Reversible, Selective Trapping of Perchlorate from Water in Record

Jan 14, 2016 - ... Low-Dimensional Cationic Coordination Polymer: Overcoming the Hofmeister Bias Selectivity. Lin Zhu , Chengliang Xiao , Xing Dai , J...
1 downloads 11 Views 901KB Size
Subscriber access provided by NEW YORK UNIV

Article

Reversible, selective trapping of perchlorate from water in record capacity by a cationic metal-organic framework Ian R Colinas, Rachel C Silva, and Scott R. J. Oliver Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b03455 • Publication Date (Web): 14 Jan 2016 Downloaded from http://pubs.acs.org on January 16, 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 17

Environmental Science & Technology

Reversible, selective trapping of perchlorate from water in record capacity by a cationic metal-organic framework Ian R. Colinas, Rachel C. Silva and Scott R. J. Oliver* University of California, Santa Cruz Department of Chemistry and Biochemistry, 1156 High Street, Santa Cruz California, 95064, United States

Supporting Information Placeholder

1

ABSTRACT: We report the capture of ppm level aqueous perchlorate in record capacity and kinetics

2

via the complete anion exchange of a cationic metal–organic framework. Ambient conditions were used

3

for both the synthesis of silver 4,4’-bipyridine nitrate (SBN) and the exchange, forming silver 4,4’-

4

bipyridine perchlorate (SBP). The exchange was complete within 90 minutes and the capacity was 354

5

mg/g, representing 99% removal. These values are greater than current anion exchangers such as the

6

resins Amberlite IRA-400 (249 mg/g), Purolite A530E (104 mg/g) and layered double hydroxides (28

7

mg/g). Moreover, unlike resins and layered double hydroxides, SBN is fully reusable and displays 96%

8

regeneration to SBN in nitrate solution, with new crystal formation allowing indefinite cycling for per-

9

chlorate. We show seven cycles as proof of concept. Perchlorate contamination of water represents a

10

serious health threat since it is a thyroid endocrine disruptor. This non-complexing anionic pollutant is

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 17

11

significantly mobile and environmentally persistent. Removal of other anionic pollutants from water

12

such as chromate, pertechnetate or arsenate may be possible by this methodology.

13 14

INTRODUCTION

15

Perchlorate (ClO4–) is an emerging trace contaminant in ground water and has gained significant atten-

16

tion as it has become widespread in many countries including the United States, Japan, Korea, India,

17

Germany and China.1 Even at 1 ppb level, this anionic pollutant can block the uptake of iodide by the

18

thyroid due to their equivalent charge and similar ionic radii. This blockage can disrupt the production

19

of thyroid hormones and affect metabolism, possibly leading to hypothyroidism or mental retardation in

20

fetuses and infants.2 Perchlorate salt is used as the conventional solid oxidant in the industrial manufac-

21

ture of rocket fuel, explosives, flares and fireworks. Bleach and nitrate fertilizers can also contain per-

22

chlorate, and the anion also occurs naturally in arid environments.3

23

Because of the high solubility and non-complexing nature of perchlorate, it is highly mobile in aqueous

24

environments and strongly resistant to traditional wastewater treatment technologies.4 The latter include

25

adsorption, ion exchange, membrane filtration, catalytic reduction and biological remediation. Among

26

these techniques, ion exchange resins are considered the most viable and efficient method.5 These ma-

27

trices consist of a divinylbenzene-crosslinked polystyrene backbone with terminal quaternary ammoni-

28

um cationic side groups that possess an exchangeable counter-anion.6 These resins have been widely

29

studied and optimized for the selective capture of perchlorate from groundwater and significant effort

30

has been directed towards increasing their regenerability.7,8,9 They also have been used in the attempt of

31

separating pollutant anions such as arsenate (AsO43–).10 These resins, however, are of limited thermal

32

and chemical stability due to their organic polymeric nature. For example, approximately 15% of their

33

ion exchange capacity is lost upon contact with water at 90 °C, and 64% and 57% of their ion exchange

34

capacity in the presence of sodium hypochlorite and hydrogen peroxide, respectively.11 A recent attempt

ACS Paragon Plus Environment

2

Page 3 of 17

Environmental Science & Technology

35

to separate perchlorate from water using an organic resin was reported by Song et al.12 It was observed

36

that such resins are significantly sensitive to pH. A perchlorate adsorption capacity of 170.4 mg/g

37

(1.713 meq/g) could only be obtained upon treatment with a resin bed depth of 3.4 cm in a solution with

38

an adjusted pH of 7, and the resin could only be effectively regenerated via concentrated hydrochloric

39

acid treatment.

40

Layered double hydroxides (LDHs) are an isostructural set of materials consisting of cationic brucite-

41

type layers that are charge balanced by interlayer anions with the general formula [M2+1-xM3+x(OH)2]An–

42

x/n·mH2O.

43

group of cationic materials, however, has limited capacity as evidenced by adsorption titration and iso-

44

therms. In a recent study, the nitrate form of Mg:Al LDH had a perchlorate adsorption capacity of only

45

1.959 mg/g (0.0197 meq/g).14 In addition, selectivity of LDHs is low toward target anions due to the

46

interfering effect of competing anions. Common anions such as carbonate, sulfate and chloride tend to

47

de-intercalate some or all of the trapped anions.15

48

Metal-organic frameworks (MOFs, also known as coordination polymers) are an emerging class of ma-

49

terials with a now vast array of topologies and properties of interest including gas storage,16 gas separa-

50

tion,17 drug delivery18 and ion exchange.19 Cationic MOFs are a subgroup of these compounds, where

51

positive metal centers are connected by neutral organic linkers and anions reside in the pores. If the

52

host-guest interaction is weak, anion exchange may occur if the framework does not collapse in the pro-

53

cess.20 Feng et al. recently demonstrated a selective anion exchange separation process using a cationic

54

indium-based MOF [In3O(COO)6+]. The material reversibly captures organic dye anions of varying

55

charge from a dimethylformamide (DMF) solution.21 Custelcean and coworkers have developed an ani-

56

on separation approach in aqueous systems by the selective crystallization of chloride, bromide or io-

57

dide into one-dimensional cationic MOFs, even in the presence of competing oxo-anions such as nitrate

58

and sulfate. The anion selectivity is governed by the anion size, coordinating ability and the overall

59

packing efficiency of the MOF.22,23

They have been studied extensively as a possible alternative to anion exchange resins.13 This

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 17

60

Ag(I) is an ideal metallic node for cationic MOF construction due to its high affinity for N- and O-

61

donors and flexible coordination number/geometry of two to eight.24 In fact, silver coordination poly-

62

mers have been previously studied for their ability to selectively trap anion via a single-crystal-to-

63

single- crystal (SC-SC) transition.25,26 Our group reported a cationic framework consisting of chains of

64

alternating Ag(I) and 4,4'-bipyridine that π-π stack into cationic layers. The layers are charge balanced

65

by ethanedisulfonate (EDS) anions and selectively exchange for permanganate and perrhenate, with a

66

significantly high adsorption capacity of 292 and 602 mg/g, respectively.27 We also reported an isostruc-

67

tural solid solution material consisting of Co(II) and Zn(II), which selectively captures chromate at a

68

capacity of 68.5mg/g.28

69

Herein, we report the perchlorate trapping profile of the cationic MOF [Ag-bipy+][NO3–] (SBN).29 The

70

material releases nitrate  an essential component of fertilizers whose maximum concentration level

71

(MCL) in water is 10 ppm  in exchange for perchlorate, a highly toxic anion with an MCL of 6 ppb

72

and is present in polluted groundwater from 10 ppb to 6 ppm.30 Hence, the exchange for perchlorate

73

would result in a nitrate concentration below the regulatory level and such treated water could potential-

74

ly be incorporated into irrigation systems. Unlike urea or ammonium, nitrate is immediately available

75

as a nutrient that can be readily absorbed by plants at high rate and concentration up to 100 mM, with

76

excess nitrate stored in the vacuole.31 In addition, SBN traps perchlorate quantitatively in far greater

77

capacity, selectivity, speed and reversibility compared to existing resins and LDHs, offering an alterna-

78

tive solution to the ongoing issue of perchlorate pollution.

79 80

MATERIALS AND METHODS

81

Reagents. Silver nitrate (AgNO3, Fisher, 99%), 4,4'-bipyridine [(C5H4N)2, Acros Organics, 98%] were

82

used as-received for the synthesis. Sodium nitrate (NaNO3, Fisher, 99%), sodium perchlorate monohy-

83

drate (NaClO4·H2O, Fluka Analytical, 98%), Purolite A530E (Purolite), Amberlite IRA-400 resin (Sig-

ACS Paragon Plus Environment

Page 5 of 17

Environmental Science & Technology

84

ma-Aldrich) and hydrotalcite [Mg6Al2(CO3)(OH)16·4H2O, Aldrich Chemistry] were used as-purchased

85

for the anion exchange reactions.

86

Synthesis. Pale gray parallelepiped crystals of SBN 400 to 600 µm in length were synthesized hydro-

87

thermally (HT-SBN, Figure S1). Alternatively, room temperature conditions (RT-SBN) yielded a micro-

88

crystalline powdered material consisting of 4 to 10 µm average sized blocks (Figure S2a). For HT-SBN,

89

a mixture of AgNO3 (0.1 g, 0.59 mmol), 4,4'-bipyridine (0.1 g, 0.64 mmol) and deionized water (10 mL)

90

was stirred at room temperature for 10 min and then transferred to a 15 mL Teflon lined autoclave to 2/3

91

filling. The autoclave was placed in a programmable oven and heated at 140 °C for 10 h, then cooled to

92

110 °C for 8 h, followed by further cooling to 90 °C for 6 h and finally cooling to room temperature at a

93

rate of 0.1 °C/min. Pale gray crystals were isolated after filtration and rinsed with water and acetone

94

(yield: 0.181 g, 94.3% based on silver nitrate). The synthesis of RT-SBN crystals was carried out by

95

simply stirring the reactants in the same ratio for 3 days in a sealed beaker and filtered in the same man-

96

ner (yield: 0.187 g, 97.6% based on silver nitrate).

97

Perchlorate Exchange. Batch experiments were carried out under ambient conditions by simply plac-

98

ing 80 mg (0.25 mmol) of as-synthesized SBN material previously ground using a mortar and pestle into

99

a beaker containing 50 mL deionized water and NaClO4 (35 mg, 0.25 mmol). The anion exchange solu-

100

tion was sealed and stirred mildly. The exchange solution was analyzed by taking liquid aliquots at var-

101

ious time intervals to quantify the residual perchlorate concentration by ion chromatography (IC). The

102

post-exchange crystalline SBP product was recovered by vacuum filtration and rinsed with water and

103

acetone prior analysis by powder X-ray diffraction (PXRD).

104

Selectivity Tests. Selectivity batch tests were performed by introducing 65.2 mg (0.2 mmol) of SBN

105

material into a sealed beaker containing 50 ml H2O and 42 mg (0.3 mmol) NaClO4. The beaker also

106

contained either 0.84 g (10 mmol) NaHCO3 or 1.42 g (10 mmol) NaSO4 (fifty-fold molar excess). The

107

selectivity reactions were mildly stirred for 2 h under ambient conditions and the filtrate was analyzed

108

quantitatively by ion chromatography (IC) to quantify the residual perchlorate concentration.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 17

109

SBN Regeneration. Upon completion of ion exchange, the solid crystalline SBP material was placed

110

into a 0.1 M NaNO3 solution and mechanically stirred in an oil bath at 70 °C. The percent regeneration

111

was evaluated by measuring the increase of perchlorate concentration in the regeneration solution via IC

112

(Figure S4) and by the PXRD peak area of the main (002) and (100) peak for SBN and SBP, respective-

113

ly.

114

Characterization. PXRD was measured on a Rigaku Americas Miniflex Plus diffractometer, scanning

115

from 2 to 40 ° (2θ) at a rate of 2 °·min–1 with a 0.04 ° step size under Cu Kα radiation (λ = 1.5418 Å).

116

IC analysis was performed to assess perchlorate concentration using a Dionex ICS-3000 with an IonPac

117

AS20 column and a detection limit of 3 µg/L (ppb). Scanning electron microscopy (SEM) data were

118

collected with a FEI Quanta 3D Dualbeam microscope.

119 120 121

RESULTS AND DISCUSSION

122

SBN Synthesis and Characterization. Crystals of SBN can be synthesized hydrothermally or at room

123

temperatures. Hydrothermal conditions produced pale gray crystals of uniform size ranging from 400-

124

600 µm with a parallelepiped morphology (Figure S1). Room temperature synthesis of SBN resulted in

125

smaller crystals with average width of 4-10 µm (Figure S2a). Both synthetic methods result in highly

126

crystalline materials with yields of 94.3% and 97.6% for the hydrothermal and room temperature routes,

127

respectively. PXRD confirms that both methods yield the identical phase and match the theoretical pat-

128

tern based on the single crystal solution (Figure 1a,b).29,32

129 130

Perchlorate Capture. We performed anion exchange batch tests to capture perchlorate through a sol-

131

vent-mediated recrystallization process described by Schröder et al. for related metal bipyridine and py-

132

razine coordination polymers.33 SBN has weakly bound nitrate anions between the cationic Ag(I)-bipy+

133

chains (vide infra) and its solubility in water (0.80 mmol• L–1) is much greater than that of SBP (0.19 ACS Paragon Plus Environment

Page 7 of 17

Environmental Science & Technology

134

mmol•L–1).33 For the SBN material, the adsorption capacity was 354.0 mg ClO4–/g SBN, accounting for

135

99% ClO4– removal based on the theoretical adsorption capacity (Table 1). The PXRD of the solid ma-

136

terial after exchange matches the theoretical structure of SBP (Figure 1c). SEM shows the partial mor-

137

phology change from SBN blocks to SBP needles after 20 min of exposure to perchlorate (Figure S2b),

138

with only SBP needles present after 70 min (Figure S2c). PXRD confirms that both structures are pre-

139

sent after 20 minutes of exchange and SBP is phase-pure after 70 min (Figure S3).

140 141

Figure 1. PXRD patterns: (a) Room Temperature SBN; (b) Hydrothermal SBN; (c) SBP.

142 143

Selectivity Tests and Kinetics. In addition to the high adsorption capacity and ambient conditions for

144

both the synthesis and exchange, SBN exhibits highly selective perchlorate capture in the presence of

145

multiple-fold excess of potentially competing common anions. The issue of selectivity is an ongoing

146

problem for LDHs due to the high affinity of the cationic layers for carbonate as well as other anions

147

including sulfate, hydroxide and chloride.15 The selectivity of SBN for perchlorate uptake over car-

148

bonate and sulfate was demonstrated by performing two separate batch tests containing 50-fold molar

149

excess of each anion. In both tests it was found that the competing anions did not interfere with the up-

150

take of perchlorate due to both its lower relative hydration energy and lower solubility of the complex.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 17

151

The adsorption capacities remained at the high levels of 337 and 342 mg ClO4–/g SBN for the carbonate

152

and sulfate batch tests, representing 95% and 97% perchlorate uptake, respectively.

153

The kinetics of perchlorate uptake by SBN were analyzed and compared to the performance of the

154

commercially available anion exchange resins Amberlite IRA-400 and Purolite A530E, as well as the

155

calcined and uncalcined forms of hydrotalcite [Mg6Al2(CO3)(OH)16·4H2O] and Ni3Al-LDH. The latter

156

was tested because it adsorbs in higher capacity than hydrotalcite for some anions including pertechne-

157

tate (TcO4–).34 All batch experiments were performed under the same conditions using equivalent stoi-

158

chiometric amounts of the anion exchanger. The kinetic plots underscore the superior performance of

159

SBN (Figure 2). A rapid decrease of approximately 98% of the perchlorate concentration was observed

160

within 60 min for SBN. This rapid perchlorate trapping can be further understood by following the ki-

161

netics of the nitrate released by SBN (Figure S4). The plot reaches a plateau after 90 min corresponding

162

to the stoichiometric amount of perchlorate exchanged into SBP. In the case of the other anion ex-

163

changers, only Amberlite showed a significant perchlorate uptake in 60 min, with 76% removed. After

164

90 min, the values were 100% for SBN and 78% for Amberlite. Both calcined hydrotalcite and Ni3Al-

165

LDH display a low per chlorate uptake of only 4 to 12%. This lower affinity for the LDHs is in part due

166

to the fact that we did not exclude carbonate from the exchange solutions (always present from atmos-

167

pheric CO2). Carbonate is known to lower anion capacity for LDHs and is in fact used to regenerate

168

exchanged LDH.13 The meq/g adsorption capacities for each of the anion exchangers are summarized in

169

Table 1.

ACS Paragon Plus Environment

Page 9 of 17

Environmental Science & Technology

170 171

Figure 2. Perchlorate uptake per gram of solid versus time: SBN (black squares), Amberlite (red

172

squares), Purolite (purple stars), uncalcined hydrotalcite (magenta triangles), calcined hydrotalcite (pur-

173

ple stars), uncalcined Ni3Al-LDH (blue triangles) and calcined Ni3Al-LDH (cyan diamonds).

174 175 176

Table 1. Perchlorate exchange capacities of the various cationic host materials tested



Anion Exchanger





% ClO4

mg ClO4 /g

mol ClO4 /mol

meq/g

SBN

353.97

1.062

3.569

99%

Uncalcined Hydrotalcite-LDH

10.62

0.065

0.107

3.2%

ACS Paragon Plus Environment

removal

a

Environmental Science & Technology

Page 10 of 17

Calcined Hydrotalcite-LDH

47.51

0.289

0.478

11%

Uncalcined Ni3AlLDH

5.85

0.048

0.059

1.6%

Calcined Ni3Al-LDH

24.0

0.195

0.241

5.1%

Amberlite IRA-400 Resin

248.87

N/A

2.502

78%

Purolite A530E Resin

103.88

N/A

1.044

30%

177 178

a

Based on the theoretical capacity of the starting solid.

179 180

Recyclability. For any industrial application, the recyclability of the anion exchanger is a critical issue.

181

The material after anion exchange must be easily recoverable and retain its structure and particle size

182

after multiple regeneration cycles. LDHs often form a gel and require ca. 30 min of centrifuging for

183

total separation from solution, while resins require strong acid or brine and form a slurry. In contrast,

184

SBP is highly crystalline after anion exchange (Figure S2c) and is easily separated from solution by

185

vacuum filtration. In order for the MOF to reversibly exchange its anions by a solvent mediated pro-

186

cess, it must have similar solubility for its original and exchanged forms.33 The lower solubility of SBP

187

must be overcome by increasing the nitrate to perchlorate concentration ratio as well increasing the

188

temperature to shift the equilibrium and allow SBP to dissolve into its individual components and re-

189

crystallize back to SBN. Indeed, the solid was quantitatively regenerated to phase-pure SBN simply by

190

stirring in a 20-fold molar excess 0.1 M sodium nitrate salt solution at 70 °C for 24 h. The evolved per-

191

chlorate as determined by IC showed that the SBN material regenerated an average of 96% (Figure S6). ACS Paragon Plus Environment

Page 11 of 17

Environmental Science & Technology

192

This perchlorate-nitrate cycling shows continued excellent reversibility on further cycles (Figure 3) and

193

IC data confirms complete exchange (Table S1). We show seven cycles for proof of concept. New SBN

194

crystals are formed upon regeneration (Figure S2d), meaning fresh crystals are formed on each cycle for

195

continued recyclability (Figures 3, S7). On the other hand, the Purolite resin displayed negligible recy-

196

clability upon treatment with 10% hydrochloric acid solution after perchlorate uptake. The theoretical

197

capacity per gram of material per cycle is 0.354 g of perchlorate, which would correspond to the ability

198

to treat (for example) 17,700 L of 20 ppb perchlorate contaminated water. In the case of the Amberlite

199

resin, its theoretical capacity per cycle of 0.249 g perchlorate would treat only 12,400 L of 20 ppb con-

200

taminated water and require concentrated brine solution or mineral acid treatment for each regeneration

201

cycle. The superior performance of SBN/SBP can be seen in a plot of the uptake capacity versus cycle

202

number (Figure 4). Although a mass loss of ~ 5% was encountered after each cycle in the process of

203

recovering the adsorbent material, such mechanical loss is attributed to the material lost in the filter pa-

204

per while performing vacuum filtration. Further, SBP crystallizes immediately on the surface of the

205

SBN block crystals, giving rise to needles that elongate from the surface of the SBN blocks (Figure

206

S2b).

207 208

Figure 3. PXRD shows the excellent cyclability of [Ag-bipy+] over seven cycles for proof of concept. ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 17

209 210

Figure 4. Plot of perchlorate uptake capacity versus cycle number, underscoring the superior perfor-

211

mance of SBN versus Amberlite.

212 213

In addition to these highly desirable exchange properties, the material must be able to capture the dilute

214

perchlorate of wastewater As an initial text of lower concentration, we performed exchange at 5 ppm

215

perchlorate, in the range of contaminated underground water plumes. 60 mg of SBN exhibited a 100%

216

perchlorate removal over 5 days (Figure S5). The perchlorate uptake is much faster rate and at a deeper

217

level for our material compared to resins and LDHs (Figure 2). These features make it possible to use

218

the material in a large stationary container, the common method for perchlorate treatment. Regeneration

219

can then occur offline at room temperature in excess nitrate solution at 70 °C condition over the longer

220

required time frame (Figure S6).

221 222

ACS Paragon Plus Environment

Page 13 of 17

Environmental Science & Technology (a)

(b)

223 224

Figure 5. Crystallographic views of two [Ag-bipy+] "layers" (top in red, bottom in green; corresponding

225

side views shown at top; anions and hydrogens omitted for clarity). (a) SBN, [Ag-bipy+][NO3–], where

226

the layers are oriented ~ 90° to each other and only half the pyridine rings π -stack to the adjacent layer;

227

(b) SBP, [Ag-bipy+][ClO4–], with the viewing angle offset slightly to emphasize the π –stacking of all

228

pyridine rings.

229 230

Structural Considerations. Fortunately, [Ag-bipy+] displays greater preference for perchlorate than

231

nitrate, allowing for the rapid trapping of ppm level perchlorate from contaminated water. Regeneration

232

of the nitrate form can occur offline in excess nitrate to obtain a concentrated perchlorate wasteform.

233

The anion preference can be understood by considering the SBN and SBP crystal structures (Figures S8,

234

S9). For both, each Ag(I) is chelated by the nitrogens of two different (µ-2)-4,4'-bipyridine units, form-

235

ing extended polymeric chains. These chains align parallel into a non π-stacked “layer” where the rings

236

are all co-planar with the plane of the page (red atoms, Figure 5). The two structures differ in the ar-

237

rangement of the next “layer” (green atoms, Figure 5). For SBN, the next layer is rotated by 90° (Figure

238

5a) whereas for SBP the layers are eclipsed (Figure 5b). For SBN, only half of the pyridine rings π-

239

stack in a staggered manner to the adjacent layer (Figure 5a, average distance 3.55 Å) whereas all pyri-

240

dine rings π-stack and are eclipsed for SBP (Figure 5b, average distance 3.47 Å). For SBN, half of the

241

silver centers dimerize [Ag-Ag distance 2.97(1) Å]25 and cross-link to the next layer to define a 3-D

242

MOF (Figure 5a, top). For SBP, the silvers have a long contact of 3.60(1) Å, which is well outside the ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 17

243

median covalent bond length of 3.047±0.193 Å by a search of Conquest/Vista in the Cambridge Crystal

244

Structure Database (CSSD).35

245

These two structures transform reversibly and are a consequence of the shape of the incoming anion that

246

must pack between the cationic polymers (Figure 5: side view of “layers” without anions). Nitrate is

247

flat while perchlorate is tetrahedral (spheroidal). Both structures are stoichiometrically equivalent, with

248

one mole of the monomeric anion in their formulae. The anion oxygen to silver distances are typical of

249

ionic bonding [2.826(17) Å and 2.782(10) Å for SBN; 2.863(1) Å and 2.945(15) Å for SBP]. SBN,

250

however, has half the π -stacking than SBP (Figure 5). SBP is therefore more stable and accounts for

251

the faster exchange of nitrate for very dilute perchlorate. Re-intercalation of nitrate requires greater

252

concentration and longer time, where the polymers reassemble around the incoming anion to give quan-

253

titative exchange and fresh crystals for continuous exchange cycles.

254

In summary, as-prepared SBN displays outstanding propensity towards the toxicologically important

255

pollutant perchlorate. The capture capacity, kinetics, selectivity, regenerability and cyclability are all

256

unprecedented. The regeneration method actually takes advantage of the metastability of MOFs by re-

257

assembling back into the regenerated form. The process is environmentally acceptable since the MCL

258

of nitrate is 10 ppm. It is possible that other problematic elements that also occur in their oxo-anion

259

form in water may be able to be trapped in this manner, such as chromium-6, arsenite, selenium and ac-

260

tinates. We are currently scaling up the synthesis from the gram to kilogram level for testing on actual

261

perchlorate contaminated at a nearby Superfund site. Scaling up of the synthesis and perchlorate uptake

262

should be straightforward because both occur in unbuffered water, at room temperature and even un-

263

stirred conditions. Overall, this discovery is a significant advance in the methods available for perchlo-

264

rate trapping and water purification.

265 266

ASSOCIATED CONTENT

267

Supporting Information ACS Paragon Plus Environment

Page 15 of 17

Environmental Science & Technology

268

Optical micrograph and SEM images before and after anion exchange, PXRD of perchlorate trapping

269

kinetics, table with adsorption capacities at each cycle, SBN regeneration plot and crystallographic

270

views of SBN and SBP. This material is available free of charge via the Internet at http://pubs.acs.org.

271 272

AUTHOR INFORMATION

273

Corresponding Author

274

*Phone: +1-831-459-5448; Fax: +1-831-459-2935 ; email: [email protected].

275

Author Contributions

276

All authors contributed equally.

277

Notes

278

The authors declare no competing financial interests.

279 280

ACKNOWLEDGMENTS

281

The PXRD data in this work were recorded on an instrument supported by the NSF Major Research In-

282

strumentation (MRI) Program under Grant DMR-1126845. We acknowledge assistance of Rob Franks

283

with ion chromatography (IC) in the Department of Earth and Marine Sciences at UC Santa Cruz. We

284

also acknowledge Dr. Tom Yuzvinsky and the W.M. Keck Center for Nanoscale Optofluidics at UC San-

285

ta Cruz for the use and assistance in image acquisition on the FEI Quanta 3D Dualbeam microscope.

286 287

REFERENCES

288 289 290 291 292

(1) Yang, Y.; Gao, N.; Chu, W.; Zhang, Y.; Ma, Y. Adsorption of perchlorate from aqueous solution by the calcination product of Mg/(Al–Fe) hydrotalcite-like compounds. J. Hazard. Mater. 2012, 209-210, 318–325. (2) Srinivasan, A.; Viraraghavan, T. Perchlorate: Health Effects and Technologies for Its Removal from Water Resources. Int. J. Environ. Res. Public. Health 2009, 6 (4), 1418–1442. ACS Paragon Plus Environment

Environmental Science & Technology

293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341

Page 16 of 17

(3) Dasgupta, P. K.; Martinelango, P. K.; Jackson, W. A.; Anderson, T. A.; Tian, K.; Tock, R. W.; Rajagopalan, S. The Origin of Naturally Occurring Perchlorate: The Role of Atmospheric Processes. Environ. Sci. Technol. 2005, 39 (6), 1569–1575. (4) Urbansky, E. T.; Schock, M. R. Issues in managing the risks associated with perchlorate in drinking water. J. Environ. Manage. 1999, 56 (2), 79–95. (5) Praharaj, S.; Nath, S.; Panigrahi, S.; Ghosh, S. K.; Basu, S.; Pande, S.; Jana, S.; Pal, T. Layer-byLayer Deposition of Bimetallic Nanoshells on Functionalized Polystyrene Beads. Inorg. Chem. 2006, 45 (4), 1439–1441. (6) Mark, R.; Findley, W. N. Thermal expansion instability and creep in amine-cured epoxy resins. Polym. Eng. Sci. 1978, 18 (1), 6–15. (7) Gu, B.; Brown, G. M.; Chiang, C.-C. Treatment of Perchlorate-Contaminated Groundwater Using Highly Selective, Regenerable Ion-Exchange Technologies. Environ. Sci. Technol. 2007, 41 (17), 6277–6282. (8) Gu, B.; Ku, Y.-K.; Brown, G. M. Sorption and Desorption of Perchlorate and U(VI) by StrongBase Anion-Exchange Resins. Environ. Sci. Technol. 2005, 39 (3), 901–907. (9) Gu, B.; Brown, G. M.; Maya, L.; Lance, M. J.; Moyer, B. A. Regeneration of Perchlorate (ClO 4 )-Loaded Anion Exchange Resins by a Novel Tetrachloroferrate (FeCl 4 - ) Displacement Technique. Environ. Sci. Technol. 2001, 35 (16), 3363–3368. (10) Dambies, L. Existing and Prospective Sorption Technologies for the Removal of Arsenic in Water. Sep. Sci. Technol. 2005, 39 (3), 603–627. (11) Neagu, V.; Bunia, I.; Plesca, I. Ionic polymers VI. Chemical stability of strong base anion exchangers in aggressive media. Polym. Degrad. Stab. 2000, 70 (3), 463–468. (12) Song, W.; Xu, X.; Tan, X.; Wang, Y.; Ling, J.; Gao, B.; Yue, Q. Column adsorption of perchlorate by amine-crosslinked biopolymer based resin and its biological, chemical regeneration properties. Carbohydr. Polym. 2015, 115, 432–438. (13) Layered double hydroxides: present and future; Rives, V., Ed.; Nova Science Publishers: Huntington, N.Y, 2001. (14) Kim, J. Y.; Komarneni, S.; Parette, R.; Cannon, F.; Katsuki, H. Perchlorate uptake by synthetic layered double hydroxides and organo-clay minerals. Appl. Clay Sci. 2011, 51 (1-2), 158–164. (15) Goh, K.-H.; Lim, T.-T.; Dong, Z. Application of layered double hydroxides for removal of oxyanions: A review. Water Res. 2008, 42 (6-7), 1343–1368. (16) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal–Organic Frameworks. Chem. Rev. 2012, 112 (2), 724–781. (17) Wang, B.; Côté, A. P.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. Colossal cages in zeolitic imidazolate frameworks as selective carbon dioxide reservoirs. Nature 2008, 453 (7192), 207–211. (18) Long, J. R.; Yaghi, O. M. The pervasive chemistry of metal–organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1213. (19) Custelcean, R.; Moyer, B. A. Anion Separation with Metal–Organic Frameworks. Eur. J. Inorg. Chem. 2007, 2007 (10), 1321–1340. (20) Oliver, S. R. J. Cationic inorganic materials for anionic pollutant trapping and catalysis. Chem. Soc. Rev. 2009, 38 (7), 1868. (21) Zhao, X.; Bu, X.; Wu, T.; Zheng, S.-T.; Wang, L.; Feng, P. Selective anion exchange with nanogated isoreticular positive metal-organic frameworks. Nat. Commun. 2013, 4. (22) Custelcean, R.; Haverlock, T. J.; Moyer, B. A. Anion Separation by Selective Crystallization of Metal−Organic Frameworks. Inorg. Chem. 2006, 45 (16), 6446–6452. (23) Rajbanshi, A.; Moyer, B. A.; Custelcean, R. Sulfate Separation from Aqueous Alkaline Solutions by Selective Crystallization of Alkali Metal Coordination Capsules. Cryst. Growth Des. 2011, 11 (7), 2702–2706.

ACS Paragon Plus Environment

Page 17 of 17

342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372

Environmental Science & Technology

(24) Yang, T.; Wang, J.; Wang, J.-G.; Wang, S.; Xiao, H.-P. Three silver(I) coordination polymers assembled from 2-sulfoterephthalic acid and N-donor ligands: syntheses, crystal structures, and luminescence properties. Transit. Met. Chem. 2013, 38 (8), 865–871. (25) Li, C.-P.; Chen, J.; Guo, W.; Du, M. Anion-directed assembly and crystal transformation of Ag(I) coordination polymers with a versatile tripyridyltriazole ligand 3,4-bis(2-pyridyl)-5-(4-pyridyl)1,2,4-triazole. J. Solid State Chem. 2015, 223, 95–103. (26) Du, M.; Zhao, X.-J.; Guo, J.-H.; Batten, S. R. Direction of topological isomers of silver(i) coordination polymers induced by solvent, and selective anion-exchange of a class of PtS-type host frameworks. Chem. Commun. 2005, No. 38, 4836. (27) Fei, H.; Bresler, M. R.; Oliver, S. R. J. A New Paradigm for Anion Trapping in High Capacity and Selectivity: Crystal-to-Crystal Transformation of Cationic Materials. J. Am. Chem. Soc. 2011, 133 (29), 11110–11113. (28) Fei, H.; Han, C. S.; Robins, J. C.; Oliver, S. R. J. A Cationic Metal–Organic Solid Solution Based on Co(II) and Zn(II) for Chromate Trapping. Chem. Mater. 2013, 25 (5), 647–652. (29) Yaghi, O. M.; Hailian, L. T-Shaped Molecular Building Units in the Porous Structure of Ag (4, 4’bpy)⊙ NO3. Journal of the American Chemical Society 1996, No. 118.1, 295–296. (30) US EPA, O. Basic Information about Nitrate in Drinking Water http://water.epa.gov/drink/contaminants/basicinformation/nitrate.cfm (accessed Oct 31, 2015). (31) Crawford, N. M.; Glass, A. D. . Molecular and physiological aspects of nitrate uptake in plants. Trends Plant Sci. 1998, 3 (10), 389–395. (32) Wang, L.-S.; Zhang, J.-F.; Yang, S.-P. catena -Poly[[silver(I)-µ-4,4′-bipyridine-κ 2 N : N ′] perchlorate]. Acta Crystallogr. Sect. E Struct. Rep. Online 2004, 60 (10), m1484–m1486. (33) Cui, X.; Khlobystov, A. N.; Chen, X.; Marsh, D. H.; Blake, A. J.; Lewis, W.; Champness, N. R.; Roberts, C. J.; Schröder, M. Dynamic Equilibria in Solvent-Mediated Anion, Cation and Ligand Exchange in Transition-Metal Coordination Polymers: Solid-State Transfer or Recrystallisation? Chem. - Eur. J. 2009, 15 (35), 8861–8873. (34) Wang, Y.; Gao, H. Compositional and structural control on anion sorption capability of layered double hydroxides (LDHs). J. Colloid Interface Sci. 2006, 301 (1), 19–26. (35) Allen, F. H. The Cambridge Structural Database: a quarter of a million crystal structures and rising. Acta Crystallogr. B 2002, 58 (3), 380–388.

373 374 TOC Art

ACS Paragon Plus Environment