Competitive Upconversion-Linked Immunosorbent Assay for the

May 11, 2016 - Photon-upconverting nanoparticles (UCNPs) emit light of shorter wavelength under near-infrared excitation and thus avoid optical backgr...
4 downloads 15 Views 668KB Size
Subscriber access provided by ORTA DOGU TEKNIK UNIVERSITESI KUTUPHANESI

Article

A competitive Upconversion-Linked Immunosorbent Assay (ULISA) for the rapid and sensitive detection of diclofenac Antonín Hlavá#ek, Zden#k Farka, Maria Hübner, Veronika Hor#áková, Daniel Nemecek, Reinhard Niessner, Petr Skládal, Dietmar Knopp, and Hans H Gorris Anal. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.analchem.6b01083 • Publication Date (Web): 11 May 2016 Downloaded from http://pubs.acs.org on May 12, 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.

Analytical Chemistry 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 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

1

A competitive Upconversion-Linked Immunosorbent Assay

2

(ULISA) for the rapid and sensitive detection of diclofenac

3 4

Antonín Hlaváček†,‡,§, Zdeněk Farka†,‡, Maria Hübnerǁ, Veronika Horňáková‡, Daniel

5

Němeček‡, Reinhard Niessnerǁ, Petr Skládal‡, Dietmar Knoppǁ, Hans H. Gorris†*

6 7 8



Institute of Analytical Chemistry, Chemo- and Biosensors, University of Regensburg, 93040 Regensburg, Germany ‡

9 10

CEITEC - Central European Institute of Technology, Masaryk University, Brno 625 00, Czech Republic

11

§

12 13

ǁ

Institute of Analytical Chemistry AS CR, v. v. i., Brno 602 00, Czech Republic

Chair for Analytical Chemistry and Institute of Hydrochemistry, Technical University of Munich, 81377 Munich, Germany

14 15 16 17 18 19 20 21 22

*Corresponding author

23

Hans H. Gorris, PhD

24 25 26 27

Institute of Analytical Chemistry, Chemo- and Biosensors University of Regensburg Universitätsstr. 31 93040 Regensburg

28

Germany

29

Tel.: +49-941-943-4015

30

Fax: +49-941-943-4064

31

e-mail: [email protected]

1 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

32

Abstract

33

Photon-upconverting nanoparticles (UCNPs) emit light of shorter wavelength under near-

34

infrared excitation and thus avoid optical background interference. We have exploited this

35

unique photophysical feature to establish a sensitive competitive immunoassay for the

36

detection of the pharmaceutical micropollutant diclofenac (DCF) in water. The so-called

37

upconversion-linked immunosorbent assay (ULISA) was critically dependent on the design of

38

the upconversion luminescent detection label. Silica-coated UCNPs (50 nm in diameter)

39

exposing carboxyl groups on the surface were conjugated to a secondary anti-IgG antibody.

40

We investigated the structure and monodispersity of the nanoconjugates in detail. Using a

41

highly affine anti-DCF primary antibody, the optimized ULISA reached a detection limit of

42

0.05 ng DCF per mL. This performance comes close to a conventional ELISA without the

43

need for an enzyme-mediated signal amplification step. The ULISA was further employed for

44

analyzing drinking and surface water samples and the results were consistent with a

45

conventional ELISA as well as LC-MS.

46 47

Introduction

48

The enzyme-linked immunosorbent assay (ELISA) is a cost efficient tool for the specific and

49

highly sensitive detection of many toxic analytes in food and environmental samples as well

50

as clinical diagnosis. There are, however, some disadvantages of a classic ELISA such as an

51

inherent instability of enzymes and time consuming signal development. Consequently, many

52

research efforts have been made to replace the enzymes by using nanoparticles (NPs) as

53

signal amplifiers, e.g. fluorescent dye-doped polymer or silica NPs,1 metal NPs,2 magnetic

54

NPs,3

55

(UCNPs) have been used as a new generation of luminescent labels for sensitive

56

immunochemical detection. UCNPs are lanthanide-doped nanocrystals that can be excited by

57

near-infrared light and emit light of shorter wavelengths (anti-Stokes emission),6,7 which

58

strongly reduces autofluorescence and light scattering. Further advantages of UCNPs include

59

(A) a very high photostability, (B) large anti-Stokes shifts allowing for an excellent separation

60

of excitation and detection channels, and (C) multiple and narrow emission bands that can be

61

tuned individually for the multiplexed detection of analytes.8-10

catalytic NPs4 or quantum dots.5 Recently, photon-upconverting nanoparticles

62

These distinct photophysical features of UCNPs have been used for the design of

63

heterogeneous microtiter plate immunoassays, e.g. for the detection of prostate-specific 2 ACS Paragon Plus Environment

Page 2 of 16

Page 3 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

64

antigen (limit of detection, LOD: 0.15 ng mL-1 / 6 pM)11 or human chorionic gonadotropin

65

(LOD: 3.8 ng mL-1 / 200 pM).12 The advantages of UCNPs in lateral flow assays e.g. for the

66

detection of worm parasite antigens (LOD: 0.01 ng mL-1 / 0.1 pM)13 have also been well

67

documented. There have been a few reports on the use of homogeneous competitive

68

immunoassays for the detection of small molecules such as estradiol (LOD: ~0.1 ng mL-1 /

69

400 pM)14 and folate (LOD: 0.4 ng mL-1 / 1000 pM)15 in blood and a bead-based

70

immunoassay for the detection of mycotoxins in food samples (LOD: ~0.01 ng mL-1 / 50

71

pM).16

72

The development and widespread availability of more sensitive analytical techniques has

73

resulted in an increasing number of pharmaceuticals that can be detected in the environment

74

after medical or veterinary use.17,18 Diclofenac (2-[2-(2,6-dichlorophenyl) aminophenyl]

75

ethanoic acid; DCF) is a widely used non-steroidal anti-inflammatory drug (NSAID). In the

76

Indian subcontinent, the widespread use of DCF for veterinary treatment of cattle since the

77

1990s has led to a precipitous decline of the indigenous vulture population because DCF leads

78

to renal failure in vultures that feed on contaminated carcasses.19 In Europe, DCF belongs to

79

the most frequently detected pharmaceuticals in the water-cycle because it is not easily

80

degraded when passing through sewage treatment plants. DCF has been detected in low µg

81

L−1 amounts in wastewater effluents and also in ng L-1 amounts in surface waters,20

82

groundwater and drinking water.21 Very low amounts of DCF can be detected by LC-TOF-

83

MS or high resolution mass spectrometers.22,23 These instrumental techniques, however, are

84

expensive, time consuming, labor intensive and need trained personnel. By contrast,

85

immunoassays are more suitable for on-site testing directly in the field or for the analysis of

86

large numbers of samples in small laboratories.24

87

Here, we have optimized the preparation of monodisperse and stable upconversion

88

reporters for the sensitive detection of DCF in water samples by a competitive upconversion-

89

linked immunosorbent assay (ULISA). Anti-mouse IgG antibodies were conjugated to silica-

90

coated UCNPs exposing carboxyl function on the surface and the conjugates were

91

characterized by gel electrophoresis.25 The competitive detection of DCF was performed

92

using a monoclonal mouse anti-DCF antibody (Figure 1). This antibody was characterized in

93

detail as described recently26 and showed about 10 % cross-reactivity with DCF metabolites

94

such as 5-OH-DCF, 4’-OH-DCF and DCF-acyl glucuronide, but only less than 1 % with other

95

structurally related non-steroidal anti-inflammatory drugs. The performance of the optimized

96

ULISA was compared with a conventional ELISA as well as LC-MS. 3 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 16

97 98 99 100 101 102 103

Figure 1. Scheme of the indirect competitive ULISA for the detection of diclofenac (DCF). (A) A microtiter plate is coated with a bovine serum albumin-DCF conjugate (BSA-DCF). (B) Dilution series of DCF are prepared in the microtiter plate followed by the addition of anti-diclofenac mouse antibody. (C) The attachment of anti-diclofenac antibody is then detected by an anti-mouse IgG-UCNP secondary antibody conjugate. The upconversion luminescence is recorded under 980 nm laser excitation.

104

Experimental section

105

Chemicals

106

All standard chemicals and diclofenac sodium salt (D6899, purity ≥98 %) were obtained from

107

Sigma-Aldrich (Steinheim, Germany). Carboxyethylsilanetriol sodium salt; 25% (w/v) in

108

water was obtained from ABCR GmbH (Karlsruhe, Germany). The horseradish peroxidase-

109

labeled horse anti-mouse IgG was from Axxora (Lörrach, Germany) and horse anti-mouse

110

IgG was from Vector Laboratories (Burlingame, USA). The monoclonal anti-diclofenac

111

antibody 12G5 was generated in mice using a DCF-thyroglobulin conjugate as described

112

previously.26 An antibody stock solution of 0.45 mg mL-1 was prepared in 20 mM NaH2PO4,

113

20 mM NaH2PO4, 0.1 M Tris-HCl, 0.02% NaN3, pH 7.4 and stored at 4° C. Buffers and

114

solutions were prepared with ultrapure water, which was obtained by reverse osmosis with

115

UV treatment (Milli-RO 5 Plus, Milli-Q185 Plus, Eschborn, Germany).

116

Synthesis of carboxyl-silica-coated UCNPs

117

UCNPs of 42.5 ± 4.9 nm in diameter were synthesized by high-temperature co-precipitation27

118

as described in the Supporting Information. The mass concentration was determined by

119

gravimetric analysis and a concentration of 1.0 mg mL-1 of UCNPs was estimated to be

120

equivalent to the molar concentration of 9.8×10-9 mol L-1 (Supporting Information).

121

Carboxyl-silica-coated

UCNPs

(COOH-UCNPs)

were

prepared

by

a

reverse

122

microemulsion method:25 UCNPs (80 mg) were diluted in cyclohexane to a final volume of

123

23 mL. This dispersion was mixed with 1800 mg of Igepal CO-520 and 100 µL of tetraethyl

124

orthosilicate (TEOS) and stirred intensively for 10 min. A mixture of 55 µL 25 % (w/v)

125

aqueous ammonium hydroxide and 55 µL water was added to form a microemulsion that was 4 ACS Paragon Plus Environment

Page 5 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

126

slowly stirred overnight. Then, 25 µL of TEOS were added and the microemulsion was again

127

stirred for 180 min. After adding 50 µL of 25% (w/v) sodium carboxyethylsilanetriol in water,

128

the microemulsion was first sonicated for 15 min and then stirred for 60 min. The COOH-

129

UCNPs were extracted with 1000 µL of dimethylformamide and washed four times with 20

130

mL of propan-2-ol, three times with 5000 µL of water and finally dispersed in water to yield a

131

final concentration of 150 mg mL-1. The COOH-UCNPs in water were stable at 4°C for

132

several months.

133

Conjugation of COOH-UCNP and secondary antibody

134

COOH-UCNPs were first activated by 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide

135

(EDC) and N-hydroxysulfosuccinimide (sulfo-NHS). In a typical synthesis, 0.5 mg (~5 pmol)

136

of COOH-UCNPs was dispersed in water to a final volume of 200 µL. The volume of 50 µL

137

of a mixture containing 2.1 µmol of EDC and 5.5 µmol of sulfo-NHS in 100 µL of 100 mM

138

sodium 2-(N-morpholino)ethanesulfonate (MES) buffer, pH 6.1 was added and mixed for 30

139

min. A dispersion of 100 µL activated COOH-UCNPs (1 mg mL-1 or ~10 nmol L-1) were

140

mixed with 100 µL of horse anti-mouse IgG in borate buffer (100 mM sodium borate, pH

141

9.0). Three IgG concentrations were employed and incubated for 90 min at room temperature:

142

1) 330 nmol L-1 IgG resulting in a ratio 33 IgG molecules per UCNP (sample IgG-UCNP-

143

33:1). 2) 67 nmol L-1 IgG resulting in a ratio of 7 IgG molecules per UCNP (IgG-UCNP-7:1),

144

and 3) 33 nmol L-1 IgG resulting in a ratio of 3 IgG molecules per UCNP (IgG-UCNP-3:1).

145

The bioconjugates were centrifuged for 10 min at 4,000 g, dispersed in UCNP assay buffer

146

(50 mM Tris, 150 mM NaCl, 0.05% NaN3, 0.01% Tween 20, 0.05% bovine γ-globulin

147

(BGG), 0.5% bovine serum albumin (BSA), 0.2% polyvinyl alcohol 6000 (PVA), pH 7.75)

148

and sonicated for 5 min.

149

Nanoparticle characterization

150

Transmission electron microscopy (TEM) was performed on a Tecnai F20 FEI instrument

151

(Eindhoven, The Netherlands). About 4 µL of UCNPs were deposited on a 400-mesh copper

152

EM grid coated with a continuous carbon layer and negatively stained with 2% (w/v) aqueous

153

solution of uranyl acetate to increase the contrast of the silica shell. The dried grids were then

154

imaged at 50,000× magnification (2.21 Å pixel-1). The size of individual particles in the TEM

155

images was measured by the imaging software ImageJ (http://imagej.nih.gov).28 The

156

hydrodynamic diameter and zeta potential of UCNP suspensions were determined on a

157

Zetasizer Nano SZ from Malvern Instruments (Malvern, UK). FT-IR spectra were recorded

158

on an Alpha FTIR Spectrometer from Bruker (Billerica, USA). 5 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

159

Agarose gel electrophoresis

160

Following our previous work,25 COOH-UCNPs and their bioconjugates were characterized by

161

agarose gel electrophoresis (0.5% w/v agarose, 45 mM Tris, 45 mM H3BO3 with pH 8.6, 15

162

min at 100 V). Samples were mixed in a ratio of 10:1 with 50 % w/w glycerol and 8 µL

163

aliquots were loaded onto the gel. A custom-built upconversion reader (CHAMELEON

164

multilabel microplate reader, Hidex, Turku, Finland) equipped with a continuous 980 nm laser

165

(4 W) was used to scan agarose gels with a spatial resolution of 0.5 mm as described earlier.29

166

Conjugation of diclofenac to BSA

167

BSA-DCF conjugates were prepared using either 1.5 µmol or 7.5 µmol DCF and 9.7 µmol

168

sulfo-NHS added to a mixture of 400 µL MES buffer and 100 µL dimethylformamide. DCF

169

was activated by the addition of 47 µmol of EDC and incubation at room temperature for 30

170

min. After adding 500 µL of 0.15 µmol BSA in water and 250 µL of 50 mM aqueous Na2CO3,

171

the mixture was incubated at room temperature for 4 hours and then dialyzed three times

172

against 150 mL of 50 mM Na2CO3. BSA-DCF was adjusted to a concentration of 1.6 mg

173

mL−1 by adding 50 mM of Na2CO3 and stored at 4 °C in presence of 0.05 % NaN3. The

174

conjugate was analyzed by MALDI-TOF-MS (Bruker, Ultraflex TOF/TOF, N2-laser, 337 nm,

175

positive mode).

176

Water samples

177

Munich tap water and two surface water samples were collected in Southern Bavaria from

178

lake Wörthsee and river Würm. The fresh water samples were filtrated over a glass microfiber

179

filter (GF/C, Whatman Cat. No. 1822 047) and stored at 4°C. The concentrations of Ca2+,

180

Mg2+, and dissolved organic content (DOC) as well as the conductivity and pH were

181

determined (Supporting Information Table 1). ELISA and ULISA were performed with

182

undiluted and spiked samples. For LC-MS, the samples were subjected to generic solid phase

183

extraction (SPE) and analyzed by an Orbitrap-based ExactiveTM Benchtop Mass Spectrometer

184

(Thermo Scientific, Dreieich, Germany) as described earlier26.

185

Upconversion-linked immunosorbent assay (ULISA)

186

A transparent 96-well microtiter plate with high protein binding capacity (Corning,

187

Wiesbaden, Germany) was coated with BSA-DCF in coating buffer (optimal concentration: 1

188

µg mL-1 BSA-DCF in 50 mM NaHCO3 /Na2CO3, 0.05% NaN3, pH 9.6; 200 µL per well) at

189

4° C over night. All subsequent steps were carried out at room temperature. The plate was

190

washed manually four times with 250 µL of washing buffer (50 mM Na/H2PO4/HPO4, 0.01% 6 ACS Paragon Plus Environment

Page 6 of 16

Page 7 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

191

Tween 20, 0.05% NaN3, pH 7.4). The free binding sites in each well were blocked with 250

192

µL of 1% BSA in 50 mM Na/H2PO4/HPO4, 0.05 % NaN3, pH 7.4 for 1 h. The plate was

193

washed four times with washing buffer. Either standard dilutions of DCF in double distilled

194

water or environmental samples (100 µL per well) were added, immediately followed by the

195

anti-DCF monoclonal mouse antibody (12G5, optimal concentration: 0.225 µg mL−1 in 100

196

mM NaPO4, 300 mM NaCl, 100 µL per well) and incubated for 1 h. After four washing steps,

197

the microtiter plate was incubated for one hour with 100 µL of the IgG-UCNP conjugate

198

(optimal concentration: 10 µg mL−1 in 50 mM Tris, 150 mM NaCl, 0.05% NaN3, 0.01%

199

Tween 20, 0.05% BGG, 0.5% BSA, 0.2% PVA, pH 7.75). After four washing steps, the

200

upconversion luminescence was read out from empty wells using a custom-built upconversion

201

microplate reader (CHAMELEON multilabel microplate reader, Hidex, Turku, Finland)

202

equipped with a continuous 980 nm laser (4 W). A collimated laser spot of ~0.8 mm was

203

focused on the bottom of the microtiter wells. Each well was scanned 100 times in a raster

204

with the step size of 0.4 mm and 500 ms signal integration time. The truncated mean was

205

calculated for each well after discarding the ten highest and ten lowest measurements of the

206

luminescence intensity to account for local irregularities on the microtiter well surface that

207

result in signal outliers.

208

Enzyme-linked immunosorbent assay (ELISA)

209

The ELISA was performed as described earlier.26 A transparent 96-well microtiter plate with

210

high protein binding capacity (Greiner Bio-one, Frickenhausen, Germany) was coated with

211

0.5 µg mL−1 of ovalbumin-DCF conjugate in coating buffer (50 mM NaHCO3/Na2CO3, 0.05%

212

NaN3, pH 9.6; 200 µL per well) at 4 °C over night. All subsequent steps were carried out at

213

room temperature. The plate was automatically washed with a plate washer (ELx405 Select,

214

Bio-Tek Instruments, Bad Friedrichshall, Germany) four times with washing buffer (50 mM

215

K/H2PO4/HPO4, 146 mM NaCl, 0.05% Tween 20, pH 7.6; PBST). The free binding sites in

216

each well were blocked with 300 µL of 1% BSA in PBST for 1 h. The plate was washed four

217

times with washing buffer. First, standard dilutions of DCF in double distilled water or

218

environmental samples (100 µL per well) were added, immediately followed by the anti-DCF

219

monoclonal mouse antibody (12G5, 0.5 µg mL−1 in PBS; 100 µL per well) and incubated for

220

30 min. After four washing steps, the secondary horseradish peroxidase-labeled antibody was

221

added (0.2 mg mL-1 in PBS; 200 µL per well) and incubated for 1 h. After final washing, the

222

substrate solution (200 µL per well) was added and the plates were shaken for about 15 min

223

for color development. The substrate solution consisted of 25 mL substrate buffer (prepared 7 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 16

224

with 46.0 mL potassium dihydrogen citrate and 0.1 g potassium sorbate in 1 L of water, pH

225

3.8), 500 µL 3,3',5,5'-tetramethylbenzidine stock solution (375 mg in 30 mL of

226

dimethylsulfoxide) and 100 µL 1% hydrogen peroxide. The enzyme reaction was stopped by

227

adding 100 µL of 5% sulfuric acid per well. The absorbance was read at 450 nm by a

228

microplate reader (Synergy HT, Bio-Tek Instruments).

229

Data analysis

230

A four-parameter logistic function (Eq. 1) was used for a regression analysis of the calibration

231

curves:

232

Y =

233

where [DCF] is the concentration of diclofenac, and Y is either the upconversion

234

luminescence or the absorbance at 450 nm. Eq. 1 yields the maximum (Ymax) and background

235

(Ybg) signal, the DCF concentration that reduces (Ymax-Ybg) by 50 % (IC50) and the slope at the

236

inflection point (s). All measurements were made at least in triplicate. The concentration of

237

DCF in real water samples was determined by utilizing an inverse function of Eq. 1 and the

238

limit of detection (LOD) was defined as before:26

239

Y(LOD) = 0.85 × (Ymax - Ybg) + Ybg

Ymax − Ybg  [ DCF ]   1 +   IC 50 

s

+ Ybg

(1)

(2)

240 241

Results and Discussion

242

Surface modification and characterization of UCNPs

243

The development of a competitive upconversion immunoassay (ULISA) for the detection of

244

DCF (Figure 1) critically depends on the design of the luminescent reporter that replaces the

245

conventional enzyme amplification steps.30 Oleic acid-capped UCNPs were coated with a

246

silica shell exposing carboxylic acid functional groups on the surface. The carboxyl groups

247

improve the dispersibility in water and serve as attachment sites for subsequent conjugation

248

steps. We previously described a one-step water-in-oil microemulsion protocol for coating the

249

surface of small UCNPs (~12 nm in diameter) with a carboxylated silica shell that showed

250

only a weak upconversion luminescence.25 For the immunoassay, we have synthesized larger

251

UCNPs of 42.5 ± 4.9 nm in diameter (Supporting Information Figure S1) that are much

252

brighter because they are less affected by surface quenching effects.31 The one step silica8 ACS Paragon Plus Environment

Page 9 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

253

coating protocol, however, resulted in aggregation when directly applied to bigger

254

nanoparticles. Therefore, we developed a two-step protocol to prepare a thicker, compact and

255

more stable silica shell on the surface of UCNPs.32 First, TEOS was added to the

256

microemulsion to generate a thin layer of bare silica (2.4 ± 0.4 nm, Supporting Information

257

Figure S2). This step alone was not sufficient to prevent aggregation. Therefore, TEOS was

258

added for a second time, which changed the thickness of the silica shell only slightly. The

259

second carboxylation step ensured an excellent dispersibility of COOH-UCNP in water.33 The

260

total diameter of COOH-UCNPs was consistent as determined by transmission electron

261

microscopy (TEM: 46.9 ± 5.0 nm; Supporting Information Figure S3 and Figure 2A) and

262

atomic force microscopy (AFM: 45.4 ± 7.6 nm; Supporting Information Figure S4). Dynamic

263

light scattering measurements confirmed an increase of the hydrodynamic diameter from 55

264

nm to 65 nm after silica coating (Supporting Information Figure S5).

265

The optimized COOH-UCNPs were then conjugated to a secondary anti-IgG antibody via

266

standard EDC/sulfo-NHS chemistry (Figure 2B).34,35 A low concentration of COOH-UCNP

267

was utilized to prevent that one antibody molecule binds to several UCNPs, which would lead

268

to aggregation. The conjugates were characterized by agarose gel electrophoresis (Figure 2C),

269

dynamic light scattering (DLS), zeta potential measurements and FT-IR spectroscopy

270

(Supporting Information Figures S6-S8). The lowest degree of aggregation was observed

271

when the concentration of COOH-UCNPs in the reaction mixture was 1 mg mL-1. The

272

conjugation of the secondary antibody reduced the negative surface potential of the COOH-

273

UCNPs as shown by zeta potential measurements and led to a stronger retardation in the

274

agarose gel.36 The shift of the electrophoretic mobility was linearly dependent on the ratio of

275

IgG molecules per UCNP and indicated the degree of surface modification (Figure 2D).

276

Additionally, larger aggregates of nanoparticles remained in the gel pockets and could not

277

enter the agarose matrix. Sample IgG-UCNP-33:1 shows a main fraction of monodisperse

278

bioconjugates separated as a distinct band and a smaller fraction of slowly moving

279

components, which are probably partially aggregated and crosslinked bioconjugates. This

280

result is consistent with a bimodal particle distribution and a higher polydispersity index

281

observed in the DLS measurement.

282

The IgG-UCNP conjugates were purified from an excess of unbound secondary anti-mouse

283

IgG and components of the reaction mixture by differential centrifugation. At first, the

284

bioconjugates were centrifuged at 10,000 g, which, however, led to strong nanoparticle

285

aggregation (Figure 2C). When the centrifugal speed was reduced to 4,000 g followed by 9 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

286

short sonication, purified and monodisperse IgG-UCNP were obtained. Further lowering of

287

the centrifugal field was not efficient since the sedimentation of IgG-UCNP was too slow.

288

The small retardation coefficient of UCNPs prepared with the lowest amount of IgG (IgG-

289

UCNP-3:1, Figure 2C, lane V) indicated an insufficient surface modification. Therefore, this

290

bioconjugate was not used for the following ULISA experiments.

291

292 293 294 295 296 297 298 299 300 301

Figure 2. Preparation and characterization of IgG-UCNP conjugates. (A) TEM image of silica-coated UCNPs exposing carboxyl groups on the surface (COOH-UCNPs, Supporting Information Figure S3). (B) The carboxyl groups are activated by EDC/sulfo-NHS and conjugated to anti-mouse IgG. (C) The conjugates are prepared by using different ratios of anti-mouse IgG and COOH-UCNPs (I/II: 33 to 1; III/IV: 7 to 1; V/VI: 3 to 1, VII/VIII: no IgG). Each sample is centrifuged either with 4,000 g (I, III, V, VII) or with 10,000 g (II, IV, VI, VIII) and characterized by agarose gel electrophoresis. The migration distance (∆) is indicated with red lines. (D) The relative electrophoretic mobility (∆ratio[x]/∆no IgG) of the conjugates is linearly dependent on the ratio of IgG molecules per UCNP.

302

Design of Upconversion-linked immunosorbent assay (ULISA)

303

In a competitive immunoassay, a low concentration of coating antigen ensures that the free

304

analyte can compete efficiently for the binding sites of the detection antibodies. On the other

305

hand, the signal generation has to be strong enough for a reliable readout. Here, we prepared

306

two different coating conjugates consisting of BSA-DCF. The conjugates were analyzed by

307

MALDI-TOF mass spectrometry, which showed a coupling density of either 5.7 or 10 DCF

308

residues per BSA molecule (Supporting Information Figures S9-S11). When the conjugate 10 ACS Paragon Plus Environment

Page 10 of 16

Page 11 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

309

with the higher degree of derivatization was used for coating in the immunoassay, the signals

310

were – as expected – about twice as high but also showed stronger signal fluctuation and a

311

hook effect (Figure 3A), which may be the consequence of two binding sites of IgG

312

molecules forming cyclic complexes (Supporting Information Figure S12).37 By contrast, the

313

conjugate exposing 5.7 DCF residues per BSA molecule yielded more stable signals and a

314

slightly lower IC50 value (1.2 ng mL-1 compared to 1.5 ng mL-1) and a lower detection limit

315

for DCF. Consequently, this coating conjugate was used in all further experiments. An

316

optimal signal generation was observed with a coating concentration of 1 µg mL-1

317

(Supporting Information Figure S13).

318 319 320 321 322 323 324 325

Figure 3. ULISA optimization. (A) Microtiter plates are coated with 1 µg mL-1 of BSA carrying either 10 (□) or 5.7 (○) DCF residues. (B) The upconversion luminescent (UCL) signal is generated by using 10 µg mL-1 of IgGUCNP-33:1 (○) or IgG-UCNP-7:1 (□), respectively. (C) The detection of DCF is optimized by using the monoclonal anti-DCF antibody in concentrations of 0.5 µg mL-1 (○) (IC50: 0.68 ng mL-1), 0.25 µg mL-1 (□) (IC50: 0.23 ng mL-1), 0.1 µg mL-1 (∆) (IC50: 0.13 ng mL-1) or 0.02 (◊) µg mL-1 (IC50: 0.08 ng mL-1). Error bars represent standard deviations in upconversion signals from three replicate wells.

326

The competition step including free DCF and anti-DCF detection antibody was performed

327

in analogy to a sensitive conventional ELISA.26 Only the enzyme-mediated color generation

328

was replaced by an IgG-UCNP conjugate as a direct luminescent reporter (Figure 3B). The

329

higher degree of UCNP surface coverage (IgG-UCNP ratio of 33:1 compared to 7:1)

330

increased the maximum signal intensity by a factor of five although both conjugates were

331

prepared with a molar excess of IgG molecules per nanoparticle. This difference can be

332

explained because not every surface-conjugated antibody may have the right orientation or be

333

fully functional in order to bind efficiently to the primary antibody. Consequently, a higher

334

degree of derivatization resulted in a proportionally higher number of functional antibodies.

335

On the downside, using IgG-UCNP-33:1 resulted in strong signal fluctuations as well as a

336

hook effect, which impedes the reproducible determination of DCF. It should also be noted

337

that the degree of surface substitution did not significantly affect IC50 or the LOD, and a

11 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

338

concentration of 10 µg mL-1 IgG-UCNP-7:1 resulted in the most reproducible upconversion

339

signal generation (Supporting Information Figure S14)

340

In contrast to the UCNP-bound secondary antibody, the primary anti-DCF antibody is

341

directly involved in the competition step. Figure 3C shows that both the upconversion signal

342

intensity and the IC50/LOD for DCF strongly depend on the antibody concentration. A higher

343

primary antibody concentration leads to a higher signal intensity because more antibodies can

344

bind to the DCF-BSA coating conjugate, but they also consume a larger amount of free DCF

345

and thus deteriorate the assay sensitivity. A concentration of 0.25 µg mL-1 primary anti-DCF

346

antibody yielded an optimal balance between signal generation and sensitivity for the

347

determination of DCF and was used in all further experiments.

348

Calibration and sensitivity of ULISA and ELISA

349

For each type of competitive immunoassay it is necessary to find the optimal balance between

350

detection sensitivity for an analyte and signal development. It should also be noted that a high

351

affinity and a low cross-reactivity of the primary antibody are the most distinctive features

352

that determine the sensitivity and specificity of the analyte detection. Figure 4 shows

353

calibration curves of ULISA and ELISA recorded under similar conditions and using the same

354

anti-DCF primary antibody. In both cases a signal to background ratio (Ymax/Ybg) of 5:1 was

355

adjusted to achieve the most sensitive detection of DCF but also to obtain a reliable signal

356

generation. The competitive ULISA (LOD: 0.05 ng mL−1 / 170 pM) has a five times higher

357

detection limit than a conventional ELISA (LOD: 0.01 ng mL−1/ 34 pM) but allows for an

358

easier and faster signal generation. As the detection sensitivity is ultimately dependent on the

359

anti-DCF antibody, it can be expected that the ULISA can be further optimized by developing

360

brighter UCNPs reporter conjugates.

361

Competitive immunoassays for small molecules are typically less sensitive than sandwich

362

immunoassays where the signal generation is directly proportional to the analyte

363

concentration. The highest sensitivity was described for the detection of Schistosoma

364

circulating anodic antigen by using micron-sized upconversion particles in a lateral flow assay

365

(LOD: 0.01 ng mL-1 / 0.1 pM).13 This particular analyte displays repetitive surface epitopes

366

and facilitates binding of several primary antibodies per analyte molecule. The competitive

367

immunoassay for DCF affords a similar sensitivity as a magnetic bead-based competitive

368

immunoassay for the detection of aflatoxin that was reported to reach an LOD of 0.01 ng mL-1

369

(50 pM) under optimal conditions.16 The additional magnetic separation step, however,

370

demands a more sophisticated instrumentation and is more time consuming. 12 ACS Paragon Plus Environment

Page 12 of 16

Page 13 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Analytical Chemistry

371 372 373 374 375

Figure 4. Normalized calibration curves of ULISA (□, replotted red curve from Figure 3C; IC50: 0.23 ng mL−1, LOD: 0.05 ng mL−1) and ELISA (○, IC50: 0.05 ng mL−1, LOD: 0.01 ng mL−1). Error bars represent standard deviations of three replicate wells.

376

Detection of diclofenac in real water samples

377

Two surface water samples and drinking water were collected in Southern Bavaria and the

378

matrix was analyzed (Supporting Information Table S1) to assess possible interferences with

379

the detection of DCF. These interferences should be as low as possible because matrix effects

380

can suppress the immunoassay signal and lead to an overestimation of analyte concentrations.

381

The monoclonal primary antibody 12G5 is resistant to matrix interferences over a wide pH

382

range, humic acid concentrations up 20 mg L-1 and Ca2+ concentrations up to 75 mg L-1 as

383

described earlier.26 The drinking water sample from Munich, however, contained relatively

384

high Ca2+ and Mg2+ concentrations of 110 mg L-1 in total, which is probably the reason for a

385

signal suppression (defined as 100 × (Ysample - Ybg) / (Ymax - Ybg)) to 60±7 % in the ULISA and

386

73±10 % in the ELISA in the undiluted samples without DCF. By contrast, the surface water

387

samples contained less Ca2+ and Mg2+ and were less affected by signal suppression.

388

The concentration of DCF was too low to be detectable in the unspiked water samples.

389

Thus, each sample was additionally spiked with either 1 ng mL-1 or 10 ng mL-1 of DCF. The

390

spiked samples were typically diluted at least by a factor of three prior to the immunoassay to

391

keep matrix effects to a minimum. Table 1 shows the concentrations of DCF as determined by

392

ULISA, ELISA and LC-MS. The ULISA led to slightly stronger deviations from the spiking

393

concentration compared to the ELISA because the matrix may also have an impact on the

394

binding of the nanoparticulate luminescent reporter unit, which is relatively large compared to

395

the enzyme antibody conjugate used for the ELISA. These differences in the immunoassay

396

performance are subject to further investigation and will be optimized to unfold the full

397

potential of the ULISA for the background-free detection of analytes.

398 13 ACS Paragon Plus Environment

Analytical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

399

Page 14 of 16

Table 1. Detection of DCF in unspiked and spiked real water samples. Sample

Lake Wörthsee

River Würm

Munich tap water

Spiked (ng mL-1)

ULISA (ng mL-1)

ELISA (ng mL-1)

LC-MS (ng mL-1)

-