Implications for Phosphorus Recovery - ACS Publications - American

Nov 13, 2016 - Dissolution of Natural Brucite: Implications for Phosphorus Recovery ... interactions of ammonium phosphate solutions with brucite...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of Otago Library

Article

In situ nanoscale imaging of struvite formation during the dissolution of natural brucite: implications for phosphorus recovery from wastewaters Jörn Hövelmann, and Christine V Putnis Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b04623 • Publication Date (Web): 13 Nov 2016 Downloaded from http://pubs.acs.org on November 14, 2016

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 30

Environmental Science & Technology

1

In situ nanoscale imaging of struvite formation during the dissolution of

2

natural brucite: implications for phosphorus recovery from wastewaters

3

Jörn Hövelmann1* and ChristineV. Putnis2,3

4 5 6

1

German Research Centre for Geosciences (GFZ), Interface Geochemistry, 14473 Potsdam, Germany 2

7 8

3

Institut für Mineralogie, University of Münster, 48149 Münster, Germany

The Institute for Geoscience Research (TIGeR), Department of Chemistry, Curtin University, Perth 6845, Australia

9 10 11

*Corresponding author:

12

Phone: +49 331 288-28703; E-mail: [email protected] (Jörn Hövelmann)

1 ACS Paragon Plus Environment

Environmental Science & Technology

13

Abstract

14

As phosphorus (P) resources are diminishing, the recovery of this essential nutrient from wastewaters

15

becomes an increasingly interesting option. P-recovery through the controlled crystallization of

16

struvite (MgNH4PO4•6H2O), a potential slow-release fertilizer, is highly attractive, but costly if large

17

amounts of Mg have to be added. In this context, natural Mg-minerals like brucite (Mg(OH)2) could

18

provide more cost effective Mg-sources compared to high-grade Mg-compounds such as MgCl2. Here

19

we used in situ atomic force microscopy (AFM) to study the interactions of ammonium phosphate

20

solutions with brucite (001) cleavage surfaces. Brucite dissolution was strongly enhanced in the

21

presence of H2PO4- ions, most likely due to the formation of negatively charged surface complexes.

22

Simultaneously with brucite dissolution, we directly observed the formation of a new phase that was

23

identified as struvite by Raman spectroscopy. Our results suggest that brucite dissolution and struvite

24

precipitation were coupled at the mineral-fluid interface within a thin fluid boundary layer. An

25

interpretation is proposed where the heterogeneous nucleation and growth of struvite occurs via a

26

particle-mediated process involving the formation of primary nanoparticles, followed by their

27

continuous aggregation, fusion and possible transformation to crystalline struvite. These observations

28

have implications for the feasibility of using brucite in phosphorus recovery processes.

2 ACS Paragon Plus Environment

Page 2 of 30

Page 3 of 30

Environmental Science & Technology

29

Introduction:

30

Phosphorus (P) is a vital nutrient for plant growth, but its availability to plants is often limited due to

31

its adsorption on soil mineral surfaces (mainly Fe- and Al-(hydr)oxides1) or its precipitation in the

32

form of sparingly soluble phosphates (e.g., Ca-phosphate2). Thus, the use of phosphate fertilizers is

33

essential in modern agriculture to ensure adequate food production for a growing global population.

34

However, global resources of phosphate ores for fertilizer production are finite and may run out by the

35

end of this century3. There is a substantial ambiguity regarding the actual extent of global phosphate

36

reserves. However, although the timing of a ʻpeak phosphorusʼ remains uncertain, there is no dispute

37

about its limited resource availability considering the growing global demand for fertilizers4. At the

38

same time, high levels of fertilizers applied to land increase the potential that excess amounts of P will

39

be washed out into the groundwater causing problems such as eutrophication of streams, rivers and

40

coastal regions, presenting a major environmental concern. Hence, a more sustainable P management

41

becomes increasingly important. In recent years, a lot of research has focused on investigating

42

effective routes for the recovery of P from various industrial, farm and municipal wastewater streams.

43

In this context, the mineral struvite (MgNH4PO4•6H2O) has gained strong interest5,6,7. Its controlled

44

precipitation from wastewaters does not only enable the safe disposal of nutrient-laden wastes, but

45

could also contribute to the conservation of natural P resources, as recovered struvite crystals may be

46

reused directly as an eco-friendly slow-release fertilizer8,9. However, despite these highly attractive

47

prospects, such struvite-based P recovery methods are not yet widely adopted because high costs for

48

wastewater pre-treatments often limit their economic efficiency. To ensure effective struvite

49

crystallization, the solution must be slightly alkaline (pH ~8-9) and contain phosphate (PO43-),

50

ammonium (NH4+) and magnesium (Mg2+) ions in close to equimolar (1:1:1) concentrations10. Thus,

51

for most wastewater sources pH has to be increased, which is typically done by adding an alkali source

52

such as sodium hydroxide (NaOH). Moreover, the addition of magnesium is generally required since

53

most nutrient-rich wastewaters are deficient in Mg2+ relative to PO43- and NH4+. Most of the

54

commercially available struvite recovery technologies use pure Mg salts such as MgCl2, MgSO4 and

55

MgO6. However, these high-grade compounds are expensive and may contribute up to 75% of the

3 ACS Paragon Plus Environment

Environmental Science & Technology

56

overall costs, making large-scale applications uneconomical11. Therefore, employing cheaper Mg

57

sources would be an effective way for cost reduction12.

58

Comparatively inexpensive Mg-sources include seawater, bittern (a by-product of salt manufacture),

59

low-grade caustic magnesia (MgO) as well as natural Mg-rich minerals12-22. The use of seawater and

60

bittern, although principally feasible17,18, appears only economical when close to the sea since

61

comparatively large volumes are required. On the other hand, the efficiency of solid Mg sources

62

depends to a large degree on their availability, Mg content, solubility and reactivity19. The common

63

mineral magnesite (MgCO3), for example, has a high Mg content (~30 wt%), but a low solubility (Ksp

64

= 10-7.8 at 25°C) and extremely slow dissolution rate in water at ambient temperature23. Therefore,

65

either high doses of magnesite are required or it has to be pre-treated by acid dissolution or thermal

66

decomposition to make enough Mg available for struvite formation14,21. Brucite (Mg(OH)2) is another

67

natural Mg-rich mineral. It is less abundant than magnesite, but has a higher Mg content (~40 wt%).

68

Brucite is widely distributed as an accessory mineral in a variety of rock types. Significant quantities

69

(up to 20 wt%) are, for example, found in ultramafic rocks such as peridotites or serpentinites24,25.

70

These common rock types are extensively mined for their chromite and nickel ores, hence making

71

brucite also an abundant component of ultramafic mining residues26. Most brucite deposits of

72

economic interest are, however, hosted by metamorphosed carbonate rocks, e.g. dolomitic marbles27.

73

Using brucite as a reactant for struvite recovery could have several advantages. First of all, brucite

74

dissolves relatively fast in the near-neutral pH region28,29,30, hence it could readily be added in the solid

75

form without prior dissolution in acid. Furthermore, brucite dissolution not only releases Mg2+ (Mg

76

source), but also OH- (alkali source), which helps to neutralize acids so as to achieve the optimal pH

77

values for struvite precipitation. At the same time, brucite particles could act as nucleation sites for

78

struvite, hence reducing induction times and increasing crystallization rates20. On the other hand,

79

extensive precipitation of struvite onto the dissolving brucite substrate may also lead to surface

80

passivation that could ultimately reduce the efficiency of the coupled dissolution-precipitation

81

process19.

82

Huang and co-workers15 have recently evaluated the use of natural brucite for struvite precipitation to

83

remove ammonium from wastewater generated in the separation process for rare-earth elements. The 4 ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30

Environmental Science & Technology

84

authors performed batch-type experiments using different brucite-to-wastewater ratios and

85

demonstrated that 93-95% of ammonium could be removed as struvite within 12 h. They also report

86

that the precipitates collected at the end of the experiments contained a large amount of non-reacted

87

brucite that could be reused in subsequent experiments for further struvite precipitation.

88

While the use of brucite as a reactive medium for struvite precipitation seems principally feasible, a

89

detailed understanding of how exactly brucite interacts with ammonium and phosphate bearing

90

solutions is currently lacking. In particular, there is no direct molecular-scale observations of the

91

mechanistic and kinetic pathways of the coupled process of brucite dissolution and subsequent struvite

92

precipitation. However, such information is crucial to fully assess the potential of using brucite as an

93

alternative Mg source and may provide hints of how the operational parameters of the struvite

94

recovery process could be optimized.

95

In this study we investigate the interactions of brucite surfaces with ammonium phosphate solutions by

96

in situ, real-time imaging at the nanometer scale using atomic force microscopy (AFM) in connection

97

with a fluid reaction cell. The objective was to elucidate and quantify the effect of phosphate and

98

ammonium on the dissolution kinetics of brucite and to characterize the coupling between brucite

99

dissolution and the following nucleation and growth of struvite at different initial pH values (6 – 9.5),

100

ammonium phosphate concentrations (5 – 100 mM) and N:P molar ratios (1:1 and 2:1). Our results

101

have implications for the efficiency of using natural brucite as well as synthetic Mg(OH)2 in struvite

102

recovery processes.

103 104

Methodology:

105

Brucite Specimen. Natural brucite from the Tallgruvan (Norberg, Sweden) was used for the

106

experiments. The initial, essentially monomineralic brucite rock sample contained minor amounts of

107

dolomite (CaMg(CO3)2), magnetite (Fe3O4) and pyroaurite (Mg6Fe2(CO3)(OH)16•4H2O), which were

108

avoided during AFM specimen preparation. Only optically transparent brucite crystals were used.

109

Immediately before each experiment a brucite crystal was cleaved parallel to the (001) cleavage plane

110

to expose a fresh surface. The final dimensions of the brucite specimens used in the AFM experiments

111

were ca. 3 x 3 x 0.2 mm. 5 ACS Paragon Plus Environment

Environmental Science & Technology

112

Solutions. Aqueous solutions of phosphate and ammonium (5-100 mM) were prepared by dissolving

113

reagent grade salts (Sigma Aldrich) of ammonium dihydrogen phosphate (NH4H2PO4), diammonium

114

hydrogen phosphate ((NH4)2HPO4) or sodium dihydrogen phosphate (NaH2PO4) into double-

115

deionized water (resistivity > 18mΩ cm-1). Adjustments of pH were made using 0.1 M NaOH or HCl.

116

Ammonium- and phosphate-free solutions with pH values and ionic strengths (adjusted with NaCl)

117

similar to the ammonium phosphate solutions were also used in some experiments.

118

Atomic Force Microscopy. Brucite (001) surfaces were imaged at room temperature (23 ± 1°C) using

119

a Bruker Multimode Atomic Force Microscope (AFM) operating in contact mode. In situ experiments

120

were performed within an O-ring sealed flow-through fluid cell from Digital Instruments (Bruker).

121

Solutions were injected at regular time intervals between each scan (lasting ~1.5 min), giving an

122

effective flow rate of 22 µL s-1. AFM images were collected using Si3N4 tips (Bruker, tip model NP-

123

S20) with spring constants of 0.12 and 0.58 N m-1. Images were analyzed using the NanoScope

124

Analysis software (version 1.50). Step retreat velocities or etch pit spreading rates (vs) were calculated

125

measuring the length increase of etch pit step edges (s) per unit time in sequential images scanned in

126

the same direction. For each experimental condition at least 5 different etch pits were analyzed in 2-10

127

pairs of sequential images. Each etch pit spreading rate value thus represents an average of 10-50

128

individual measurements.

129

For several experiments, the outlet fluid was sampled, collecting a sequence of 2-4 aliquots of 8 ml

130

(the outflow from four consecutive scans) that were later analyzed for magnesium concentration using

131

ICP-OES (inductively coupled plasma – optical emission spectroscopy). The analytical uncertainties

132

of the ICP-OES measurements were below 3% based on repeated measurements of aqueous standards.

133

In some of the experiments scanning was stopped from time to time and the solution in the fluid cell

134

was kept static for several minutes, hence, allowing the system to approach equilibrium.

135

Following in situ AFM experiments, some samples were placed directly into a beaker filled with 10 ml

136

of the solution used in that specific in situ experiment. After 12 – 36 hours of reaction the samples

137

were recovered from the solution and immediately dried by absorbing the fluid with filter paper. The

138

crystals from these ex situ experiments were then re-examined in air in the AFM.

6 ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30

Environmental Science & Technology

139

Scanning Electron Microscopy. Samples from the ex situ experiments were also imaged using an

140

Ultra 55 Plus (Carl Zeiss SMT) scanning electron microscope (SEM) equipped with an energy

141

dispersive X-ray (EDX) detector for elemental analyses of the reacted brucite surfaces and newly

142

formed precipitates. Before imaging, all samples were coated with a 20-nm-thick layer of carbon.

143

Raman Spectroscopy. A confocal Raman spectrometer (Horiba Jobin Yvon XploRA) operating with

144

the 638 nm line of a He–Ne laser was used for the analysis of surface precipitates on brucite after

145

contact with ammonium phosphate bearing solutions. Spectra were taken with a 500 µm hole, 100 µm

146

slit and 1200 grooves per millimeter grating using an acquisition time of 2 x 60 s. Corrections for

147

system drift were made using the 520.7 cm-1 Raman band of a silicon standard taken at the beginning

148

and the end of the Raman session. Reference spectra of struvite were obtained from the RRUFF

149

database (http://www.rruff.info).

150

Geochemical Modelling. The hydro-geochemical software PHREEQC31 (version 3.2.0-9820) was

151

used to calculate the chemical speciation of the initial solutions used in the AFM experiments as well

152

as to simulate the reactions with brucite. All calculations were done using the sit.dat database (version

153

9a, July 2014, www.thermochimie-tdb.com) that was supplemented with thermodynamic data for

154

struvite taken from Bhuiyan et al.32.

155 156

Results and Discussion:

157

Dissolution Features. In situ AFM observations show that dissolution of brucite (001) cleavage

158

surfaces is very slow when in contact with pure water or moderately saline NaCl solutions. In 100 mM

159

NaCl (pH 7) dissolution only occurred by the retreat of pre-existing step edges. However, no etch pits

160

were observed to form, even after a contact time of 22 min (Fig. 1A). Dissolution then rapidly

161

increased upon injection of a 10 mM (NH4)2HPO4 solution (pH 7.9) with the immediate formation and

162

spreading of equilateral triangular etch pits (Fig. 1B,C). Large variations in etch pit density (ranging

163

from 100 in a scanned area of 5x5 µm) were observed between different surfaces, but also

164

between different areas of the same surface, as well as at different reaction times, possibly indicating a

165

substantial heterogeneity in the distribution of crystal defects. Most etch pits were initially shallow

166

(single stepped) and randomly distributed on the surface. Lateral spreading of etch pits eventually 7 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 30

167

resulted in their coalescence leaving behind small islands that disappeared upon further dissolution

168

leading to a layer-by-layer dissolution mechanism. With time, some etch pits developed into deeper,

169

concentric (multi stepped) etch pits (Fig. 1D). These etch pits most likely originate from structural

170

defects that penetrate several layers, whereas the monolayer etch pits nucleated at either defect-free

171

surfaces or point defects33. Occasionally, also spiral etch pits were observed, which are likely to be

172

related to screw dislocations. The height profile in Fig. 1E shows that individual steps are ~0.5 nm

173

deep.

174

The observed etch pit and step edge morphologies are in agreement with previous studies28,29 and can

175

be understood by considering the crystal structure of brucite (Fig. 1F). Brucite is composed of sheets

176

of edge-sharing Mg(OH)6 octahedra. Each hydroxyl (OH) group is coordinated by three Mg atoms

177

with the O-H-vector perpendicular to the (001) plane. The H-atom forms hydrogen bonds to three

178

oxygen atoms on the adjacent brucite sheet. The thickness of a single brucite sheet is 0.47 nm, which

179

corresponds to the depth of individual steps observed in AFM. The equilateral triangular shape of the

180

etch pits results from the 3-fold rotation axis normal to the (001) plane. The edges of the triangles are

181

parallel to the symmetrically equivalent [100], [010] and [110] directions, which are characterized by

182

strong periodic bond chains (PBCs) consisting of edge-sharing Mg(OH)6 octahedra.

183

Effects of phosphate and ammonium on brucite dissolution. In general, the presence of ammonium

184

and phosphate strongly increased the dissolution rate of brucite as shown by the much faster spreading

185

of etch pits. As well, increasing the ammonium phosphate concentration resulted in a continuous

186

acceleration of etch pit spreading (Fig. S1).

187

For equilateral triangular etch pits, the spreading rate (vs) can be defined as  =

188

denotes the distance from the center of a triangle to its side and s is the side length of the triangle (Fig.

189

1F). In our experiments, the reproducibility of absolute spreading rate values was somewhat limited as

190

shown by the fact that rate measurements on different surfaces, but under the same conditions (i.e.,

191

same concentrations and pH) gave variable results. For example, spreading rates in 20 mM NH4H2PO4

192

(pH 8.5) measured on three different surfaces were 0.24 (±0.04), 0.33 (±0.07) and 0.49 (±0.04) nm s-1

193

(Fig. 2A). The reason for this surface dependent variability of spreading rates remains elusive, but may

194

be related to inherent variations in surface energy, e.g., due to differences in the number, distribution 8 ACS Paragon Plus Environment

 

=

  , √ 

where x

Page 9 of 30

Environmental Science & Technology

195

and nature of reactive sites34, possibly also implying variable degrees of residual stress/strain in the

196

naturally grown brucite crystals. Nevertheless, an adequate comparison of spreading rates was possible

197

for values measured in the same area of the same surface.

198

On a given surface, measured etch pit spreading rates generally showed an approximately linear

199

increase with increasing ammonium phosphate concentration at a fixed pH, e.g., from 0.07 (±0.02) nm

200

s-1 in 5 mM NH4H2PO4 to 0.51 (±0.06) nm s-1 in 50 mM NH4H2PO4 at pH 8.5 (Fig. 2A). For

201

comparison, measured values in NaCl solutions with the same pH and similar ionic strength only

202

ranged between 0.01 and 0.03 nm s-1, thus any increase in etch pit spreading rates must be due to the

203

presence of phosphate and ammonium. On the other hand, increasing the pH from 8 to 9.5 at a fixed

204

concentration resulted in a nearly exponential decrease in measured etch pit spreading rates, e.g., from

205

0.76 (± 0.11) nm s-1 at pH 8 to 0.04 (± 0.02) nm s-1 at pH 9.5 in 20 mM NH4H2PO4 (Fig. 2B). This

206

trend may be explained by an increased activity of OH- at higher pH values resulting in a lower

207

thermodynamic driving force for brucite dissolution. Another likely explanation for our observations

208

is that, depending on pH, the adsorption of phosphate ions on the brucite substrate either enhances or

209

inhibits its dissolution. It has been shown that the dihydrogen phosphate (H2PO4-) ion accelerates the

210

dissolution of metal (oxy)hydroxides such as brucite35 and goethite36,37 by increasing their surface

211

protonation through the formation of mononuclear negatively charged surface complexes. Conversely,

212

the fully deprotonated phosphate ion (PO43-) may inhibit brucite dissolution by forming binuclear

213

surface complexes that bridge two Mg centers, thereby increasing the energy needed for the

214

detachment of Mg atoms from the brucite crystal lattice35. Thus, it seems reasonable to assume that in

215

our system brucite dissolution kinetics are mostly controlled by the presence of H2PO4- and PO43- ions

216

in solution and their respective adsorption on the brucite surface. This is further corroborated by

217

speciation calculations (Table S1 and Fig. S2) showing that an increase in solution pH from 8 to 9.5

218

decreases the activity of the dissolution enhancing H2PO4- species while increasing the activity of the

219

inhibiting PO43- species, i.e., consistent with the observed decrease in etch pit spreading rates at higher

220

pH values. On the other hand, the activity of the (dominant) singly protonated phosphate (HPO42-) ion

221

remains virtually constant in the considered pH range, suggesting that this species most likely does not

222

affect brucite dissolution. From Figure 2C it can be seen that there is generally a good positive linear 9 ACS Paragon Plus Environment

Environmental Science & Technology

223

correlation (Rsq = 0.945 - 0.981) between the measured spreading rate values of Fig.2A and B and the

224

calculated H2PO4- activities in the corresponding solution.

225

It should be noted that the enhancement of brucite dissolution in the presence of phosphate may also

226

be related, to some degree, to the formation of Mg-phosphate bearing complexes or clusters in solution

227

(Table S2), that lower the saturation state with respect to brucite, i.e., maintaining far-from

228

equilibrium conditions and hence promoting further dissolution. A similar effect will arise from the

229

precipitation of a phosphate-bearing phase such as struvite.

230

We also observed a tendency of higher etch pit spreading rates in the presence of (NH4)2HPO4 relative

231

to those measured in the presence of NH4H2PO4 at the same pH and total phosphate concentration

232

(Fig. 2A). This effect is further confirmed when comparing etch pit spreading rates in NaH2PO4,

233

NH4H2PO4 and (NH4)2HPO4 measured in the same 5x5 µm surface area during a continuous

234

experiment (Fig. 2D). At a total phosphate concentration of 10 or 20 mM and a pH of 8.5, rates in

235

NaH2PO4 (no NH4+ and NH3 present) were consistently ~20% lower compared to rates in NH4H2PO4

236

and ~50% lower compared to rates in (NH4)2HPO4. Moreover, measured Mg concentrations in the

237

outflow solutions showed a slightly increasing trend from NaH2PO4 to (NH4)2HPO4 (Table S3, Fig.

238

S3), thus corroborating the in situ AFM observations. As in the case of phosphate already described,

239

these observations may be explained in several ways. Firstly, an increase in the NH4+ activity will

240

increase the ion activity product of struvite, thus favoring struvite precipitation. The formation of

241

struvite will consume dissolved Mg2+ lowering the ion activity product of brucite and leading to a

242

larger thermodynamic driving force for brucite dissolution. The consumption of dissolved Mg2+ by

243

aqueous Mg-NH3 bearing complexes, on the other hand, likely has a negligible effect on the saturation

244

with respect to brucite as such complexes are predicted to be only present in very low concentration

245

(Table S2). Another possibility is that NH4+ promotes brucite dissolution via the formation of surface

246

complexes or modifications in the water structure and surface hydration dynamics38,39. However, this

247

would have to be tested separately from phosphate, i.e., using ammonium chloride or nitrate solutions.

248 249

10 ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30

Environmental Science & Technology

250

Struvite nucleation and growth onto brucite (001) cleavage surfaces. Simultaneously with brucite

251

dissolution a new phase was observed to form in the presence of all ammonium phosphate solutions.

252

In the earliest nucleation stages, small, isolated particles ( 6 struvite should

304

always be the first phase to become stable, thus being the most likely phase to precipitate. On the other

struvite)

are

newberyite

(Mg(PO3OH)·3H2O),

bobbierite

12 ACS Paragon Plus Environment

(Mg3(PO4)2·8H2O)

and

Page 13 of 30

Environmental Science & Technology

305

hand, for pH values ≤ 6, newberyite may become stable first. Depending on the initial fluid

306

composition and pH, the amount of Mg that needs to be released to reach struvite saturation ranges

307

between 0.06 – 57.5 mg/L (= 0.0025 – 2.4 mmol/L) (Table S4). However, Mg concentrations in the

308

outflow solutions were in most cases below the detection limit ( 8.5) brucite

343

dissolution slows down quickly due to the inhibitory effect of PO43- and/or OH- ions. On the other

344

hand, in the pH range 6 – 8.5 the dissolution promoting effect of H2PO4- ions dominates, thus ensuring

345

high Mg release rates, while the potential of precipitating undesired phases is minimized.

346

While brucite is a relatively cheap Mg source, its solubility in water (Ksp = 10-10.9 at 25°C) is much

347

lower compared to more expensive artificial compounds like MgCl2. This fact may have negative

348

consequences for the efficiency of the struvite formation and harvesting process. However, several

349

aspects also have to be taken into account when assessing the usefulness of brucite as compared to the

350

highly soluble salts. Firstly, brucite dissolution is strongly enhanced in the presence of phosphate.

351

Thus, sufficient amounts of brucite may readily be dissolved without the use of strong acids. In

352

addition, the release of OH- ions during brucite dissolution would lower the amount of NaOH needed

353

to achieve the optimal, slightly alkaline pH conditions for struvite precipitation. Secondly, if brucite

354

dissolution and struvite precipitation are coupled at the reaction interface (as observed in our AFM

355

experiments), absolute solubilities are not necessarily the most relevant factor, but rather the

356

solubilities of the phases in the solution at the mineral interface. The total amount of dissolved Mg

357

may be very low at any time, but the reaction could still be efficient because even the dissolution of

358

just a few monolayers of brucite may highly supersaturate the solution at the reaction interface with

359

respect to struvite43,44. Moreover, as pointed out by Adnan et al.45, high magnesium concentrations in

360

the process fluid may also be of concern because of an increased risk of unintentional struvite 14 ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30

Environmental Science & Technology

361

precipitation elsewhere in the system, especially if the effluent of the reactor is continuously pumped

362

back to the inlet of the treatment plant. This risk could be effectively lowered if Mg is not dosed via

363

the external addition of dissolved MgCl2, but via the dissolution of solid brucite particles inside the

364

reactor. A tight interfacial coupling between brucite dissolution and struvite precipitation, would allow

365

to better control the location of struvite precipitation because supersaturated conditions will only be

366

reached within the small solution volume at the particle-fluid interface, but not within the bulk

367

solution.

368

In our experiments, the elevated Mg concentrations at high dissolution sites most likely played the

369

most important role in controlling the struvite nucleation rates. However, it seems also reasonable to

370

assume that the underlying brucite substrate provided energetically favorable sites for struvite

371

nucleation considering the clear alignment of the formed crystals. Thus, we may infer that the presence

372

of steps on the brucite surfaces triggers heterogeneous nucleation leading to faster precipitation as

373

compared to homogeneous nucleation. This would be consistent with results by Stolzenburg et al.20

374

who observed faster struvite precipitation and higher phosphorous recovery rates when using

375

MgO/Mg(OH)2 suspensions instead of MgCl2 solutions.

376

The formation of a dense surface layer of struvite could eventually armour the brucite from further

377

reaction as has been previously observed in other systems46,47,48. In our case, however, the struvite

378

precipitates did not completely cover the brucite surface. Indeed, the preferential precipitation along

379

deep step edges may eventually block these highly reactive sites and consequently slow down the

380

reaction. However, a complete inhibition is not expected because the flatter areas of the brucite surface

381

remained uncovered during our experiments and continued dissolution (i.e., formation and spreading

382

of new etch pits) in these areas was observed. .

383

Overall, our experimental results suggests that natural brucite is a highly suitable reactive medium for

384

struvite crystallization. It should be noted, however, that our experiments were performed with

385

synthetic ammonium phosphate solutions. The reactions occurring in real wastewaters, which contain

386

many more dissolved species, remain to be studied in more detail to verify whether using brucite is

387

indeed feasible for phosphorus recovery applications. For wastewaters from the rare-earth industry,

388

Huang et al.15 already demonstrated that high struvite recovery rates can indeed be achieved with 15 ACS Paragon Plus Environment

Environmental Science & Technology

389

brucite as a Mg source. Another aspect that needs to be considered in feasibility studies is the

390

availability of brucite. The current worldwide market for brucite is comparatively small and there are

391

only a few large, high-grade brucite deposits in production27. However, ample occurrences of brucite

392

in nature24,25 and a potentially large number of still unexplored and undeveloped deposits27 make the

393

availability of Mg(OH)2 an unlikely limiting factor.

394

Despite the fact that the present study focused on natural brucite, it should be emphasized that our

395

findings are also relevant for the use of synthetic Mg(OH)2 and MgO. Low grade MgO, a by-product

396

of the magnesite industry, has already been tested as a cheap alternative Mg source in a range of

397

laboratory and pilot-scale studies12,19-22. Results by Stolzenburg et al.20 and Castro et al.22, for example,

398

have shown that MgO particles quickly convert to Mg(OH)2 when suspended in water. The subsequent

399

precipitation of struvite in the presence of ammonium and phosphate was found to be largely

400

controlled by the dissolution of the newly formed Mg(OH)2 demonstrating that the new knowledge

401

gained from our study is readily transferable to such MgO-based struvite crystallization approaches.

402 403

Supporting Information. Additional AFM image sequences of brucite dissolution, thermodynamic

404

calculations (solution speciation, saturation indices), chemical analyses of fluid composition, XRD

405

analyses of reacted brucite powders and further experimental details. This material is available free of

406

charge via Internet at http://pubs.acs.org.

407 408

Acknowledgements. The authors thank V. Rapelius for help with ICP-OES analyses and A. M.

409

Schleicher for help with XRD analyses. The financial support from the Helmholtz Recruiting initiative

410

provided to Liane G. Benning and JH is kindly acknowledged. CVP acknowledges funding received

411

through the European Union Marie Curie initial training networks, CO2React and Flowtrans.

412 413

References.

414

(1)

surface soils. Nature 1964, 201 (4916), 321–322.

415 416

Bromfield, S. M. Relative contribution of iron and aluminium in phosphate sorption by acid

(2)

Wang, L.; Ruiz-Agudo, E.; Putnis, C. V.; Menneken, M.; Putnis, A. Kinetics of calcium 16 ACS Paragon Plus Environment

Page 16 of 30

Page 17 of 30

Environmental Science & Technology

417

phosphate nucleation and growth on calcite: implications for predicting the fate of dissolved

418

phosphate species in alkaline soils. Environ. Sci. Technol. 2012, 46 (2), 834–842.

419

(3)

for thought. Glob. Environ. Chang. 2009, 19 (2), 292–305.

420 421

(4)

Heffer, P.; Prud’homme, M. Fertilizer Outlook 2015-2019. In 83rd IFA Annual Conference; International Fertilizer Industry Association, Istanbul (Trukey), 2015.

422 423

Cordell, D.; Drangert, J.-O.; White, S. The story of phosphorus: global food security and food

(5)

Kataki, S.; West, H.; Clarke, M.; Baruah, D. C. Phosphorus recovery as struvite from farm,

424

municipal and industrial waste: feedstock suitability, methods and pre-treatments. Waste

425

Manag. 2016, 49, 437–454.

426

(6)

Kataki, S.; West, H.; Clarke, M.; Baruah, D. C. Phosphorus recovery as struvite: recent

427

concerns for use of seed, alternative Mg source, nitrogen conservation and fertilizer potential.

428

Resour. Conserv. Recycl. 2016, 107, 142–156.

429

(7)

nutrient-rich wastewater: a review. Env. Sci Pollut Res Int 2015, 22 (22), 17453–17464.

430 431

Kumar, R.; Pal, P. Assessing the feasibility of N and P recovery by struvite precipitation from

(8)

El Diwani, G.; El Rafie, S.; El Ibiari, N. N.; El-Aila, H. I. Recovery of ammonia nitrogen from

432

industrial wastewater treatment as struvite slow releasing fertilizer. Desalination 2007, 214 (1-

433

3), 200–214.

434

(9)

Uysal, A.; Yilmazel, Y. D.; Demirer, G. N. The determination of fertilizer quality of the

435

formed struvite from effluent of a sewage sludge anaerobic digester. J. Hazard. Mater. 2010,

436

181 (1), 248–254.

437

(10)

precipitation. Waste Manag. 1999, 19 (6), 409–415.

438 439

(11)

Dockhorn, T. About the economy of phosphorus recovery. In International conference on nutrient recovery from wastewater streams; 2009; pp 145–158.

440 441

Li, X. Z.; Zhao, Q. L.; Hao, X. D. Ammonium removal from landfill leachate by chemical

(12)

Quintana, M.; Colmenarejo, M. F.; Barrera, J.; Sánchez, E.; García, G.; Travieso, L.; Borja, R.

442

Removal of phosphorus through struvite precipitation using a by-product of magnesium oxide

443

production (BMP): effect of the mode of BMP preparation. Chem. Eng. J. 2008, 136 (2), 204–

444

209. 17 ACS Paragon Plus Environment

Environmental Science & Technology

445

(13)

separated urine in Nepal. Water Res 2011, 45 (2), 852–862.

446 447

(14)

Gunay, A.; Karadag, D.; Tosun, I.; Ozturk, M. Use of magnesit as a magnesium source for ammonium removal from leachate. J. Hazard. Mater. 2008, 156 (1), 619–623.

448 449

Etter, B.; Tilley, E.; Khadka, R.; Udert, K. M. Low-cost struvite production using source-

(15)

Huang, H. M.; Xiao, X. M.; Yang, L. P.; Yan, B. Removal of ammonium from rare-earth

450

wastewater using natural brucite as a magnesium source of struvite precipitation. Water Sci

451

Technol 2011, 63 (3), 468–474.

452

(16)

Lahav, O.; Telzhensky, M.; Zewuhn, A.; Gendel, Y.; Gerth, J.; Calmano, W.; Birnhack, L.

453

Struvite recovery from municipal-wastewater sludge centrifuge supernatant using seawater NF

454

concentrate as a cheap Mg(II) source. Sep. Purif. Technol. 2013, 108, 103–110.

455

(17)

wastewater by addition of bittern. Chemosphere 2003, 51 (4), 265–271.

456 457

Lee, S. I.; Weon, S. Y.; Lee, C. W.; Koopman, B. Removal of nitrogen and phosphate from

(18)

Matsumiya, Y.; Yamasita, T.; Nawamura, Y. Phosphorus removal from sidestreams by

458

crystallisation of magnesium-ammonium-phosphate using seawater. Water Environ. J. 2000,

459

14 (4), 291–296.

460

(19)

Romero-Guiza, M. S.; Tait, S.; Astals, S.; Del Valle-Zermeno, R.; Martinez, M.; Mata-Alvarez,

461

J.; Chimenos, J. M. Reagent use efficiency with removal of nitrogen from pig slurry via

462

struvite: a study on magnesium oxide and related by-products. Water Res 2015, 84, 286–294.

463

(20)

a precursor: application to wastewater treatment. Chem. Eng. Sci. 2015, 133, 9–15.

464 465

Stolzenburg, P.; Capdevielle, A.; Teychené, S.; Biscans, B. Struvite precipitation with MgO as

(21)

Krahenbuhl, M.; Etter, B.; Udert, K. M. Pretreated magnesite as a source of low-cost

466

magnesium for producing struvite from urine in Nepal. Sci Total Env. 2016, 542 (Pt B), 1155–

467

1161.

468

(22)

Castro, S. R.; Araújo, M. A. C.; Lange, L. C. Evaluation of the hydration process of an

469

industrial magnesia compound to obtain struvite crystals: a technique for recovering nutrients.

470

Rem Rev. Esc. Minas 2015, 68 (1), 77–84.

471 472

(23)

Duckworth, O. W.; Martin, S. T. Dissolution rates and pit morphologies of rhombohedral carbonate minerals. Am. Mineral. 2004, 89, 554–563. 18 ACS Paragon Plus Environment

Page 18 of 30

Page 19 of 30

473

Environmental Science & Technology

(24) Hostetler, P. B.; Coleman, R. G.; Mumpton, F. A. Brucite in alpine serpentinites. Am. Mineral. 1966, 51 (1-2), 75–98.

474 475

(25)

Kawahara, H.; Endo, S.; Wallis, S. R.; Nagaya, T.; Mori, H.; Asahara, Y. Brucite as an

476

important phase of the shallow mantle wedge: evidence from the Shiraga unit of the

477

Sanbagawa subduction zone, SW Japan. Lithos 2016, 254-255, 53–66.

478

(26)

for carbon sequestration. Env. Sci Technol 2013, 47 (1), 126–134.

479 480

(27)

Simandl, G. J.; Paradis, S.; Irvine, M. Brucite – industrial mineral with a future. Geosci. Canada 2007, 34, 57–64.

481 482

Harrison, A. L.; Power, I. M.; Dipple, G. M. Accelerated carbonation of brucite in mine tailings

(28)

Hövelmann, J.; Putnis, C. V.; Ruiz-Agudo, E.; Austrheim, H. Direct nanoscale observations of

483

CO2 sequestration during brucite [Mg(OH)2] dissolution. Environ. Sci. Technol. 2012, 46 (9),

484

5253–5260.

485

(29)

in situ atomic force microscopy observations. Clays Clay Miner. 2006, 54 (5), 598–604

486 487

Kudoh, Y.; Kameda, J.; Kogure, T. Dissolution of brucite and the (001) surface at neutral pH:

(30)

Pokrovsky, O. S.; Schott, J. Experimental study of brucite dissolution and precipitation in

488

aqueous solutions: surface speciation and chemical affinity control. Geochim. Cosmochim.

489

Acta 2004, 68 (1), 31–45.

490

(31)

Parkhurst, D. L.; Appelo, C. A. J. User’s guide to PHREEQC (Version 2) −a computer

491

program for speciation, batch-reaction, one-dimensional transport, and inverse geochemical

492

calculations. U.S. Geological Survey, Water Resources: Denver, CO 1999, p 99−4259.

493

(32)

struvite. Env. Technol 2007, 28 (9), 1015–1026.

494 495

(33)

(34)

500

Fischer, C.; Arvidson, R. S.; Lüttge, A. How predictable are dissolution rates of crystalline material? Geochim. Cosmochim. Acta 2012, 98, 177–185.

498 499

Ruiz-Agudo, E.; Putnis, C. V. Direct observations of mineral fluid reactions using atomic force microscopy: the specific example of calcite. Mineral. Mag. 2012, 76 (1), 227–253.

496 497

Bhuiyan, M. I.; Mavinic, D. S.; Beckie, R. D. A solubility and thermodynamic study of

(35)

Pokrovsky, O. S.; Schott, J.; Castillo, A. Kinetics of brucite dissolution at 25°C in the presence of organic and inorganic ligands and divalent metals. Geochim. Cosmochim. Acta 2005, 69 (4), 19 ACS Paragon Plus Environment

Environmental Science & Technology

905–918.

501 502

(36)

Surfaces A Physicochem. Eng. Asp. 1997, 120, 143–166.

503 504

(37)

Bondietti, G.; Sinniger, J.; Stumm, W. The reactivity of Fe(III) (hydr)oxides: effects of ligands in inhibiting the dissolution. Colloids Surfaces A Physicochem. Eng. Asp. 1993, 19, 157–167.

505 506

Stumm, W. Reactivity at the mineral-water interface: dissolution and inhibition. Colloids

(38)

Kowacz, M.; Putnis, A. The effect of specific background electrolytes on water structure and

507

solute hydration: consequences for crystal dissolution and growth. Geochim. Cosmochim. Acta

508

2008, 72 (18), 4476–4487.

509

(39)

Ruiz-Agudo, E.; Kowacz, M.; Putnis, C. V.; Putnis, A. The role of background electrolytes on

510

the kinetics and mechanism of calcite dissolution. Geochim. Cosmochim. Acta 2010, 74 (4),

511

1256–1267.

512

(40)

Chernov, A. A. Modern Crystallography III: crystal growth; Springer-Verlag: Berlin, 1984.

513

(41)

Rodriguez-Navarro, C.; Burgos Cara, A.; Elert, K.; Putnis, C. V; Ruiz-Agudo, E. Direct

514

nanoscale imaging reveals the growth of calcite crystals via amorphous nanoparticles. Cryst.

515

Growth Des. 2016, 16, 1850–1860.

516

(42)

dolomite dissolution. Mineral. Mag. 2014, 78 (6), 1355–1362.

517 518

(43)

(44)

Ruiz-Agudo, E.; Putnis, C. V; Putnis, A. Coupled dissolution and precipitation at mineral–fluid interfaces. Chem. Geol. 2014, 383, 132–146.

521 522

Putnis, A.; Putnis, C. V. The mechanism of reequilibration of solids in the presence of a fluid phase. J. Solid State Chem. 2007, 180 (5), 1783–1786.

519 520

Putnis, C. V; Ruiz-Agudo, E.; Hövelmann, J. Coupled fluctuations in element release during

(45)

Adnan, A.; Mavinic, D. S.; Koch, F. A. Pilot-scale study of phosphorus recovery through

523

struvite crystallization — II: Applying in-reactor supersaturation ratio as a process control

524

parameter. J. Environ. Eng. Sci. 2003, 2 (6), 473–483.

525

(46)

Béarat, H.; McKelvy, M. J.; Chizmeshya, A. V. G.; Gormley, D.; Nunez, R.; Carpenter, R. W.;

526

Squires, K.; Wolf, G. H. Carbon sequestration via aqueous olivine mineral carbonation : role of

527

passivating layer formation. Environ. Sci. Technol. 2006, 40 (15), 4802–4808.

528

(47)

Harrison, A. L.; Dipple, G. M.; Power, I. M.; Mayer, K. U. Influence of surface passivation and 20 ACS Paragon Plus Environment

Page 20 of 30

Page 21 of 30

Environmental Science & Technology

529

water content on mineral reactions in unsaturated porous media: implications for brucite

530

carbonation and CO2 sequestration. Geochim. Cosmochim. Acta 2015, 148, 477–495.

531

(48)

Harrison, A. L.; Dipple, G. M.; Power, I. M.; Mayer, K. U. The impact of evolving mineral-

532

water-gas interfacial areas on mineral-fluid reaction rates in unsaturated porous media. Chem.

533

Geol. 2016, 421, 65–80.

534 535 536 537

Figure captions.

538

Figure 1: (A-D) Time-lapse in situ AFM deflection images of a brucite surface during dissolution in a

539

flow-through cell. (A) After 22 min in 100 mM NaCl (pH 7). No etch pits are observed. Dissolution

540

occurred via slow retreat of pre-existing step edges. (B - D) Injection of 10 mM (NH4)2HPO4 (pH 8)

541

resulted in the immediate formation and spreading of equilateral triangular etch pits. White arrow in

542

(A-C) serves as reference point. (D) Development of concentric etch pits after 34 min in contact with

543

10 mM (NH4)2HPO4 (pH 8). (E) Depth profile along section a → b as indicated by the dashed line in

544

(D). The height of single steps is ~0.5 nm. (F) Left: the brucite structure projected along the b-axis.

545

The thickness of one Mg(OH)2 layer is 0.47 nm. Right: the brucite structure projected along the c-axis.

546

The morphology and crystallographic orientation of the edges of the triangular etch pits is outlined.

547 548

Figure 2: Measured etch pit spreading rates as a function of (A) total phosphate concentration, (B) pH

549

and (C) H2PO4- activity. (D) Comparison of etch pit spreading rates in NaH2PO4, NH4H2PO4 and

550

(NH4)2HPO4. Values measured on the same surface during a continuous experiment are indicated by

551

same colors. Symbols indicate different salt solutions (squares: NH4H2PO4, diamonds: (NH4)2HPO4,

552

triangles: NaH2PO4). Error bars are standard deviations of the measured values.

553 554

Figure 3: Time-lapse in situ AFM deflection images showing the dissolution of a brucite surface

555

(indicated by the continuous growth of etch pits) and the simultaneous nucleation of a new phase in

556

100 mM (NH4)2HPO4 (pH 8). Fresh solution (2 ml) was injected before each image. Slow growth of 21 ACS Paragon Plus Environment

Environmental Science & Technology

557

nucleated particles is observed (e.g., particle indicated by blue arrows). After a short growth period

558

some particles were detached by the scanning tip (e.g., particle indicated by red arrows) or started to

559

dissolve again (e.g., particle indicated by white arrows).

560 561

Figure 4: (A) AFM deflection image showing globular shaped nanoparticles nucleated on a brucite

562

surface during reaction with 100 mM NH4H2PO4 (pH 8.5). (B) Same area as in (A) after the solution

563

had been kept stagnant for 20 min. Growth of particles is observed, e.g., in the area outlined by the

564

dashed rectangle. (C, D) Height profiles along sections a → b (dashed lines in (A) and (B)) showing

565

the height increase of a nanoparticle from ~8 to ~16 nm. (E) AFM deflection image showing larger

566

particle aggregates formed after 40 min in contact with 100 mM (NH4)2HPO4 (pH 8). (F) Higher

567

magnification AFM deflection image of the area outlined by the dashed square in (E) showing the

568

cluster nature of the larger aggregates. (G) Height profile along section a → b as indicated by the

569

dashed line in (F). The total height of the particle aggregate is ~30 nm.

570 571

Figure 5: AFM deflection images showing precipitates formed on brucite surfaces after longer periods

572

of reaction during ex situ experiments. (A) After ~38 h in 100 mM (NH4)2HPO4 (pH 8). Large particle

573

clusters have formed, mainly along deep steps. (B) After ~14 h in 100 mM NH4H2PO4 (pH 8.5). Some

574

elongated precipitates composed of fused-together nanoglobular particles have formed in areas of high

575

step density. (C) After ~14 h in 100 mM (NH4)2HPO4 (pH 8.5). Some thick precipitates with more

576

well-defined straight edges have formed on the brucite surface. (D-F) Height profiles along sections a

577

→ b as indicated by the dashed lines in (A-C), respectively. Some precipitates have reached

578

thicknesses of more than 200 nm. (G) Higher magnification AFM deflection image of the area

579

outlined by the dashed square in (C) revealing the nanoglobular texture of the precipitated phase.

580 581

Figure 6: (A) SEM image of a brucite surface after ~14 h of reaction in 100 mM (NH4)2HPO4 (pH

582

8.5). Numerous elongated crystals have formed along the step edges. (B) Higher magnification SEM

583

image of the same surface revealing a clear alignment of the lath-shaped crystals on the brucite

584

substrate. Insets display EDX analyses taken from the brucite surface and the newly formed crystals, 22 ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30

Environmental Science & Technology

585

demonstrating that high amounts of P were incorporated in the new phase. (C) Representative Raman

586

spectrum of the new phase (red curve), showing perfect agreement with a reference spectrum for

587

struvite (black curve).

23 ACS Paragon Plus Environment

Environmental Science & Technology

(A-D) Time-lapse in situ AFM deflection images of a brucite surface during dissolution in a flow-through cell. (A) After 22 min in 100 mM NaCl (pH 7). No etch pits are observed. Dissolution occurred via slow retreat of pre-existing step edges. (B - D) Injection of 10 mM (NH4)2HPO4 (pH 8) resulted in the immediate formation and spreading of equilateral triangular etch pits. White arrow in (A-C) serves as reference point. (D) Development of concentric etch pits after 34 min in contact with 10 mM (NH4)2HPO4 (pH 8). (E) Depth profile along section a → b as indicated by the dashed line in (D). The height of single steps is ~0.5 nm. (F) Left: the brucite structure projected along the b-axis. The thickness of one Mg(OH)2 layer is 0.47 nm. Right: the brucite structure projected along the c-axis. The morphology and crystallographic orientation of the edges of the triangular etch pits is outlined. 109x85mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30

Environmental Science & Technology

Measured etch pit spreading rates as a function of (A) total phosphate concentration, (B) pH and (C) H2PO4activity. (D) Comparison of etch pit spreading rates in NaH2PO4, NH4H2PO4 and (NH4)2HPO4. Values measured on the same surface during a continuous experiment are indicated by same colors. Symbols indicate different salt solutions (squares: NH4H2PO4, diamonds: (NH4)2HPO4, triangles: NaH2PO4). Error bars are standard deviations of the measured values. 95x81mm (300 x 300 DPI)

ACS Paragon Plus Environment

Environmental Science & Technology

Time-lapse in situ AFM deflection images showing the dissolution of a brucite surface (indicated by the continuous growth of etch pits) and the simultaneous nucleation of a new phase in 100 mM (NH4)2HPO4 (pH 8). Fresh solution (2 ml) was injected before each image. Slow growth of nucleated particles is observed (e.g., particle indicated by blue arrows). After a short growth period some particles were detached by the scanning tip (e.g., particle indicated by red arrows) or started to dissolve again (e.g., particle indicated by white arrows). 61x27mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 26 of 30

Page 27 of 30

Environmental Science & Technology

(A) AFM deflection image showing globular shaped nanoparticles nucleated on a brucite surface during reaction with 100 mM NH4H2PO4 (pH 8.5). (B) Same area as in (A) after the solution had been kept stagnant for 20 min. Growth of particles is observed, e.g., in the area outlined by the dashed rectangle. (C, D) Height profiles along sections a → b (dashed lines in (A) and (B)) showing the height increase of a nanoparticle from ~8 to ~16 nm. (E) AFM deflection image showing larger particle aggregates formed after 40 min in contact with 100 mM (NH4)2HPO4 (pH 8). (F) Higher magnification AFM deflection image of the area outlined by the dashed square in (E) showing the cluster nature of the larger aggregates. (G) Height profile along section a → b as indicated by the dashed line in (F). The total height of the particle aggregate is ~30 nm. 127x237mm (300 x 300 DPI)

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 5: AFM deflection images showing precipitates formed on brucite surfaces after longer periods of reaction during ex situ experiments. (A) After ~38 h in 100 mM (NH4)2HPO4 (pH 8). Large particle clusters have formed, mainly along deep steps. (B) After ~14 h in 100 mM NH4H2PO4 (pH 8.5). Some elongated precipitates composed of fused-together nanoglobular particles have formed in areas of high step density. (C) After ~14 h in 100 mM (NH4)2HPO4 (pH 8.5). Some thick precipitates with more well-defined straight edges have formed on the brucite surface. (D-F) Height profiles along sections a → b as indicated by the dashed lines in (A-C), respectively. Some precipitates have reached thicknesses of more than 200 nm. (G) Higher magnification AFM deflection image of the area outlined by the dashed square in (C) revealing the nanoglobular texture of the precipitated phase. 82x52mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 28 of 30

Page 29 of 30

Environmental Science & Technology

(A) SEM image of a brucite surface after ~14 h of reaction in 100 mM (NH4)2HPO4 (pH 8.5). Numerous elongated crystals have formed along the step edges. (B) Higher magnification SEM image of the same surface revealing a clear alignment of the lath-shaped crystals on the brucite substrate. Insets display EDX analyses taken from the brucite surface and the newly formed crystals, demonstrating that high amounts of P were incorporated in the new phase. (C) Representative Raman spectrum of the new phase (red curve), showing perfect agreement with a reference spectrum for struvite (black curve). 123x246mm (300 x 300 DPI)

ACS Paragon Plus Environment

Environmental Science & Technology

61x44mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 30 of 30