

express the carbon attach to two non-carbonyl group oxygen (O-C-O) or a carbonyl

338

oxygen (C=O). C1, C2 and C3 of C 1s peaks of leaves could indicate cuticle,

339

cutinized layer and oxidation state, respectively (Cai et al., 2014). In the present study, 13

ACCEPTED MANUSCRIPT C1 was the dominant component, followed by C2 and C3 (Fig. 4). Percentages of C2

341

and C3 were higher on leaves surface of control than that of ammonia-treated plants.

342

Component Cl reflect aliphatic chains, while components C2 and C3 reflect hydroxyl,

343

epoxy and ether functions (Genet et al., 2002). The decrease of C2 and C3 in this

344

study can be explained that ammonia treatment decreased the contents of

345

nonstructural carbohydrate, soluble sugar, sucrose, fructose, and starch in leaves of

346

submerged macrophytes (Genet et al., 2002).

RI PT

340

Figure 5.

348

High-resolution XPS O 1s peak was decomposed into two components fixed at

349

about 531.2 eV ( O1) and 532.6 eV (O2), and attributed to oxygen making a double

350

bond (O=C) and one or two single bonds (-O-) with carbon, respectively (Fig. 5).

351

After ammonia treatment, there lative peak area ratio of C1/C2 and C1/C3 increased

352

from 1.7 to 3.86 and from 5.06 to 7.42 for V. asiatica (Fig. 4a), respectively, and

353

increased from 1.35 to 1.77 and 3.56 to 5.53 for H. verticillata (Fig. 4b), respectively;

354

while that of O1/O2 decreased from 1.05 to 0.96 and from 1.05 to 0.99 for V. asiatica

355

and H. verticillata after 10 mg L-1 NH4+-N, respectively. The reduction of C3 and O1

356

(the same function group) suggests that the surface chemical states were more active

357

in control plants than in plants exposed to 10 mg L-1 NH4+-N. A reduction in chemical

358

active reflects a reduced capacity of leaves to adsorb the nutrients from water for two

359

species of plant under ammonium stress.

360

3.4.Alterations in intracellular oxidation/reduction status

362

M AN U

TE D

EP

AC C

361

SC

347

Figure 6

Reactive oxygen species (including H2O2) can be actively induced under various

363

stresses. An increase of H2O2 concentration, respectively, 2.92 and 1.41 times of their

364

control, was detected in leaves of V. asiatica and H. verticillata treated with 10 mg L-1

365

NH4+-N for 13 days (Fig. 6a), suggesting that the cell tended to be an oxidation status

366

under ammonium stress. Excess reactive oxygen species could induce the production

367

of MDA, which can be produced from polyunsaturated fatty acid oxidation as a 14

ACCEPTED MANUSCRIPT secondary end product (Wang et al., 2008; Wang et al., 2010). MDA content (Fig. 6b)

369

increased in two species of plant under ammonium stress, indicating cell membrane

370

were damaged contributing to the leakage of plant cells (Wang et al., 2008; Wang et

371

al., 2010). The activities of SOD, CAT and POD were visibly higher in the leaves of

372

two species of plant at the 10 mg L-1 NH4+-N than those in their controls (Fig. 6d-f).

373

SOD dismutase superoxide to H2O2, which can be decomposed to water and oxygen

374

by the activities of CAT and POD (Wang et al., 2008). These results indicate that

375

excessive energy is consumed by plants for biosynthesis of these enzymes to cope

376

with ammonia-induced oxidative stress.

SC

RI PT

368

Compared with the control, the total chlorophyll content of the V. asiatica and H.

378

verticillata decreased significantly with 10 mg L-1 NH4+-N at the 13th day (p<0.05)

379

(Fig. 6c). In the present study, dissolved oxygen concentration was not altered

380

obviously in overlying water at 1-7 days post ammonium application, but decreased

381

vastly after 7 days of ammonia exposure (Fig. S2a). The decrease of total chlorophyll

382

content may be ascribed to the interaction effect of biofilms and ammonia toxicity and

383

contribute to the decrease of dissolved oxygen concentrations in overlying water (Fig.

384

S2a).

TE D

M AN U

377

Biofilms growth can have negative impacts on plant cells. For example, increase

386

of biofilms area can directly increase light attenuation (Xie et al., 2013), and

387

periphyton can attenuate most of the light reaching the adaxial leaves surface of

388

Potamogeton perfoliatus (Toth, 2013). Low light and high nutrients can induce the

389

activities of antioxidant enzymes including SOD and POD in submerged macrophytes

390

Potamogeton crispus (Zhang et al., 2010a). Ammonium application stimulated the

391

biofilms growth, which potentially changed the nutrient compositions and the

392

chemical activity of leaf biofilms surface (revealed by XPS). Excess biofilms can also

393

enhance oxidative stress (Song et al., 2015). These data suggested that biofilms

394

contributed to ammonium-induced stress on submersed macrophytes.

395

4. Conclusions

AC C

EP

385

15

ACCEPTED MANUSCRIPT Effects of ammonium on biofilms and antioxidant responses in leaves of V.

397

asiatica and H. verticillata were investigated. The batch experiments illustrated that

398

high ammonia nitrogen promoted the growth of biofilms on the leaf surface, while

399

biofilms growth altered the surface topography and chemistry activity of leaf surfaces

400

of two species of plant exposed to high ammonium. In addition, the leaf surface

401

mainly consisted of C, N, O and P. The ratio of O/C on leaves surface of V. asiatica

402

and H. verticillata decreased after treating with 10 mg L-1 NH4+-N for 13 days,

403

suggesting that the chemical properties of the leaf surface were more active in the

404

control group than that of plants exposed to 10 mg L-1 NH4+-N. Moreover, 10 mg L-1

405

NH4+-N induced oxidative stress and antioxidant system stress response in two

406

species of plant. This study provided useful information for understanding the

407

oxidative mechanism of ammonium stress on submerged macrophytes.

408

Acknowledgements

M AN U

SC

RI PT

396

This study was, in part, supported by Grants from the National Natural Science

410

Foundation of China (Grant No. 51379063 and 51579075), the Natural Science

411

Foundation of JiangSu Province for Excellsent Young Scholars (BK20160087), the

412

Fundamental Research Funds for the Central Universities (2016B06714) and a Project

413

Funded by the Priority Academic Program Development of Jiangsu Higher Education

414

Institutions.

415

Competing financial interests

417 418

EP

AC C

416

TE D

409

The authors declare no competing financial interests.

Supplementary materials Fig.S1. The daily alterations in dissolved oxygen concentration (a),

419

oxidation-reduction potential (ORP) (b), pH (c) and electrical conductivity (d) during

420

experiment.

421 422

Fig. S2. Representative Plots of ln(χq(ε)) ~ ln(ε) for values of q for leaf-biofilms of macrophytes at initial (a) and at 13 day (b). 16

ACCEPTED MANUSCRIPT 423

Table S1. Elements compositions on leaf surface of submerged macrophytes (%)

424

5. References

425 426 427

Barker, T., Hatton, K., O'Connor, M., Connor, L., Moss, B., 2008. Effects of nitrate load on submerged

428 429

Bialowiec, A., Randerson, P.F., Kopik, M., 2010. Using fractal geometry to determine phytotoxicity of

430 431 432

Cai, H.M., Peng, C.Y., Chen, J., Hou, R.Y., Gao, H.J., Wan, X.C., 2014. X-ray photoelectron

433 434 435

Cao, T., Leyi, N.I., Xie, P., Jun, X.U., Zhang, M., 2011. Effects of moderate ammonium enrichment on

436 437 438

Drake, L.A., Dobbs, F.C., Zimmerman, R.C., 2003. Effects of epiphyte load on optical properties and

439 440

Fang, J.Y., Wan, X.C., 2008. XPS analysis of tea plant leaf and root surface. Spectrosc. Spect. Anal. 28

441 442

Gao, J., Li, L., Hu, Z., Zhu, S., Zhang, R., Xiong, Z., 2015. Ammonia stress on the carbon metabolism

443 444 445

Genet, M.J., Jacques, C., Mozes, N., Hove, C.V., Lejeune, A., Rouxhet, P.G., 2002. Use of differential

446 447 448

Gunes, A., Inal, A., Bagci, E.G., Coban, S., Sahin, O., 2007. Silicon increases boron tolerance and

449 450

James, C., Fisher, J., Russell, V., Collings, S., Moss, B., 2005. Nitrate availability and hydrophyte

451 452

Jones, J.I., Hardwick, E.K., 2000. The influence of periphyton on boundary layer conditions: a pH

453 454 455

Jones, J.I., Young, J.O., Eaton, J.W., Moss, B., 2002. The influence of nutrient loading, dissolved

plant biomass and species richness: results of a mesocosm experiment. Fund. Appl. Limnol. 173 (2),

RI PT

89-100.

landfill leachate on willow. Chemosphere 79 (5), 534-540.

spectroscopy surface analysis of fluoride stress in tea ( Camellia sinensis (L.) O. Kuntze) leaves. J.

SC

Fluorine Chem. 158 (158), 11-15.

1620–1629.

M AN U

three submersed macrophytes under contrasting light availability. Freshwater Biol. 56 (8),

photosynthetic potential of the seagrasses Thalassia testudinum Banks ex Konig and Zostera marina L. Limnol. Oceanogr. 48 (1-2), 456-463.

TE D

(9) , 2196-2200.

of Ceratophyllum demersum. Environ. Toxicol. Chem. 34 (4), 843-849.

AC C

601-606.

EP

charging to analyse microscopic specimens of Azolla leaves by XPS. Surf. Interface Anal. 33 (7),

reduces oxidative damage of wheat grown in soil with excess boron. Biol. Plantarum 51 (3), 571-574.

species richness in shallow lakes. Freshwater Biol. 50 (6), 1049-1063.

microelectrode investigation. Aquat. Bot. 67 (3), 191-206.

inorganic carbon and higher trophic levels on the interaction between submerged plants and periphyton. J. Ecol. 90 (1), 12-24.

17

ACCEPTED MANUSCRIPT 456

Kontturi, E.J., 2005. Surface chemistry of cellulose:from natural fibres to model surfaces.

457 458

Li, C.L., Ni, D.J., 2009. Effect of fluoride on chemical constituents of tea leaves. Fluoride 42 (3),

459 460 461

Liang, M., Wang, T., Li, Y., Yang, P., 2013. Structural and fractal characteristics of biofilm attached on

462 463 464 465

Lyche-Solheim, A., Feld, C.K., Birk, S., Phillips, G., Carvalho, L., Morabito, G., Mischke, U., Willby,

466 467

Mandelbrot, B., 1967. How long is the coast of britain? Statistical self-similarity and fractional

468 469

Matuana, L.M., Balatinecz, J.J., Rns, S., Park, C.B., 2001. Surface characterization of esterified

470 471 472

Meng, Z., Cao, T., Ni, L., Ping, X., Li, Z.Q., 2010. Carbon, nitrogen and antioxidant enzyme responses

473 474

Michael, T.S., Shin, H.W., Hanna, R., Spafford, D.C., 2008. A review of epiphyte community

475 476 477

Nuttens, A., Chatellier, S., Devin, S., Guignard, C., Lenouvel, A., Gross, E.M., 2016. Does nitrate

478

Peitgen, H.O., Jurgens, H., Saupe, D., 2004. Chaos and Fractals. SpringerVerlag.

479 480 481

Peterson, B.J., Frankovich, T.A., Zieman, J.C., 2007. Response of seagrass epiphyte loading to field

482 483

Phillips, G., Willby, N., Moss, B., 2016. Submerged macrophyte decline in shallow lakes: What have

484 485

Raoufi, D., Fallah, H.R., Kiasatpour, A., Rozatian, A.S.H., 2008. Multifractal analysis of ITO thin films

486 487

Raun, A.L., Borum, J., Sand-Jensen, K., 2010. Influence of sediment organic enrichment and water

488 489

Sinn, G., Reiterer, A., Stanzl-Tschegg, S.E., 2001. Surface analysis of different wood species using

237-243.

the surfaces of aquatic plants and gravels in the rivers and lakes reusing reclaimed wastewater.

RI PT

Environ. Earth Sci. 70 (5), 2319-2333.

N., Søndergaard, M., Hellsten, S., 2013. Ecological status assessment of European lakes: a comparison of metrics for phytoplankton, macrophytes, benthic invertebrates and fish.

SC

Hydrobiologia 704 (1), 57-74.

M AN U

dimension. Science 156 (3775), 636.

cellulosic fibers by XPS and FTIR spectroscopy. Wood Sci. Technol. 35 (3), 191-201.

of Potamogeton crispus to both low light and high nutrient stresses. Environ. Exp. Bot. 68 (1), 44-50.

TE D

development: Surface interactions and settlement on seagrass. J. Environ. Biol. 29 (4), 629-638.

co-pollution affect biological responses of an aquatic plant to two common herbicides? Aquat.

EP

Toxicol. 177, 355-364.

AC C

manipulations of fertilization, gastropod grazing and leaf turnover rates. J. Exp. Mar. Biol. Ecol. 349 (1), 61-72.

we learnt in the last forty years? Aquat. Bot. 135, 37-45.

prepared by electron beam deposition method. Appl. Surf. Sci. 254 (7), 2168-2173.

alkalinity on growth of aquatic isoetid and elodeid plants. Freshwater Biol. 55 (9), 1891-1904.

X-ray photoelectron spectroscopy (XPS). J. Mater. Sci. 36 (19), 4673-4680. 18

ACCEPTED MANUSCRIPT 490 491 492

Song, Y.Z., Wang, J.Q., Gao, Y.X., Xie, X.J., 2015. The physiological responses of Vallisneria natans

493 494 495

Stevens, P.J.G., Baker, E.A., 1987. Factors affecting the foliar absorption and redistribution of

496 497

Toth, V.R., 2013. The effect of periphyton on the light environment and production of Potamogeton

498 499 500

Wang, C., Zhang, S.H., Li, W., Wang, P.F., Li, L., 2011. Nitric oxide supplementation alleviates

501 502 503

Wang, C., Zhang, S.H., Wang, P.F., Hou, J., Li, W., Zhang, W.J., 2008. Metabolic adaptations to

504 505

Wang, C., Zhang, S.H., Wang, P.F., Li, W., Lu, J., 2010. Effects of ammonium on the antioxidative

506 507

Wepasnick, K.A., Smith, B.A., Bitter, J.L., Fairbrother, D.H., 2010. Chemical and structural

508 509 510

Wersal, R.M., Madsen, J.D., 2011. Influences of water column nutrient loading on growth

511 512 513

Wickens, D., Lynch, S., West, G., Kelly, P., Verran, J., Whitehead, K.A., 2014. Quantifying the pattern

514 515

Xiao, S., 2014. The Relationship Between Submerged Plant and Eutrophic Waters Recovery. J. Green

516 517 518

Xie, D., Yu, D., You, W.H., Wang, L.G., 2013. Algae mediate submerged macrophyte response to

519 520 521

Yuan, G., Cao, T., Fu, H., Ni, L., Zhang, X., Li, W., Song, X., Xie, P., Jeppesen, E., 2013. Linking

522 523 524

Zhang, M., Cao, T., Ni, L., Xie, P., Li, Z., 2010a. Carbon, nitrogen and antioxidant enzyme responses

to epiphytic algae with the increase of N and P concentrations in water bodies. Environ. Sci. Pollut. Res. Int. 22 (11), 8480-8487.

pesticides. 1. Properties of leaf surfaces and their interactions with spray droplets. Pest Manage. Sci.

RI PT

19 (4), 265-281.

perfoliatus L. in the mesotrophic basin of Lake Balaton. Aquat. Sci. 75 (4), 523-534.

ammonium toxicity in the submerged macrophyte Hydrilla verticillata (L.f.) Royle. Ecotox. Environ.

SC

Safe. 74 (1), 67-73.

Hara. Aquat. Toxicol. 87 (2), 88.

M AN U

ammonia-induced oxidative stress in leaves of the submerged macrophyte Vallisneria natans (Lour.)

response in Hydrilla verticillata (L.f.) Royle plants. Ecotox. Environ. Safe. 73 (2), 189-195.

TE D

characterization of carbon nanotube surfaces. Anal. Bioanal. Chem. 396 (3), 1003-1014.

characteristics of the invasive aquatic macrophyte Myriophyllum aquaticum (Vell.) Verdc. Hydrobiologia 665 (1), 93-105.

EP

of microbial cell dispersion, density and clustering on surfaces of differing chemistries and

AC C

topographies using multifractal analysis. J. Microbiol. Meth. 104, 101-108.

Sci. Technol.

nutrient and dissolved inorganic carbon loading: A mesocosm study on different species. Chemosphere 93 (7), 1301-1308.

carbon and nitrogen metabolism to depth distribution of submersed macrophytes using high ammonium dosing tests and a lake survey. Freshwater Biol. 58 (12), 2532-2540.

of Potamogeton crispus to both low light and high nutrient stresses. Environ. Exp. Bot. 68 (1), 44-50. 19

ACCEPTED MANUSCRIPT 525 526 527

Zhang, S., Pang, S., Wang, P., Wang, C., Guo, C., Addo, F.G., Li, Y., 2016. Responses of bacterial

528 529 530

Zhang, Y.H., Bai, B.F., Chen, J.B., Shen, C.Y., Li, J.Q., 2010b. Multifractal analysis of fracture

community structure and denitrifying bacteria in biofilm to submerged macrophytes and nitrate. Sci. Rep-uk. 6.

morphology of poly(ethylene-co-vinyl acetate)/carbon black conductive composite. Appl. Surf. Sci. 256 (23), 7151-7155.

AC C

EP

TE D

M AN U

SC

RI PT

531

20

ACCEPTED MANUSCRIPT Figure captions Figure 1. SEM image (a-d) and biomass (e and f) of biofilms attached to leaves of V. asiatica (a, b and e) and H. verticillata (c, d and f) treated without (a and c) or

extracellular polymers;②, diatoms;③, crystal.

RI PT

with ammonia nitrogen (10 mg L-1, b and d). ① , organic matter or

Figure 2. Multifractal spectra f(α) of leaf-biofilm of Vallisneria asiatica (a) and Hydrilla verticillata (b) treated with/without 10 mg L-1 NH4+-N.

SC

Figure 3. Surface XPS spectra of Binding energy-Intensity on Vallisneria asiatica (a

mg L-1 NH4+-N (b and d).

M AN U

and b) and Hydrilla verticillata (c and d) treated without (a and c) or with 10

Figure 4. High-resolution XPS C1s spectrum of Binding energy-Intensity on from Vallisneria asiatica (a and b) and Hydrilla verticillata (c and d) treated without (a and c) or with 10 mg L-1 NH4+-N (b and d).

TE D

Figure 5. High-resolution XPS O 1s spectrum of Binding energy-Intensity on from Vallisneria asiatica (a and b) and Hydrilla verticillata (c and d) treated without (a and c) or with 10 mg L-1 NH4+-N (b and d).

AC C

13th day

EP

Figure 6. Changes of different concentrations of NH4+-N on antioxidant system at

ACCEPTED MANUSCRIPT

RI PT

Table 1. Multifractal parameters of leaf surface from two submersed macrophytes. Plants treatments αmin αmax ݂(αmin) ݂(αmax) △α △݂ Control 1.63 3.22 1.06 0.26 1.59 0.8 -1 V. asiatica 10 mg L 1.87 2.18 1.61 1.31 0.31 0.3 NH4-N Control 1.78 3.51 1.47 0.1 1.73 1.37 H. -1 10 mg L verticillata 1.63 2.91 1.18 0.49 1.28 0.69 NH4-N

AC C

EP

TE D

M AN U

SC

Table 2. Elements compositions on leaf surface of submersed macrophytes (%) V. asiatica H. verticillata Element -1 Control 10 mg L NH4-N Control 10 mg L-1 NH4-N C 70.42±0.11 74.71±0.29* 78.45±0.42 83.66±0.77* O 22.92±0.27 16.11±1.24* 15.53±0.65 11.9±1.21* N 5.64±0.42 7.99±1.17* 5.52±0.51 4.11±0.62* P 0.21±0.02 0.26±0.03 0.21±0.12 0.33±0.09 Cl 0.26±0.03 0.30±0.06 \ \ K 0.55±0.05 0.60±0.02 \ \ O/C 0.33 0.22 0.2 0.14 N/C 0.08 0.11 0.07 0.05 Note: "\" in the table indicates that the element content is below the detection limit * indicate that mean values are significantly different between NH4-N treatments and controls for the same plant (p<0.05).

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 1

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 2.

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 3.

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 4

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 5

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Figure 6

ACCEPTED MANUSCRIPT Highlights

EP

TE D

M AN U

SC

RI PT

High ammonium enhanced biofilm growth Biofilm growth altered the topography leaf surface High ammonium altered chemical composition of leaf surface Biofilm growth might contribute to ammonium-induced toxicity

AC C