Microcystin-LR binds iron and iron promotes self-assembly

The microcystin producing Microcystis aeruginosa PCC 7806 and its close strain, the ... with microcystin implicated in the fitness of the photosynthet...
0 downloads 0 Views 3MB Size
Subscriber access provided by MacEwan University Libraries

Article

Microcystin-LR binds iron and iron promotes self-assembly Laura Ceballos-Laita, Carlos Marcuello, Anabel Lostao, Laura Calvo-Begueria, Adrián Velázquez-Campoy, María Teresa Bes, María Francisca Fillat, and María Luisa Peleato Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b05939 • Publication Date (Web): 03 Apr 2017 Downloaded from http://pubs.acs.org on April 5, 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.

Environmental Science & Technology 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 32

Environmental Science & Technology

Microcystin-LR

No metal

Fe3+

3  

1  

4   6  

2   Monomers and dimers

oligomers

ACS Paragon Plus Environment

Environmental Science & Technology

1

Page 2 of 32

Microcystin-LR binds iron and iron promotes self-assembly

2 3

Laura Ceballos-Laita1,3#, Carlos Marcuello2#, Anabel Lostao2,4, Laura Calvo-

4

Begueria1,3, Adrián Velazquez-Campoy1,3,4, María Teresa Bes1,3, María F. Fillat1,3 and

5

María-Luisa Peleato1,3

6

1

7

Microscopías Avanzadas, Instituto de Nanociencia de Aragón. Universidad de

8

Zaragoza. (Spain). 3Institute for Biocomputation and Physics of Complex Systems

9

(BIFI)-Joint Unit BIFI-IQFR (CSIC). Universidad de Zaragoza (Spain).4 Fundación

10

*

Biochemistry Department, Universidad de Zaragoza, (Spain).

2

Laboratorio de

ARAID, Aragón, Spain.

11 12

*Corresponding author:

13

María Luisa Peleato

14

Departamento de Bioquímica

15

Facultad de Ciencias

16

Pedro Cerbuna 12

17

50009-Zaragoza, Spain

18

Tf:+34976762479

19

[email protected]

20 21

#Both authors contributed equally to this work

22

ACS Paragon Plus Environment

1

Page 3 of 32

Environmental Science & Technology

23

Abstract

24 25

The microcystin producing Microcystis aeruginosa PCC 7806 and its close strain, the

26

non-producing Microcystis aeruginosa PCC 7005 grow similarly in the presence of 17

27

µM iron. Under severe iron deficient conditions (0.05 µM), the toxigenic strain grows

28

slightly less than in iron-replete conditions, while the non-producing microcystin strain

29

is not able to grow. Isothermal titration calorimetry performed at cyanobacterial cytosol

30

or meaningful environmental pHs values shows a microcystin-LR Kd for Fe2+ and Fe3+

31

of 2.4 µM. Using Atomic Force Microscopy, 40% of microcystin-LR dimers were

32

observed, and the presence of iron promoted its oligomerization up to 6 units.

33

Microcystin-LR binds also Mo6+, Cu2+ and Mn2+. Polymeric microcystin binding iron

34

may be related with a toxic cell colony advantage, providing enhanced iron

35

bioavailability and perhaps affecting the structure of the gelatinous sheath. Inside cells,

36

with microcystin implicated in the fitness of the photosynthetic machinery under stress

37

conditions, the toxin would be involved in avoiding metal-dependent Fenton reactions

38

when photooxidation causes disassembly of the iron-rich photosystems. Additionally, it

39

could be hypothesized that polymerization/depolymerization dynamics may be an

40

additional signal that could trigger changes, for example in the binding of microcystin

41

to proteins.

42 43

Key words: Iron; Microcystin; Microcystis;

44

ACS Paragon Plus Environment

2

Environmental Science & Technology

Page 4 of 32

45 46

1. Introduction

47 48

Microcystins are potent cyanotoxins that inhibit eukaryotic protein phosphatases 1 and

49

2A, and produce adverse effects in water quality and potential human and animal

50

poisoning1. The toxic strains producing microcystins deviate a considerable amount of

51

carbon and nitrogen to the synthesis of such secondary metabolites. For this reason it

52

seems that microcystins must confer adaptative advantages to producer species.

53

However, a coherent global model of its role in the cyanobacteria is not yet well

54

established. Proteomic studies pointed to a different expression profile in toxic and non-

55

toxic cells, affecting many proteins involved in photosynthesis, nitrogen and carbon

56

metabolism, and redox control1,2. Microcystin was usually considered an endotoxin.

57

Supplying radioactive inorganic carbon to M. aeruginosa cultures and following the fate

58

of the intracellular radioactive microcystin pool, no export from the intracellular

59

microcystin pool was observed3. On the other hand, the mcyH gene encoding an ABC

60

transporter is essential for microcystin synthesis4, suggesting microcystin exportation

61

out of the cell. In congruence with the toxin exportation, solid evidences linking

62

microcystin presence and enhanced colony formation and size have been reported5, 6, 7.

63

Cell-wall proteins could be implicated in the effect microcystin has on colony

64

formation8,9, likely involving the microcystin-related protein C (MrpC)9. MrpC

65

accumulates in microcystin-deficient cells, suggesting a role of microcystin in colony

66

arrangement. Microcystin also causes upregulation of genes involved in polysaccharide

67

synthesis7, involved in colony formation.

68

Microcystin has been proposed to have allelopathic effects against other members of the

69

ecosystem, but few data indicate harmful effects of microcystins at concentrations that

70

are typical for the environment10. As a defence mechanism to microfauna, phylogenetic

71

analysis suggested that the microcystin synthetase predated the metazoan lineage, thus

72

dismissing the possibility that microcystins emerged as a means of defence against

73

grazing11. Also, it has been suggested that microcystin may be involved in quorum

74

sensing or act as a signalling molecule12. However, several studies have failed to find

75

evidence for such a role13, 4.

76

Regarding a possible physiological role inside the cell, even though microcystins have

77

been involved in causing oxidative stress as an additional toxicity mechanism, an

ACS Paragon Plus Environment

3

Page 5 of 32

Environmental Science & Technology

78

interesting piece of work proposes that the toxin increases the fitness of the toxic

79

Microcystis strain under oxidative stress conditions14. Under stresses such as high light

80

or carbon deficiency, microcystin synthesis becomes advantageous15,16. Improvement of

81

competition for light has been also considered as a potential advantage provided by

82

microcystin, but competition experiments with toxic and non-toxic Microcystis strains

83

showed that non-toxic strains were better competitors for light than toxic strains17.

84

Moreover, electron microscopy of immuno-gold labelled Microcystis cells indicates that

85

two thirds of the intracellular microcystin is located around the thylakoids18, 19, 20. A role

86

in the maintenance of the photosynthetic machinery, protecting against photoinhibition

87

has been proposed21, in relation to oxidative stress protection14.

88 89

Microcystins have been described to bind to many proteins, sometimes in a bulk

90

manner14,22,23. It has been proposed that there is a covalent interaction of the toxin with

91

cysteine residues of proteins14, and in this way, an interesting proposal is that

92

microcystin competes for binding sites in thioredoxin-regulated proteins during

93

oxidative stress (low Fe)24. Lately, eukaryotic proteasome has been described as a target

94

for microcystin, affecting its trypsin-like activity, and delaying protein degradation25.

95 96

The microcystin synthesis has been linked to iron metabolism for many years. Lyck and

97

colleagues26 showed that during iron depletion, toxic strains of Microcystis maintained

98

cell vitality much longer than the non-toxic strains. Moreover, Utkilen and Gjolme27

99

found that toxic strains exhibited higher rates of iron uptake than non-toxic strains.

100

They proposed that the microcystin is an intracellular chelator of Fe2+, as well as

101

predicted that the synthesis of the toxin would be controlled by the amount of free iron

102

present in the cells. Structural similarities between microcystin and bacterial

103

siderophores28 lead also to propose a putative role as an extracellular iron-scavenging

104

molecule.

105

Another relevant aspect linking iron and microcystin is the regulation of the microcystin

106

gene cluster. The mcy operon, responsible for microcystin synthesis seems to be

107

regulated by Fur (ferric uptake regulator)29, an iron-dependent transcriptional regulator

108

of a wide number of genes involved in iron metabolism and oxidative stress30. As a

109

consequence, iron deficiency leads to an increased amount of microcystin27,31. Since a

110

long time ago, several metal-microcystin complexes have been described32,33, and Klein

111

and co-workers34 provided evidence for the formation of at least two Fe3+-microcystin-

ACS Paragon Plus Environment

4

Environmental Science & Technology

Page 6 of 32

112

LR complexes, at acidic pHs.

113

Previous results 32-34 formed the basis for examining microcystin-metal binding by ITC.

114

As an extension, the potential influence on colony formation of microcystin-LR in the

115

presence of iron and other metals through a hypothesised oligomerization of

116

microcystin-metal complexes was examined..

117 118

2. Experimental

119

2.1. Strains and culture conditions.

120

This study was carried out with M. aeruginosa PCC 7806 and PCC 7005 provided by

121

the Pasteur Culture Collection. Cells were photoautotrophically grown in BG11

122

medium supplemented with 2 mM NaNO3 and 10 mM NaHCO3, as described

123

previously by Rippka35, and as recommended by the Pasteur Institute. The culture

124

medium was air bubbled at 25oC with continuous illumination (25 µmol of

125

photon/m2.s). Light was measured using a Quantum Sensor photometer (Skye

126

Instruments, SKP 200). Iron-deficient (0.5 µM iron) and iron-replete cultures were

127

started from an equal aliquot from an iron-replete culture in late exponential phase.

128

After washing two times with iron-deficient media or iron-replete medium respectively,

129

cells were inoculated and grown in batch conditions (1 L Roux flasks with initial

130

Abs700nm= 0.15). This experiment was performed 3 times, and measurements were done

131

in duplicated samples.

132 133

2.2. Cell counting

134

The cell density was determined daily both spectrophotometrically and by counting

135

cells with a Neubauer cell. To avoid cell aggregation, the samples were treated with

136

KOH following described methodology36.

137 138

2.3. Analytical methods

139

Samples of 1 mL and 5 mL were collected for chlorophyll a and protein determination,

140

respectively. Chlorophyll a was quantified according to Mackinney37 and protein

141

contents were estimated by the bicinchoninic acid method (BCATM Protein Assay

ACS Paragon Plus Environment

5

Page 7 of 32

Environmental Science & Technology

142

Reagent Kit from Pierce). Microcystin-LR was quantified using MicroCystest®

143

(Zeulab). HPLC procedure is indicated in Supporting Information section.

144 145

2.4. Isothermal titration calorimetry (ITC)

146

Metal binding to microcystin-LR was assessed in a high-sensitivity isothermal titration

147

calorimeter Auto-iTC200 (MicroCal, GE Healthcare). Assays were performed with a

148

final concentration of 150 µM metal and 10 µM microcystin-LR. For Fe3+ binding

149

assays, a stock solution of Fe-NTA (nitrilotriacetic acid) was prepared according to

150

previously described procedure38. The solution was made anoxic by purging with

151

nitrogen and stored in a sealed vial until needed39. Experiments were carried out at 25°C

152

in 150 mM sodium phosphate (pH 7), 100 mM Hepes (pH 7.6) or 50 mM Tris-HCl (pH

153

8, 8.5) buffers, as indicated in each experiment. Assays with Fe2+ were performed with

154

fresh made solutions in the presence of 1.5 mM ascorbate for maintaining reducing

155

conditions. Microcystin-LR solution (10 µM) in the calorimetric cell was titrated with

156

metal solution (150 µM). Control experiments (either dilution of microcystin-LR or

157

dilution of metal ions, Fe2+, Fe3+, Mo6+, Mn2+, Cu2+, Ca2+. Zn2+, Co2+, Ni2+) were

158

performed under the same experimental conditions. The heat evolved after each ligand

159

injection was obtained from the integrated calorimetric signal. The heat due to the

160

binding reaction was obtained as the difference between the reaction heat and the

161

corresponding heat of dilution, the latter estimated as a constant heat throughout the

162

experiment, and included as an adjustable parameter in the analysis. The dissociation

163

constant (Kd) and the interaction enthalpy (∆H) were obtained through non-linear

164

regression of experimental data to a model considering a single ligand-binding site. A

165

brief description of the procedure has been included in Supporting Information. Data

166

were analyzed using software developed in our laboratory implemented in Origin 7

167

(OriginLab).

168 169

2.5. Cross-Linking of Microcystin-LR

170

1-Ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC) is a zero-length crosslinker

171

agent widely used to reveal protein-protein specific interactions. It is generally used as a

172

carboxyl activating agent for the coupling of primary amines to yield amide bonds.

173

Microcystin-LR cross-linking reactions (30 µL 1 mg/mL) were performed in the

ACS Paragon Plus Environment

6

Environmental Science & Technology

Page 8 of 32

174

presence of 2.7 mM FeCl2, or FeCl3. The mixture was incubated with 10 mM 1-Ethyl-3-

175

(3-dimethylaminopropyl)carbodiimide, EDC, (freshly dissolved in H2O) and 10 mM

176

Sulfo-NHS (N-hydroxysulfosuccinimide) in 50 mM acetic-acetate buffer pH 5.5, for 30

177

min at 25 ºC. Adding 20 mM 2-mercaptoethanol stopped the reaction. Other metals

178

tested (Mn2+, Zn2+, Cu2+) were used at the same concentration as iron.

179 180

2.6. Atomic force microscopy imaging

181

Sample-nanoflat surface immobilization is a vital step to carry out atomic force

182

microscopy (AFM) measurements without dragging biomolecules through the tip

183

movement on the sample. 1 cm2 cleaved muscovite mica pieces (Electron Microscopy

184

Sciences) were incubated with 1 µg/mL of microcystin-LR samples, for 10 min at room

185

temperature. The surfaces were washed three times using the solvent to remove not well

186

attached molecules, introduced in a liquid cell and covered with the same washing

187

solution ready for measurements. Microcystin-LR molecules in the different

188

experimental conditions were measured in different solvents, 20 % methanol (v/v),

189

milli-Q water and PBS, pH 7.4. The immobilization of biomolecules on mica was

190

previously evaluated observing clearly they preserve their functionality40.

191

AFM measurements were performed with a Cervantes Fullmode SPM (Nanotec

192

Electrónica S.L, Spain). AFM images were taken using the force based Jumping Mode

193

under very small applied forces barely touching the molecules in a non-intrusive

194

manner41. V-shaped silicon nitride cantilevers with spring constants from 0.01 to 0.03

195

N/m (Bruker Probes, MSNL lever Probes) calibrated using the thermal noise method

196

were used. Image processing was performed with the WSxM software42. Percentages

197

corresponding for the different oligomeric features were calculated taking several

198

sample areas and zooming with the WSxM functions to identify the monomeric units in

199

the features without losing data as reported43. The quantification was carried out by

200

counting a minimum of 250 oligomers/species from at least 10 different locations of

201

each sample. At the same time, each type of sample was measured –at least- by

202

triplicate to ensure reproducibility. Each associate was analyzed in detail using the

203

zoom and the profile functions of the WSxM program.

204

The error was calculated from the results dispersion analysis of different AFM images

205

corresponding to several areas of the sample.

206

ACS Paragon Plus Environment

7

Page 9 of 32

Environmental Science & Technology

207 208 209

3. Results

210

3.1. A toxic Microcystis endures iron deficiency more effectively than a non-toxic strain

211

Toxic and non-toxic M. aeruginosa (PCC7806 and PCC7005 respectively) were

212

transferred from an iron-replete medium in middle exponential phase (samples were

213

washed twice) to an iron-deficient or iron-replete BG11 medium. The consequences of

214

iron depletion in toxic (PCC 7806) and non-toxic (PCC 7005) M. aeruginosa strains

215

were much more evident in the non-toxic strain. Figure 1 shows that growth of cells is

216

quite different (number of cells is shown in Fig 1A, total protein in cells is shown in

217

Figure 1B, and chlorophyll in Figure 1C). The toxic strain was virtually unaffected by

218

the low iron availability in the culture medium while the growth of the non-toxic cells is

219

almost zero when iron is limited (Fig.1A and 1B). Microcystin-containing cells were

220

only slightly affected in chlorophyll content, while the non-toxic strain exhibited a very

221

low level of the pigment (Figure 1C) under the same conditions, due to the fact that iron

222

deficiency impairs chlorophyll biosynthesis, causing bleaching. Our results indicated

223

that the toxic strain, M. aeruginosa PCC7806, has some kind of advantage either

224

concerning intracellular iron storage or a more efficient iron uptake. Under iron-replete

225

conditions, the non-toxic strain, PCC7005 performed slightly better (Fig.1A and 1B).

226 227

3.2. Microcystin-LR binds iron and other metals

228

The ability of the microcystin-producing strain to overcome iron deficiency led us to

229

investigate the binding activity microcystin-LR-iron. We were especially interested in

230

measurements at physiological and environmental meaningful pHs, for instance pH of

231

cytosol or pH of alkaline environments, with low iron availability. A different

232

experimental approach than the previous described in the literature32-34,44 was chosen:

233

isothermal titration calorimetry (ITC). In order to investigate if microcystin-LR plays a

234

role by interacting with metals, assays were performed with several metal ions, Fe2+ and

235

Fe3+ in particular, and binding dissociation constants were measured. According to these

236

assays, microcystin-LR binds Fe2+, Fe3+, Mo6+ (Figure 2), Cu2+ and Mn2+ (Table 1); on

237

the other hand, in our experimental conditions microcystin-LR does not specifically

238

interact with Ca2+, Zn2+, Co2+ and Ni2+ (no interaction was observed in our calorimetric

ACS Paragon Plus Environment

8

Environmental Science & Technology

Page 10 of 32

239

experiments; however, although we could observe a weak interaction of microcystin-LR

240

with Zn2+, it was too weak to report it). Titrations were performed at different pHs (7.0,

241

7.6, 8.0 and 8.5), and the higher affinity for iron was found at pH 7. Microcystin-LR

242

exhibits a similar Kd, 2.4 µM, when binding Fe3+ and Fe2+; however, a more negative

243

enthalpy in the case of Fe2+ would suggest a more specific interaction with the reduced

244

metal form. Other metals interacting with microcystin-LR exhibited dissociation

245

constants in the low-to-moderate micromolar range (Table 1). The interaction of other

246

metals at different pHs (e.g., Cu, Mn, Mo) has been measured. In the case of Cu and Mn

247

it is observed that the interaction affinity decreases with the pH, while for Mo it is

248

observed that the interaction affinity increases with pH. A complete set of all the data of

249

the interactions observed by calorimetry is shown in Supporting Information (Figure

250

1S).

251 252

3.3. Microcystin-LR oligomerizes, being Fe3+ the main promoter of chain formation

253

After confirming that microcystin-LR interacts with iron and other metals, it was

254

considered interesting to study the complex, and HPLC of the samples were performed.

255

In order to study changes in microcystin-LR in the presence of iron, studies of

256

chromatographic mobility of the different samples were carried out. Figure 2S in

257

Supporting Information section shows HPLC reverse phase chromatograms of cross-

258

linked pure microcystin-LR in the presence and absence of Fe3+, after subtraction of the

259

EDC-NHS baseline. A single peak was obtained when microcystin-LR and EDC-NHS

260

cross-linked microcystin-LR were run, whereas several peaks and a shoulder in the

261

main peak were observed when microcystin-LR was EDC-NHS treated in the presence

262

of Fe3+. These results suggest that different species of microcystin-LR (with different

263

hydrophobicity) were generated in the presence of iron.

264 265

3.4. Oligomerization patterns analyzed by AFM imaging

266

The different molecular species of microcystin-LR observed by HPLC was studied

267

using an experimental approach that allows visualizing single molecules: Atomic Force

268

Microscopy Imaging (AFM). This technique would confirm unequivocally the presence

269

of microcystin-LR oligomers, and ascertain the influence of metals on microcystin-LR

270

oligomerization. The effect of incubating microcystin-LR with different metal cations in

271

20 % methanol was investigated and the results of the cross-linked samples are

ACS Paragon Plus Environment

9

Page 11 of 32

Environmental Science & Technology

272

summarized in Table 2. A clear set of images were obtained for all the measurements,

273

and despite the small size of the microcystin-LR molecules, the species were

274

distinguishable at the single molecule level. Control experiments were also performed

275

in milli-Q water and PBS buffer, pH 7.4 (Table 2). The results indicated that

276

microcystin-LR appears mainly as a monomer, and only a maximum of 10 % of the

277

molecules are dimerized independently of the three liquid media used (Figure 3, image

278

A).

279

The carbodiimide used (EDC) is a zero-length cross-linker; it causes direct conjugation

280

of closely positioned carboxylates (–COOH) to primary amines (–NH2) without

281

becoming part of the final cross-link (amide bond) between such interacting target

282

molecules. EDC showed the dimerization of a portion of microcystin-LR molecules,

283

with an average of about 40 % (Figure 3, image D). The height found for microcystin-

284

LR molecules in the samples with EDC is a bit higher due to the bound cross-linker, 2.0

285

nm (Figure 3, images D, E, F) with respect to 1.0 nm on average for the untagged

286

molecules (Figure 3, images A, B, C). This result can be corroborated using the Z-

287

height profile function on each feature (Figure 4, images B, C, E, F). Oligomerization

288

develops in the same plane through the building of chains, and no other type of

289

molecular architecture was detected. The zoom function made possible to unequivocally

290

attribute the association pattern of each oligomeric chain. It is important to remark that

291

the measurement in the Z-height (height profiles) presents a high accuracy, with

292

uncertainty lower than 1 nm, whereas the XY size of the feature, the width, is less

293

accurate due to the well documented broadening effect related to the AFM tip dilation45.

294

Despite this broadening effect, XY width can be compared in relative terms and it can

295

be observed that dimers have approx. twice the width of the monomer features (Figure

296

3, images B, E).

297

The presence of Fe2+ ions during the cross-linking process induces the addition of a

298

third molecule to the dimers to form trimers up to 21 % referred to total peptide (Figure

299

4, images A, B, C). The maximum degree in the oligomerization process was achieved

300

when Fe3+ was added. This agent induces the formation of chains with six microcystin-

301

LR units (Figure 4, images G, H, I). Visualization of AFM topography images shows

302

that a considerable 28% of the analyzed features were hexameric chains. This result

303

corresponds to the fact that a 62 % of the molecules are building these chains, whereas

304

no trimeric or tetrameric arrangements were found in that sample.

ACS Paragon Plus Environment

10

Environmental Science & Technology

Page 12 of 32

305

Other divalent metals such as Mn2+ or Zn2+ used in the cross-linking procedure resulted

306

in no polymerization of microcystin-LR and even inhibition of dimerization. Strands of

307

microcystin-LR with a maximum of four molecules appeared when Cu2+ was present

308

during the cross-linking (Figure 4, images D, E, F). The percentage of these longer

309

chains reaches 10 % considering features and rises up 26 % if the data is translated to

310

total peptide.

311 312 313

4. Discussion

314

The selective pressure of iron limitation has led to sophisticated systems to overcome

315

the deprivation of such essential element46 endowing cyanobacteria to compete very

316

favourably for iron or adapt to its deficiency. As part of these adaptive mechanisms,

317

microcystin was proposed long time ago to provide a selective advantage to toxin-

318

producer cells under iron deficient conditions27. Over time, many other functions for

319

microcystins have been proposed, some of them clearly demonstrated, such as its role in

320

colony formation5-7 or in protection against oxidative stress14,47 and protection to

321

photoxidation16,17,49. Besides, many data in the literature pointed also to a function

322

related with iron metabolism26,27,33-34,44. In this work, we show that using iron-limited

323

BG11 medium, the toxic strain has overcome iron deficiency without serious influence

324

on its growth, while the non-toxic strain is affected with severely reduced growth

325

(Figure 1A and 1B). Fujii et al.48 described that the siderophore-independent iron uptake

326

rates were similar in the toxic strain PCC 7806 and in a genetically modified strain

327

unable to produce microcystin, but the growth rates measured for the microcystin

328

deficient strain were lower than those of the wild-type strain, even under conditions

329

where similar iron uptake rates were expected48. In addition, previous data pointed that

330

the microcystin-producing strain of M. aeruginosa (CYA 228/1) acquires iron about 2

331

fold faster than the non-toxic strain CYA 4327. Recent reports pointed to the fact that

332

microcystin-producing strains have been shown to outcompete their nontoxic

333

counterparts under iron-limiting conditions47, in good agreement with the data presented

334

in this paper. The ecophysiological advantage of microcystin production under iron-

335

limiting conditions could be justified if microcystin can act as iron-chelator or iron

336

storage molecule. Our data indicate that microcystin-LR binds iron and other cations at

337

physiological or environmentally meaningful pHs. According to calorimetric titrations

ACS Paragon Plus Environment

11

Page 13 of 32

Environmental Science & Technology

338

microcystin-LR specifically interacts with Fe2+, Fe3+, Mo6+, Cu2+ and Mn2+. The

339

stoichiometry of the interaction is 1:1 and the dissociation constants are in the low-to-

340

moderate micromolar range. Interestingly, Fe2+ and Fe3+ display similar dissociation

341

constants (Kd 2.4 µM); however, a more negative enthalpy in the case of Fe2+ is

342

observed. Apparently microcystin-LR does not specifically interact with Ca2+, Zn2+,

343

Co2+ and Ni2+, although unspecific interactions cannot be ruled out since we have

344

observed a weak interaction with Zn2+. Klein et al.34 found electrochemical evidence for

345

the complexation of Fe3+ by microcystin-LR at acidic pHs, but not evidence for Fe2+.

346

However, Saito et al.33 using cryospray ionization Fourier transform ion cyclotron

347

resonance mass spectroscopy (CSI-FTIC-MS) identified a microcystin-Fe2+ complex,

348

but not complex formation of microcystin with Fe3+. On the other hand, Zn2+, Cu2+ and

349

Mg2+ were observed previously to form complexes with microcystin-LR32,33.

350

We have classified the interaction of microcystin with the metal ions assessed in our

351

work as having low-to-moderate binding affinity, in comparison with the standard

352

affinity ranking in biomolecular interactions: low affinity (Kd above 100 µM), moderate

353

affinity (Kd from 10 nM to 100 µM), and high affinity (Kd below 10 nM). Of course,

354

regarding iron chelation, siderophores usually exhibit extremely high affinities (Kd in

355

the range from 10-50 M to 10-30 M)50,51. However, in some cases the binding affinity of

356

the siderophore for iron is not as high as that in the more common siderophores, with

357

reported Kd is in the range from 5 to 300 µM52. Our results are in agreement with some

358

published results, but in disagreement with other ones. Humble et al.32 reported

359

dissociation constants of 10 µM approximately for Cu2+ and Zn2+ binding to

360

microcystin-LR at pH 7.5, 100 mM KCl, with no significant differences with

361

microcystin-LW and microcystin-LF. They concluded that microcystin-LR is a

362

medium-strength ligand for these metal ions. This is a statement equivalent to our

363

classification of microcystin-LR as having low-to-moderate binding affinity towards

364

metal ions. On the other hand, Yan et al.44 measured much smaller dissociation

365

constants for the binding of the same metal ions to microcystin-LR at pH 7 (10 nM for

366

Cu2+ and 0.01 nM for Zn2+). Klein et al.34 reported the formation of a complex between

367

microcystin-LR and Fe3+ following a two-step pathway, with a labile, rapidly forming

368

initial complex followed by a more stable, slowly forming complex with dissociation

369

constant of 0.1 pM in 60% v/v methanol/water, but no complex was observed between

370

microcystin-LR and Fe2+. Importantly, this dissociation constant corresponds to the

ACS Paragon Plus Environment

12

Environmental Science & Technology

Page 14 of 32

371

formation of the more stable complex, while a much larger dissociation constant would

372

correspond to the first, labile complex representing the direct encounter between

373

microcystin-LR and iron, which must be that one observed by ITC.

374

Using AFM we achieved the direct visualization of the molecular arrangements of the

375

samples using “zero-length” cross-linkers, such as EDC. The cross-linker helps to

376

mimic the conditions for the oligomerization in vivo53. The first observation was that the

377

nature of the solvent did not affect the association pathway of microcystin-LR.

378

Regarding the metals, Mn2+ and Zn2+ not only did not promote association but also even

379

seemed to behave as inhibitors of dimerization. Fe2+ ions were able to bind a monomer

380

to a dimer composing a certain portion of trimers, so it could play a role as inductor in

381

the microcystin-LR strand formation. On the other hand, Cu2+ ions seem to promote the

382

formation of tetrameric chains binding dimers among themselves. Other metals as Mn2+

383

binds microcystin-LR but did not promote oligomerization, while Cu2+ induced a

384

considerable amount of trimers and tetramers. Finally, we can see that Fe3+ is a potent

385

inducer of microcystin-LR oligomerization. Ferric ions make a majority of the

386

molecules to form hexameric chains (more than 60%). Figure 3S in Supporting

387

Information summarizes the observed oligomerization in the presence of Fe2+ and Fe3+.

388

These chains show an irregular conformation, and do not appear as rigid nanotubes (Fig.

389

4 images G, H). This effect is favoured by the degree of freedom of the chains provided

390

by the electrostatic adsorption on the mica surface. Microcystin hexamers are not rigid,

391

and certain angles between monomers are observed (Figure 4), maybe related with the

392

flexibility of the arginine side-chain in microcystin-LR54, if arginine is involved in the

393

bond.

394

A possible consequence of the ability of microcystin-LR to form complexes with iron

395

can be related to the toxic cells colony advantage. In microcystin containing colonies,

396

microcystin can act as siderophore, with increased colony size and extracellular

397

polysaccharides7, providing enhanced iron availability versus the environment. Iron-

398

promoted oligomerization has been observed in vitro, but the data leads to speculation

399

regarding a possible role of microcystin-LR chains in the structure of the colony

400

gelatinous sheath. Gan et al.7, found that depletion of extracellular microcystin

401

concentrations caused small colonies. However, the authors indicated that induction of

402

extracellular polysaccharides by microcystin exposure is the major mechanism

403

explaining the increased colony size in presence of the toxin. On the other hand, Amano

ACS Paragon Plus Environment

13

Page 15 of 32

Environmental Science & Technology

404

et al.55 described that added extracellular polysaccharides to M. aeruginosa (UTEX LB

405

2061) did not promote colony formation or aggregation of cells. When the

406

polysaccharides where added with iron 6.8 mg/L, colonies were formed. Unfortunately,

407

the strain used in such experiments does not produce microcystin but the authors infer

408

that the extracellular polysaccharides themselves would not act as binders for M.

409

aeruginosa cells and only when high amount of iron is present colony formation

410

occurs56.

411

Microcystin was inmunolocated around the thylakoids18-20. Many hypotheses may arise,

412

for instance microcystin would prevent metal-dependent Fenton reactions when

413

photooxidation causes disassembly of photosystems and iron-containing photosynthetic

414

components release iron. In the other hand, low iron or high light, as well as other

415

environmental stresses, causes oxidative stress in cyanobacteria46, and the ROS-induced

416

microcystin binding to proteins14 may be related with iron-induced changes in the

417

oligomerization status of the peptide.

418

The role of a polymeric microcystin binding iron in an intracellular scenario may be

419

related

420

conditions16,17,48,56,57. Beyond the consideration of the long-standing idea of microcystin

421

acting as an intracellular siderophore or chelator for storing iron27, it seems clear that

422

microcystin may have a relevant role in the adaptation of photosynthesis to

423

environmental conditions. Several studies pointed to an active role of microcystins in

424

preventing light photoxidation48,56,57, and providing ecophysiological fitness under

425

environmental stresses. Moreover, the localization of exogenous microcystin-LR taken

426

by a non-producing cyanobacterium, Synechocystis PCC6803, located the toxin in the

427

thylakoid membranes, causing a decrease in photosystem II activity58. Phycobiliproteins

428

may be the major proteins that have binding sites for these oligopeptides59, and

429

curiously, a molar ratio near to 6 was observed for microcystins to the phycobilin (αβ)

430

monomer59.

431

Either low iron or high light, as well as other environmental stresses, causes in

432

cyanobacteria oxidative stress46, and the ROS-induced microcystin binding to proteins14

433

may be related with iron-induced changes in the oligomerization status of the peptide.

434

When photooxidation causes disassembly of the iron-rich photosystems, the

435

oligomerization may occur, avoiding metal-dependent Fenton reactions. Also, the large

with

the

fitness

of

the

photosynthetic

machinery

ACS Paragon Plus Environment

under

stress

14

Environmental Science & Technology

Page 16 of 32

436

subunit of RubisCO showed a lower susceptibility to proteases in the presence of

437

microcystin14, maybe keeping up the machinery under stress conditions. Metal-driven

438

oligomerization of microcystin may represent an additional signalling: as iron is

439

released from disassembly of photosynthetic machinery, the equilibrium is shifted

440

towards the oligomerization state, which is achieved through metal binding.

441

Finally, our work would inspire future experiments for elucidating the possible role of

442

the microcystin as an ecophysiological molecule, either as iron concentrator agent or as

443

involved in reversible inhibition of the photosystem II. Microcystin present in the

444

surface of the cells and gelatinous sheaths surrounding colonies may function as a

445

capturing iron agent from the environment, able to concentrate the scarce metal around

446

the colony and playing a role in the structure of the sheath. Microcystin, iron and

447

extracellular polysaccharides should be related with MrpC9, necessary to piece the

448

microcystin-role puzzle. Inside the cells, microcystin would protect photosystem II and

449

photosynthetic machinery against photoxidation, triggered by iron release and

450

subsequent oligomerization. However, to better understand the implications of this

451

study, in the near future is necessary to study the ability of other microcystin variants as

452

well as nodularin to oligomerize in the presence of iron or other metals.

453 454 455

Acknowledgements

456

This work was funded by the Spanish Ministerio de Educación y Ciencia (BFU2006-

457

03454 and BFU2009-07424). L.C.B. was partially supported by a fellowship from the

458

BIFI (Zaragoza, Spain). A.L. and A.V-C thank ARAID for financial support. C.M. is

459

indebted to Gobierno de Aragón for receiving a PhD fellowship. We thank to Laura

460

Vela, Irene Gregorio and María del Mar Pascual for performing initial experiments on

461

microcystin cross-linking; we also thank J. L. Díez and I. Echániz for technical support.

462

The research was partially funded by Gobierno de Aragón and European Social Fund

463

(Grupo Consolidado B-18).

464 465

Supporting Information:

466

-Determination of dissociation constants by isothermal titration calorimetry

ACS Paragon Plus Environment

15

Page 17 of 32

Environmental Science & Technology

467

-Microcystin HPLC analysis

468

-Figure 1S: Calorimetric titrations for the interaction of microcystin LR with different

469

metal ions at several pH values.

470

-Figure 2S: HPLC of cross-linked microcystin-LR.

471

-Figure 3S: Summary of the main oligomeric chain species composed by microcystin-

472

LR under different conditions.

473 474 475 476 477

1.

478

expression in different strains of the bloom-forming cyanobacterium Microcystis

479

aeruginosa. Mol. Cell. Proteomics 2011, 10: M110 003749.

References Alexova, R.; Haynes, P.A.; Ferrari, B.C.; Neilan, B.A. Comparative protein

480 481

2.

Tonietto, A.; Petriz, B.A.; Araujo, W.C., Mehta, A.; Magalhaes, B.S.; Franco,

482

O.L. Comparative proteomics between natural Microcystis isolates with a focus on

483

microcystin synthesis. Proteome. Sci. 2012, 10: 38

484 485

3.

Rohrlack, T.; Hyenstrand, P. Fate of intracellular microcystins in the

486

cyanobacterium Microcystis aeruginosa (Chroococcales, Cyanophyceae). Phycologia

487

2007, 46: 277-283.

488 489

4.

Pearson, L.A.; Hisbergues, M.; Borner, T.; Dittmann, E.,; Neilan, B.A.

490

Inactivation of an ABC transporter gene, mcyH, results in loss of microcystin

491

production in the cyanobacterium Microcystis aeruginosa PCC 7806. Appl. Environ.

492

Microbiol. 2004, 70: 6370-6378.

493 494

5.

Kurmayer, R.; Christiansen, G.; Chorus, I. The abundance of microcystin-

495

producing genotypes correlates positively with colony size in Microcystis sp. and

496

determines its microcystin net production in Lake Wannsee. Appl. Environ. Microbiol.

497

2003, 69: 787-795.

498

ACS Paragon Plus Environment

16

Environmental Science & Technology

499

6.

500

changes in selected representative phytoplanktons. Microb. Ecol. 2005, 50: 298-305.

Page 18 of 32

Sedmak, B.; Elersek, T. Microcystins induce morphological and physiological

501 502

7.

Gan, N.; Xiao, Y.; Zhu, L.; Wu, Z.; Liu, J.; Hu, C.; Song, L. The role of

503

microcystins in maintaining colonies of bloom-forming Microcystis spp. Environ.

504

Microbiol. 2012, 14: 730-742.

505 506

8.

Kehr, J.C.; Zilliges, Y.; Springer, A.; Disney, M.D.; Ratner, D.D.; Bouchier, C.

507

et al. A mannan binding lectin is involved in cell-cell attachment in a toxic strain of

508

Microcystis aeruginosa. Mol. Microbiol. 2006, 59: 893-906.

509 510

9.

Zilliges, Y.; Kehr, J.C.; Mikkat, S.; Bouchier, C.; de Marsac, N.T.; Borner, T.;

511

Dittmann, E. An extracellular glycoprotein is implicated in cell-cell contacts in the toxic

512

cyanobacterium Microcystis aeruginosa PCC 7806. J. Bacterio.l 2008, 190: 2871-2879.

513 514

10.

Babica, P.; Blaha, L.; Marsalek, B. Exploring the natural role of microcystins; a

515

review of effects of photoautotrophic organism. J. Phycol. 2006, 42: 9-20.

516 517

11.

Rantala, A.; Fewer, D.P.; Hisbergues, M.; Rouhiainen, L.; Vaitomaa, J.; Borner,

518

T.; Sivonen, K. Phylogenetic evidence for the early evolution of microcystin synthesis.

519

Proc. Natl. Acad. Sci. U S A 2004, 101: 568-573.

520 521

12.

Schatz, D.; Keren, Y.; Vardi, A.; Sukenik, A.; Carmeli, S.; Borner, T. et al.

522

Towards clarification of the biological role of microcystins, a family of cyanobacterial

523

toxins. Environ. Microbiol. 2007, 9: 965-970.

524 525

13.

Dittmann, E.; Erhard, M.; Kaebernick, M.; Scheler, C.; Neilan, B.A., von

526

Dohren, H.; Borner, T. Altered expression of two light-dependent genes in a

527

microcystin-lacking mutant of Microcystis aeruginosa PCC 7806. Microbiology 2001,

528

147: 3113-3119.

529 530

14.

Zilliges, Y.; Kehr, J.C.; Meissner, S.; Ishida, K.; Mikkat, S.; Hagemann, M. et

531

al. The cyanobacterial hepatotoxin microcystin binds to proteins and increases the

532

fitness of microcystis under oxidative stress conditions. PLoS One 2011, 6: e17615.

ACS Paragon Plus Environment

17

Page 19 of 32

Environmental Science & Technology

533 534

15.

Jahnichen, S.; Ihle, T.; Petzoldt, T.; Benndorf, J. Impact of inorganic carbon

535

availability on microcystin production by Microcystis PCC 7806. Appl. Environ.

536

Microbiol. 2007, 73: 6994-7002

537 538

16.

Phelan, R.R.; Downing, T.G. A Growth Advantage for Microcystin Production

539

by Microcystis PCC7806 under High Light. J. Phycol. 2011, 47: 1241-1246.

540 541

17.

Kardinaal, W.E.; Tonk, L.; Janse, I.; Hol, S.; Slot, P.; Huisman, J.; Visser, P.M.

542

Competition for light between toxic and nontoxic strains of the harmful cyanobacterium

543

Microcystis. Appl. Environ. Microbiol. 2007, 73: 2939-2946.

544 545

18.

Shi, L.; Carmichael, W.W.; Miller, I. Immonugold Localization of Hepatotoxins

546

in Cyanobacterial Cells. Arch. Microbiol. 1995, 163: 7-15.

547 548

19.

Young, F.M.; Thomson, C.; Metcalf, J.S.; Lucocq, J.M.; Codd, G.A.

549

Immunogold localisation of microcystins in cryosectioned cells of Microcystis. J.

550

Struct. Biol. 2005, 151: 208-214.

551 552

20.

Gerbersdorf, S.U. An advanced technique for immuno-labelling of microcystins

553

in cryosectioned cells of Microcystis aeruginosa PCC 7806 (cyanobacteria):

554

implementations of an experiment with varying light scenarios and culture densities.

555

Toxicon 2006, 47: 218-228.

556 557

21.

Downing, T.G.; Phelan, R.R.; Downing, S. A potential physiological role for

558

cyanotoxins in cyanobacteria of arid environments. J. Arid. Environ. 2015, 112: 147-

559

151.

560 561

22.

Vela, L.; Sevilla, E.; Gonzalez, C.; Bes, M.T.; Fillat, M.F.; Peleato, M.L.

562

Exploring the interaction of microcystin-LR with proteins and DNA. Toxicol. In Vitro

563

2008, 22: 1714-1718.

564

ACS Paragon Plus Environment

18

Environmental Science & Technology

Page 20 of 32

565

23.

Meissner, S.; Fastner, J.; Dittmann, E. Microcystin production revisited:

566

conjugate formation makes a major contribution. Environ. Microbiol. 2013, 15: 1810-

567

1820.

568 569

24.

Alexova, R.; Dang, T.C.; Fujii, M.; Raftery, M.J.; Waite, T.D.; Ferrari, B.C.;

570

Neilan, B.A. Specific global responses to N and Fe nutrition in toxic and non-toxic

571

Microcystis aeruginosa. Environ. Microbiol. 2015, 18: 401-413.

572 573

25.

Zhu, Z.; Zhang, L.; Shi, G.Q. Proteasome as a Molecular Target of Microcystin-

574

LR. Toxins 2015, 7: 2221-2231.

575 576

26.

Lyck, S.; Gjolme, N.; Utkilen, H. Iron starvation increases toxicity of

577

Microcystis aeruginosa CYA 228/1 (Chroococcales, Cyanophyceae). Phycologia 1996,

578

35: 120-124.

579 580

27.

Utkilen, H.; Gjolme, N. Iron-stimulated toxin production in Microcystis

581

aeruginosa. Appl. Environ. Microbiol. 1995, 61: 797-800.

582 583

28.

Crosa, J.H.; Walsh, C.T. Genetics and assembly line enzymology of siderophore

584

biosynthesis in bacteria. Microbiol. Mol. Biol. Rev. 2002, 66: 223-249

585 586

29.

Martin-Luna, B.; Sevilla, E.; Hernandez, J.A.; Bes, M.T.; Fillat, M.F.; Peleato,

587

M.L. Fur from Microcystis aeruginosa binds in vitro promoter regions of the

588

microcystin biosynthesis gene cluster. Phytochemistry 2006, 67: 876-881.

589 590

30.

Martin-Luna, B.; Hernandez, J.A.; Bes, M.T.; Fillat, M.F.; Peleato, M.L.

591

Identification of a Ferric uptake regulator from Microcystis aeruginosa PCC7806.

592

FEMS Microbiol. Lett. 2006, 254: 63-70.

593 594

31.

Sevilla, E.; Martin-Luna, B.; Vela, L.; Bes, M.T.; Fillat, M.F.; Peleato, M.L.

595

Iron availability affects mcyD expression and microcystin-LR synthesis in Microcystis

596

aeruginosa PCC7806. Environ. Microbiol. 2008, 10: 2476-2483.

597

ACS Paragon Plus Environment

19

Page 21 of 32

Environmental Science & Technology

598

32.

Humble, A.V.; Gadd, G.M.; Codd, G.A. Binding of copper and zinc to three

599

cyanobacterial microcystins quantified by differential pulse polarography. Water. Res.

600

1997, 31: 1679-1686.

601 602 603

33.

Saito, K.; Sei, Y.; Miki, S.; Yamaguchi, K. Detection of microcystin-metal

604

complexes by using cryospray ionization-Fourier transform ion cyclotron resonance

605

mass spectrometry. Toxicon 2008, 51: 1496-1498.

606 607

34.

Klein, A.R.; Baldwin, D.S.; Silvester, E. Proton and iron binding by the

608

cyanobacterial toxin microcystin-LR. Environ. Sci. Technol. 2013, 47: 5178-5184.

609 610

35.

Rippka, R. Isolation and purification of cyanobacteria. Methods Enzymol. 1988,

611

167: 3-27.

612 613

36.

Janse, I.; Kardinaal, W. E. A.; Meima, M.; Fastner, J.; Visser, P. M.; Zwart, G.

614

Toxic and Nontoxic Microcystis Colonies in Natural Populations Can Be Differentiated

615

on the Basis of rRNA Gene Internal Transcribed Spacer Diversity. Appl. Environ.

616

Microb. 2004, 70: 3979–3987.

617 618

37.

Mackinney, G. Absorption of light by chlorophyll solutions. J. Biol. Chem.

619

1941, 140: 315-322.

620 621

38.

Lin, L.N.; Mason, A.B.; Woodworth, R.C. ; Brandts, J. F. Calorimetric studies

622

of the binding of ferric ions to human serum transferrin. Biochemistry, 1993, 9398-

623

9406.

624 625

39.

Liu, C.X.; Gorby, Y.A.; Zachara, J.M.; Fredrickson, J.K.; Brown, C.F.

626

Reduction kinetics of Fe(III), Co(III), U(VI) Cr(VI) and Tc(VII) in cultures of

627

dissimilatory metal-reducing bacteria. Biotechnol. Bioeng. 2002, 80: 637-649.

628 629

40.

Marcuello, C.; Arilla-Luna, S.; Medina, M.; Lostao, A. Detection of a

630

quaternary organization into dimer of trimers of C. ammoniagenes FAD synthetase at

ACS Paragon Plus Environment

20

Environmental Science & Technology

Page 22 of 32

631

the single-molecule level and at the in cell level. Biochim. Biophys. Acta 2013, 1834,

632

665-676

633 634

41.

Sotres, J.; Lostao, A.; Gómez-Moreno, C.; Baró, A.M. Jumping mode AFM

635

Imaging of biomolecules in the repulsive electrical double layer. Ultramicroscopy 2007,

636

107, 1207-1212.

637 638

42.

Horcas, I.; Fernandez, R.; Gomez-Rodriguez, J.M.; Colchero, J.; Gomez-

639

Herrero, J.; Baro, A.M. WSXM: a software for scanning probe microscopy and a tool

640

for nanotechnology. Rev. Sci. Instrum. 2007, 78 013705.

641 642

43.

Lostao, A.; Peleato M.L.; Gómez-Moreno, C.; Fillat, M.F. Oligomerization

643

properties of FurA from the cyanobacterium Anabaena sp. PCC 7120: direct

644

visualization by in situ atomic force microscopy under different redox conditions.

645

Biochim. Biophys. Acta 2010, 1804: 1723-9

646 647

44.

Yan, F.; Ozsoz, M.; Sadik, O.A. Electrochemical and conformational studies of

648

microcystin-LR. Anal. Chim. Acta 2000, 409: 247-255.

649 650

45.

Meinander, K.; Jensen, T.N.; Simonsen, S.B.; Helveg, S.; Lauritsen, J.V.

651

Quantification of tip-broadening in non-contact atomic force microscopy with carbon

652

nanotube tips. Nanotechnology 2012, 23: 405705

653 654

46.

Strauss, N. Iron deprivation: Physiology and gene regulation. In "The

655

Molecular Biology of Cyanobacteria" Ed D. A. Bryan. Kluwer. pp 731-750. 2006

656 657

47.

Yeung, A.C.; D´Agostino, P.M.; Poljak, A.; McDonald, J.; Bligh, M.W.; Waite,

658

T.D.; Neilan, B.A. Physiological and Proteomic Responses of Continuous Cultures of

659

Microcystis aeruginosa PCC 7806 to Changes in Iron Bioavailability and Growth Rate.

660

App. Environ. Microbiol. 2016, 82: 5918-5929.

661 662

48.

Fujii, M.; Rose, A.L.; Waite, T.D. Iron uptake by toxic and nontoxic strains of

663

Microcystis aeruginosa. Appl. Environ. Microbiol. 2011, 77: 7068-7071.

664

ACS Paragon Plus Environment

21

Page 23 of 32

Environmental Science & Technology

665

49.

Sedmak, B.; Kosi, G., The role of microcystins in heavy cyanobacterial bloom

666

formation (vol 20, pg 691, 1998). J. Plankton Res. 1998, 20, (7), 1421-1421.

667 668

50.

Zheng, T.; Nolan, E.M. Siderophore-based detection of Fe(III) and microbial

669

pathogens. Metallomics 2012, 4: 886-880.

670 671

51.

Que, E.L.; Domaille, D.W.; Chang, C.J. Metals in Neurobiology: probing their

672

chemistry and biology with molecular imaging. Chem. Rev. 2008, 108: 1517-1549.

673 674

52.

Miethke, M.; Marahiel, M.A. Siderophore-based iron acquisition and pathogen

675

control. Microbiol. Mol. Biol. Rev. 2007, 71:413-51.

676 677

53.

Sriswasdi, S.; Harper, S.L.; Tang, H.Y.: Speicher, D.W. Enhanced

678

Identification of Zero-Length Chemical Cross-Links Using Label-Free Quantitation and

679

High-Resolution Fragment Ion Spectra. J. Proteome Res. 2014, 13: 898-914.

680 681

54.

Trogen, G.B.; Eriksson, A.A.; Kontteli, M.; Meriluoto, J.; Sethson, I.; Zdunek,

682

J.; Edlund, U. Conformational studies of microcystin-LR using NMR spectroscopy and

683

molecular dynamics calculations. 1996. Biochemistry 35:3197-3205.

684 685

55.

Amano, Y.; Hosoi, T.; Machida, M.; Imazeki, F. Effects of extracellular

686

polymeric substances (EPS) an iron ion on colony formation of unicellular Microcystis

687

aeruginosa. J. Japan Soc. Civil. Eng., Ser.G (Environ. Res.) 2013, 69: III_39-III_44.

688 689

56

Downing, T. G.; Phelan, R. R.; Downing, S., A potential physiological role for

690

cyanotoxins in cyanobacteria of arid environments. J. Arid Environ. 2015, 112, 147-

691

151.

692 693

57.

Wang, Z. C.; Li, D. H.; Cao, X. Y.; Song, C. L.; Zhou, Y. Y., Photosynthesis

694

regulates succession of toxic and nontoxic strains in blooms of Microcystis

695

(Cyanobacteria). Phycologia 2015, 54, (6), 640-648.

696 697

58.

Phelan, R. R.; Downing, T. G., The localization of exogenous microcystin LR

698

taken up by a non-microcystin producing cyanobacterium. Toxicon 2014, 89, 87-90.

ACS Paragon Plus Environment

22

Environmental Science & Technology

Page 24 of 32

699 700

59.

Juttner, F.; Luthi, H., Topology and enhanced toxicity of bound microcystins in

701

Microcystis PCC 7806. Toxicon 2008, 51, (3), 388-397.

702 703

ACS Paragon Plus Environment

23

Page 25 of 32

Environmental Science & Technology

704 705

Figure legends

706 707

Figure 1. Microcystis aeruginosa strains PCC 7806 (microcystin-producer) and PCC

708

7005 (microcystin non-producer) growth. Growth was determined in iron-replete and

709

iron-deficient culture medium. Panel A: number of cells/mL. Panel B: total soluble

710

protein. Panel C: chlorophyll a content.

711 712

Figure 2: Microcystin-metal interaction monitored by ITC. Microcystin-LR solution (10

713

µM) was titrated with metal solution (100-150 µM) at 25°C: (A) Fe2+ in 100 mM

714

phosphate buffer, 1.5 mM ascorbate, pH 7; (B) Fe3+ in 100 mM phosphate buffer, 1.5

715

mM ascorbate, pH 7; (C) Mo6+ in 100 mM Tris-HCl, pH 8.5. Inset: Control experiment

716

injecting Fe2+ into 100 mM phosphate, pH 7. Upper panels show the thermograms

717

(thermal power as a function of time) and the lower panels show the binding isotherm

718

(normalized heat as a function of the molar ratio).

719 720

Figure 3: AFM topography images of microcystin-LR samples diluted in 20 %

721

methanol (A) and presence of EDC (D). A majority of monomers can be observed in

722

(A), whereas the ratio of dimers increases in (D). (B) 2D image of a monomeric feature

723

from (A) and height profile (corresponding to the blue line) (C). (E) 2D image of a

724

representative dimer from (D) and height profile for the two protomers of the dimer (F).

725

Heights in the Z axis for each protomer are numbered. Monomers and dimers are

726

highlighted in white and red circles, respectively.

727 728

Figure 4: AFM topography images of crosslinked microcystin-LR samples upon

729

incubation with different metals in 20 % methanol: image showing different species

730

(A), a zoomed image showing a trimer in detail (B) and its corresponding height profile

731

highlighting the different units (C), in the presence of Fe2+; image showing different

732

species (D), a zoomed image showing a tetramer in detail (E) and its corresponding

733

height profile highlighting the different units (F), in the presence of Cu2+; image

734

showing different species (G), a zoomed image showing a hexameric chain in detail (H)

ACS Paragon Plus Environment

24

Environmental Science & Technology

Page 26 of 32

735

and its corresponding height profile highlighting the different units (I), in the presence

736

of Fe3+. Monomers are rounded in white, dimers in red, trimers in green, tetramers in

737

yellow and hexamers in pink.

738 739

ACS Paragon Plus Environment

25

Page 27 of 32

Environmental Science & Technology

740

Table 1. Binding parameters for microcystin-LR interacting with different metals at

741

25°C. The buffers are 150 mM sodium phosphate, 100 mM HEPES or 50 mM Tris-

742

HCl. Buffer

pH

Kd (µM)

∆H (kcal/mol)

Fe2+

Phosphate

7.0

2.4

-5.6

Fe3+

Phosphate

7.0

2.4

-2.8

Mo6+

Phosphate

7.0

37

-8.1

Hepes

7.6

15

-3.2

Tris-HCl

8.0

3.3

-14.3

Tris-HCl

8.5

1.3

-21.8

Hepes

7.6

1.2

-2.3

Tris-HCl

8.0

4.6

-8.5

Tris-HCl

8.5

10

-7.0

Hepes

7.6

1.3

-1.2

Tris-HCl

8.0

31

-8.0

Tris-HCl

8.5

44

-11.9

Cu2+

Mn2+

743

Relative error in Kd is 15%.

744

Absolute error in ∆H is 0.5 kcal/mol.

745

ACS Paragon Plus Environment

26

Environmental Science & Technology

Page 28 of 32

746 747

Table 2. Percentages of microcystin-LR oligomeric species detected by AFM. All

748

measurements were performed in 20 % methanol except indicated. The samples were

749

cross-linked (EDC) in the presence of indicated metals. Features correspond to image

750

units, while molecules refer to the amount of individual peptide monomers in the

751

corresponding image units. Error associated to percentage determination was ± 5–15%.

752 CHAIN ASSOCIATION Conditions

Units

Microcystin- Features

Monomers Dimers Trimers Tetramers Hexamers (%)

(%)

(%)

%

%

95

5

----

----

----

LR

Molecules 90

10

----

----

----

+Milli-Q

Features

5

---

----

----

water

Molecules 90

10

---

----

----

+PBS

Features

98

2

----

----

----

pH 7.4

Molecules 96

4

----

----

----

Features

72

27

1

----

----

Molecules 56

42

2

----

----

+EDC

Features

97

3

----

----

----

Mn (II)

Molecules 95

5

----

----

----

+EDC

Features

97

3

----

----

----

Zn (II)

Molecules 95

5

----

----

----

+EDC

Features

65

25

10

----

----

Fe (II)

Molecules 45

34

21

----

----

+EDC

Features

67

20

3

10

----

Cu (II)

Molecules 43

25

6

26

----

+EDC

Features

42

30

----

----

28

Fe (III)

Molecules 16

22

----

----

62

+EDC

95

753 754

ACS Paragon Plus Environment

27

Page 29 of 32

Environmental Science & Technology

Figure 1 M.  aeruginosa  PCC  7806  Control   M.  aeruginosa  PCC  7806  -­‐Fe   M.  aeruginosa  PCC  7005  Control   M.  aeruginosa  PCC  7005  -­‐Fe  

Number of cells (10-6/mL)

50   40  

A

30   20   10   0   0  

2  

4  

6  

8  

9  

12  

Total protein/cell (µg/cell)

Time (days)

60   50  

M.  aeruginosa  PCC  7806  Control   M.  aeruginosa  PCC  7806  -­‐Fe   M.  aeruginosa  PCC  7005  Control   M.  aeruginosa  PCC  7005  -­‐Fe  

B

40   30   20   10   0   0  

2  

4  

6  

8  

12  

Chlorophyll a (µg/ml)

Time (days)

6  

M.  aeruginosa  PCC  7806  Control   M.  aeruginosa  PCC  7806  -­‐Fe   M.  aeruginosa  PCC  7005  Control   M.  aeruginosa  PCC  7005  -­‐Fe  

C

4   2   0   0  

2  

4  

6  

8  

9  

Time (days)

ACS Paragon Plus Environment

12  

Environmental Science & Technology

Page 30 of 32

Figure 2!

[Fe(II)]T/[microcystin-LR]T!

dQ/dt (µcal/s)!

B

Time (min)!

C

Q (kcal/mol of injectant)!

dQ/dt (µcal/s)!

Time (min)!

Q (kcal/mol of injectant)!

A

Q (kcal/mol of injectant)!

dQ/dt (µcal/s)!

Time (min)!

[Fe(III)]T/[microcystin-LR]T!

ACS Paragon Plus Environment

[Mo(VI)]T/[microcystin-LR]T!

Page 31 of 32

Environmental Science & Technology

Figure 3

A

B

C

D

E

F

ACS Paragon Plus Environment

Environmental Science & Technology

Page 32 of 32

Figure 4

A

B

C

D

E

F

G

H

I

E ACS Paragon Plus Environment