Supplementary Data - Springer Static Content Server

advertisement

1 Journal of Nanoparticle Research

2

Supplementary Material

3

4

Facilitated transport of titanium dioxide nanoparticles by humic

5

substances in saturated porous media under acidic conditions

6

7 Ruichang Zhang a, c

, Haibo Zhang b

, Chen Tu b

, Xuefeng Hu

b

, Lianzhen Li

b

,

8 Yongming Luo a, b,

*

9

10 a

Key Laboratory of Soil Environment and Pollution Remediation, Institute of Soil

11 Science, Chinese Academy of Sciences, Nanjing 210008, China

12 b

Key Laboratory of Coastal Environmental Processes and Ecological Remediation,

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

14 264003, China

15 c

University of Chinese Academy of Sciences

16

17

18

19 *Corresponding author. Tel./fax: +86-535-2109 007. E-mail: ymluo@yic.ac.cn (Luo

20 YM).

1

21

22

23

1500

1000

500

Anatase

0

10 20 30 40

2



50 60 70 80

Fig. S1 XRD pattern of TiO

2

NPS in the present study

2

24

25

60

40

20

0

-20

-40

TiO

2

NPs

HS

Quartz

-60

2 3 4 5 pH

6 7 8

Fig. S2 Zeta potentials of bare TiO

2

NPs, HS and quartz in deionized water

3

26

27

3455

2937

2874

937

1106

1036

1387

1631

4000 3500 3000 2500 2000 1500 1000 500

Wave number (cm

-1

)

Fig. S3 FTIR spectrum of HS in this study

4

28

29

Table S1 The assignmen of absorption bands in FTIR spectrum of HS

Wavenumber (cm -1 )

3455

2937

2874

1631

1387

1106

1036

937

Assignment

OH stretching vibration

Aliphatic CH

2

stretching vibration

CH

2

symmetric stretching vibration

Aromatic C=C stretching vibration

COO symmetric stretching vibration

Aliphatic C-OH stretching vibration

C-O-C stretching

Aliphatic C-C vibration

5

30

31

1.0

y=0.039x+0.006

R

2

=0.999

0.8

0.6

y=0.011x+0.003

R

2

=0.999

0.4

0.2

TiO

2

NPs at 343 nm

HS at 228 nm

0.0

0 20 40

Concentration (mg L

-1

)

60 80

Fig. S4 Calibration curves for UV light absorbance of TiO

2

NPs and HS

6

32 Table S2 Physiochemical parameters of transport experiments of TiO

2

NPs.

Background condition pH Na +

(mmol L -1 )

Ca 2+

(mmol L -1 )

HS

(mg L -1 )

Porosity Flow velocity

(cm min -1 )

0 mg L -1 HS

0.5 mg L -1 HS

1 mg L -1 HS

5 mg L -1 HS

10 mg L -1 HS pH 4.0

4.0

4.0

4.0

4.0

4.0

4.0

0.1

0.1

0.1

0.1

0.1

0.1

0

0.5

1

5

10

0.5

0.45

0.42

0.47

0.46

0.45

0.42

0.36

0.36

0.36

0.38

0.37

0.36 pH 5.0 5.0 pH 6.0

1 mmol L -1 NaCl

6.0

5.0

10 mmol L -1 NaCl

100 mmol L -1 NaCl

5.0

5.0

250 mmol L -1 NaCl 5.0

0.5 mmol L -1 CaCl

2

5.0

1 mmol L -1 CaCl

2

2 mmol L -1 CaCl

2

5 mmol L -1 CaCl

2

5.0

5.0

5.0

250

0.1

0.1

1

10

100

0.5

1

2

5

5

5

5

5

5

0.5

0.5

5

5

5

0.42

0.44

0.42

0.42

0.46

0.45

0.43

0.43

0.45

0.43

0.37

0.37

0.37

0.36

0.36

0.36

0.36

0.36

0.37

0.35

7

33

Adsorption of HS onto TiO

2

NPs

34 To quantitatively evaluate the effects of HS on the stability and transport behavior of

35 TiO

2

NPs observed in column experiments, adsorption studies were conducted to

36 determine the amount of HS adsorbed onto TiO

2

NPs in conditions identical to those

37 used in transport experiments. The suspended particles in TiO

2

NPs suspension

38 prepared above were pelleted by sequential centrifugation (Chen et al. 2012). Briefly,

39 10 mL suspensions were added to Teflon centrifuge tubes and centrifuged for 20

40 minutes at 9,400

g (3K 15, Sigma Laborzentrifugen). Then 8 mL of supernatant

41 were carefully withdrawn from each tube and transferred into another clean centrifuge

42 tube for centrifugation. This procedure was repeated until the TiO

2

NPs was

43 completely removed from the solution. The HS concentration in the supernatant was

44 determined using a spectrophotometer (GENESYS 10S UV-Vis, Thermo Scientific)

45 at 228 nm. The adsorbed HS were then determined by the difference between the

46 initial and final HS concentrations in the aqueous phase. Control experiments with

47 TiO

2

NPs-free solutions showed no variations in HS concentrations before and after

48 the centrifugation processes in the range of HS concentrations tested.

8

0.8

a

0.6

0.4

pH 4.0

pH 5.0

pH 6.0

0.2

0.0

1 10

HS (mg L

-1

)

0.35

b

0.30

0.25

0.20

0.15

49

50

0.10

1 10

NaCl (mmol L

-1

)

100

Fig. S5 Amounts of HS a dsorbedonto the TiO

2

NPs surface as a function of pH and

51 HS concentration (0.1 mmol L

-1

NaCl) (a), and NaCl concentration (5 mg L

-1

HS and

52 pH 5.0) (b) in the TiO

2

suspension

9

3413 2928

HS coated TiO

2

NPs

1625

1389

1042

1118

514

TiO

2

NPs

53

54

55

56

57

4000 3500 3000 2500 2000 1500 1000 500

Wave number (cm

-1

)

Fig. S6 FTIR spectra of TiO

2

NPs and HS coated TiO

2

NPs. Peaks at 2928 cm

-1

for aliphatic CH

2

, 1118 cm

-1 for aliphatic C-OH, and 1042 cm

-1

for C-O-C in FTIR spectra of HS coated TiO

2

NPs could reasonably inferred the adsorption of HS onto the surface of TiO

2

NPs.

10

58

Particles attachment efficiency

59 Using the single collector efficiency model, a dimensionless contact efficiency

η

0

60

61 could be calculated as follows (Tufenkji and Elimelech 2004),

   

0

  

D I G

(S1)

62 Where

η

D

,

η

I

, and

η

G

are collector efficiency components for particles transported to

63 the collector due to diffusion, interception, and gravity, respectively. Diffusion is

64 associated with smaller particles when they undergo Brownian motion due to random

65 bombardment by molecules of the suspending medium and as a result come in contact

66 with the collector surface. Interception occurs when the moving particles contact with

67 the collector grain while traveling along a streamline. Gravitational sedimentation is

68 related to the settling of particles on collector grains by the combined effects of the

69 buoyant weight of the particle and the fluid drag on the particles (Rahman et al. 2013;

70 Yao et al. 1971).

71 Single collector contact efficiency can also be calculated using the following equation

72

73 in detail (Tufenkji and Elimelech 2004):

0

2.4

1/3 A N s R

0.081

N

Pe

0.715

N vdW

0.052

0.55

1.675

N N

R A

0.125

0.22

0.24

1.11

N N N

R G vdW

0.053

(S2)

74 Where A s

is the porosity-dependent parameter of Happel’s model:

75

76

A s

 p 5 2(1 )

2 3 p

3 p 5

2 p 6

(S3)

Here p is determined according to p

  

) 1/3 , where ε is the porosity of a porous

77 medium.

78 N

R

is the aspect radio, which is defined as:

11

79 N

R

 d p d c

(S4)

80 Where d p

and d c

are the diameter of TiO

2

NPs aggregate and quartz packed in column

81 respectively.

82 N

Pe

is Peclet number, which is described as:

83 N

Pe

Ud c

D

(S5)

84 Where U is the approach (superficial) velocity of the TiO

2

NPs suspension, and D

is

85 the diffusion coefficient in an infinite medium, which is calculated using:

86 D

 kT

6

 a p

(S6)

87 Here, k is the Boltzmann constant, 1.3805×10

-23

, T is the temperature in degree of

88 Kelvin, 293 K,

μ is the absolute viscosity of fluid, and a p

is the radius of TiO

2

NPs.

89

90

91

N vdW

is the van der Waals number, which is defined as:

N vdW

A kT

(S7)

Where A is the Hamaker constant for the interacting system of TiO

2

NPs -water-quartz,

92 and was assumed to be 1.0×10 -20

.

93

94

N

G

is the gravity number obtained by the following equation:

N

G

 a p

2 2 (

 p

 

9

U f

) g

(S8)

95 Where

ρ p

and

ρ f

is the density of TiO

2

NPs and fluid respectively, and g is the

96 gravitational acceleration, 9.81 m s

-2

.

97 N

A

is the attraction number, which is calculated using:

98 N

A

12



A

2 a U p

(S9)

12

99 Essentially, η

0

is the probability that a particle will collide with the collector grain

100 through one of the three modes of transport (diffusion, interception, and transport),

101 depending on system hydrodynamics, particle size, density, and van der Waal’s forces

102 (Saleh et al. 2008).

103 Under conditions relevant to most aquatic systems, the single collector removal

104 efficiency η is lower than the single collector contact efficiency η

0

due to repulsive

105 colloidal interactions between particles and collector grains (Tufenkji and Elimelech

106 2004). The actual single collector removal efficiency is often expressed using eq. S10,

107

  

2 d

 c

3(1

) L ln( C C i

/

0

) (S10)

108 Here, d c is the average diameter of the collector particles,

ε is packed bed porosity, L

109 is the length of the column, and C

0 and C i are influent and effluent particle

110 concentration, respectively.

111

The attachment efficiency or “sticking coefficient” α

in eq. S11 represents the fraction

112 of collisions between particles and collectors that result in attachment, i.e., describes

113 the ratio between the experimental single collector removal efficiency

η

and the

114

115 predicted single collector contact efficiency

η

0

.

 

0

(S11)

116

13

117

Derjaguin–Landau–Verwey–Overbeek (DLVO) theory

118 DLVO theory was applied to evaluate the role of electrostatic and van der Waals

119 interactions on the interaction between the nanoparticles and the nanoparticle-quartz

120 surfaces.

121

Total

  vdW

  dl

(S12)

122 DLVO interaction energies between TiO

2

NPs were calculated assuming

123 sphere-sphere geometry by utilizing the following equations (Gregory 1981):

124

125

 vdW

 dl

( )

2

  a r p p

2

A a

101 p

(S13)

    h )

(S14)

126 Interaction profiles for nanoparticles and quartz sand particles were developed

127 assuming sphere-plate geometry and the following equations were used for

128 calculation (Gregory 1981):

129

130

 dl

 vdW

 

A a

102 p

(S15)

  

0 a r p

2

  p c ln

   h

   h )

)

2

 p

 c

2

     h )

(S16)

131 In DLVO interaction energy profiles, positive interaction energy values represent

132 repulsive condition whereas negative interaction energy values correspond to

133 attraction.

134 When DLVO interaction energy between TiO

2

NPs is calculated, a p

is the radius of

135 the initial TiO

2

NPs, 30 nm, and in the case of the energy between TiO

2

NPs and

136 quartz surface, the radius of an equivalent sphere for the nanoparticle aggregates

137 which were measured by DLS has been used as the nanoparticle radius ( a p

). h denotes

14

138 the (minimum) surface-to-surface separation distance between the spheres (for

139 sphere–sphere geometry) or between a sphere and a plate (for sphere–plate geometry).

140 A characteristic wavelength (

λ

) of 100 nm was assumed in the calculations.

141 Permittivity of free space (

ε

0

) and dielectric constant (

ε r

) of water are 8.854×10

-12

C

142 V

-1

m

-1

and 81.5 respectively,

κ is the inverse Debye length (m

-1

) which was estimated

143 for each electrolyte solution using eq. S17, and ψ p

and ψ c

are the surface potentials of

144 TiO

2

NPs and quartz collector (V), respectively. For the calculation of interaction

145 profiles, zeta potentials of TiO

2

NPs and quartz were measured under different

146 chemical conditions and these values were used instead of surface potentials. The

147 Hamaker constant for TiO

2

NPs–water–TiO

2

NPs interaction system ( A

101

) used was

148 3.7×10 -20 J (Shih et al. 2012) and for TiO

2

NPs–water–quartz system ( A

102

) 1.0×10 -20 J

149 was used (Chowdhury et al. 2011).

150

  

10 3 2 e N (2 )

A

 

0 r kT

 1/ 2

151 (S17)

152 Where e is the electron charge, 1.60×10

-19

C, N

A

is Avogadro’s constant, 6.02×10

23

153 mol

-1

, and I is the ionic strength of the solution.

15

0

-5

-5

-10

5

0

20

15

10 a

0

10

5 b

0 mg L

-1

0.5 mg L

-1

1 mg L

-1

5 mg L

-1

10 mg L

-1

10 20 30

Separation Distance (nm)

pH 4.0

pH 5.0

pH 6.0

40

154

155

-10

0 10 20 30

Separation Distance (nm)

40

Fig. S7 Calculated DLVO interaction energy between TiO

2

NPs (based on primary

156 size) under varying HS concentrations (pH 4.0 and 0.1 mmol L

-1

NaCl) (a), and under

157 varying pH (5 mg L -1 HS and 0.1 mmol L -1 NaCl) (b)

16

158

159

160

0.6

0.4

0.2

1.0

0.8

0.0

0 1 2 3 4 5

PV

Fig. S8 Breakthrough curve of conservative Br-tracer (0.1 mmol L

-1

, pH 4.0 and 0.1 mmol L

-1

NaCl) in quartz sands.

17

1.0

a

0.8

0.6

0.4

0.2

0.6

0.4

0.2

0.0

0

1.0

b

0.8

1 2

PV

3

0 mg L

-1

1 mg L

-1

5 mg L

-1

10 mg L

-1

4

0 mg L

-1

1 mg L

-1

5 mg L

-1

10 mg L

-1

5

161

162

163

0.0

0 1 2 3 4 5

PV

Fig. S9 Breakthrough curves for TiO

2

NPs under different HS concentrations at pH

5.0 (a) and pH 6.0(b)

18

600

400

200

0

-200

-400 a

0 10 20 30

Separation Distance (nm)

0 mg L

-1

0.5 mg L

-1

1 mg L

-1

5 mg L

-1

10 mg L

-1

40

400 b

pH 4.0

pH 5.0

pH 6.0

200

0

-200

164

165

166

167

0 10 20 30 40

Separation Distance (nm)

Fig. S10 Calculated DLVO interaction energy between TiO

2

NPs (based on aggregated size) and quartz under varying HS concentrations (pH 4.0 and 0.1 mmol

L -1 NaCl) (a), and under varying pH (0.5 mg L -1 HS and 0.1 mmol L -1 NaCl) (b)

19

30

25

20

15

10

pH 5.0

1 mg L

-1

HS

2 mmol L

-1

CaCl

2

168

169

170

171

5

0

100 1000 10000

Hydrodynamic Diameter (nm)

Fig. S11 Representative number-weighted hydrodynamic diameter distribution of the

TiO

2

NPs in the influent samples at 0.5 mg L

-1

HS, 1 mg L

-1

HS and 2 mmol L

-1

CaCl

2

. The solid lines are drawn to provide visual guides

20

172

References

173 Chen GX, Liu XY, Su CM (2012) Distinct effects of humic acid on transport and

174

175 retention of TiO

2

rutile nanoparticles in saturated sand columns. Environ Sci

Technol 46:7142-7150 doi:Doi 10.1021/Es204010g

176 Chowdhury I, Hong Y, Honda RJ, Walker SL (2011) Mechanisms of TiO

2

nanoparticle

177 transport in porous media: Role of solution chemistry, nanoparticle

178 concentration, and flowrate. J Colloid Interf Sci 360:548-555 doi:DOI

10.1016/j.jcis.2011.04.111 179

180 Gregory J (1981) Approximate expressions for retarded Vanderwaals interaction. J

181 Colloid Interf Sci 83:138-145 doi:Doi 10.1016/0021-9797(81)90018-7

182 Rahman T, George J, Shipley HJ (2013) Transport of aluminum oxide nanoparticles in

183

184 saturated sand: Effects of ionic strength, flow rate, and nanoparticle concentration. Sci Total Environ 463:565-571 doi:DOI

185 10.1016/j.scitotenv.2013.06.049

186 Saleh N, Kim HJ, Phenrat T, Matyjaszewski K, Tilton RD, Lowry GV (2008) Ionic

187

188 strength and composition affect the mobility of surface-modified Fe-0 nanoparticles in water-saturated sand columns. Environ Sci Technol

189

191

42:3349-3355 doi:Doi 10.1021/Es071936b

190 Shih YH, Liu WS, Su YF (2012) Aggregation of stabilized TiO

2

nanoparticle suspensions in the presence of inorganic ions. Environ Toxicol Chem

192 31:1693-1698 doi:Doi 10.1002/Etc.1898

193 Tufenkji N, Elimelech M (2004) Correlation equation for predicting single-collector

21

194 efficiency in physicochemical filtration in saturated porous media. Environ Sci

Technol 38:529-536 doi:Doi 10.1021/Es034049r 195

196 Yao KM, Habibian MM, Omelia CR (1971) Water and waste water filtration:

197 Concepts and applications. Environ Sci Technol 5:1105-1113 doi:Doi

10.1021/Es60058a005 198

199

22

Download