34
Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung, 1 Pei-Ying Lin, 1 Chiung-Wen Hsieh, 1 and Shuchen Hsieh 1, 2 () Nano Res., Just Accepted Manuscript • DOI: 10.1007/s12274-014-0479-y http://www.thenanoresearch.com on April 21, 2014 © Tsinghua University Press 2014 Just Accepted This is a “Just Accepted” manuscript, which has been examined by the peer-review process and has been accepted for publication. A “Just Accepted” manuscript is published online shortly after its acceptance, which is prior to technical editing and formatting and author proofing. Tsinghua University Press (TUP) provides “Just Accepted” as an optional and free service which allows authors to make their results available to the research community as soon as possible after acceptance. After a manuscript has been technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Please note that technical editing may introduce minor changes to the manuscript text and/or graphics which may affect the content, and all legal disclaimers that apply to the journal pertain. In no event shall TUP be held responsible for errors or consequences arising from the use of any information contained in these “Just Accepted” manuscripts. To cite this manuscript please use its Digital Object Identifier (DOI® ), which is identical for all formats of publication. Nano Research DOI 10.1007/s12274-014-0479-y

Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

Nano Res

1

Aqueous Self-Assembly and Surface-Functionalized

Nanodots for Live Cell Imaging and Labeling

Mei-Lang Kung,1 Pei-Ying Lin,

1 Chiung-Wen Hsieh,

1 and Shuchen Hsieh

1, 2 ()

Nano Res., Just Accepted Manuscript • DOI: 10.1007/s12274-014-0479-y

http://www.thenanoresearch.com on April 21, 2014

© Tsinghua University Press 2014

Just Accepted

This is a “Just Accepted” manuscript, which has been examined by the peer-review process and has been

accepted for publication. A “Just Accepted” manuscript is published online shortly after its acceptance,

which is prior to technical editing and formatting and author proofing. Tsinghua University Press (TUP)

provides “Just Accepted” as an optional and free service which allows authors to make their results available

to the research community as soon as possible after acceptance. After a manuscript has been technically

edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP

article. Please note that technical editing may introduce minor changes to the manuscript text and/or

graphics which may affect the content, and all legal disclaimers that apply to the journal pertain. In no event

shall TUP be held responsible for errors or consequences arising from the use of any information contained

in these “Just Accepted” manuscripts. To cite this manuscript please use its Digital Object Identifier (DOI® ),

which is identical for all formats of publication.

Nano Research

DOI 10.1007/s12274-014-0479-y

Page 2: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

1

Aqueous Self-Assembly and Surface-Functionalized

Nanodots for Live Cell Imaging and Labeling

Mei-Lang Kung, Pei-Ying Lin, Chiung-Wen Hsieh, and

Shuchen Hsieh*

“National Sun Yat-sen University”, Kaohsiung

Schematic illustration of a conjugated self-assembled nanodot (SAND)

system for targeted cell imaging

Page 3: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

2

Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling

Mei-Lang Kung,

1 Pei-Ying Lin,

1 Chiung-Wen Hsieh,

1 and Shuchen Hsieh

1, 2 ()

1Department of Chemistry and Center for Nanoscience and Nanotechnology, “National Sun Yat-sen University”, Kaohsiung, 80424 2School of Pharmacy, College of Pharmacy, Kaohsiung Medical University, Kaohsiung, 80707

Received: day month year / Revised: day month year / Accepted: day month year (automatically inserted by the publisher)

© Tsinghua University Press and Springer-Verlag Berlin Heidelberg 2011

ABSTRACT Nanoparticles have enormous potential for bioimaging and biolabeling applications, in which conventional

organically based fluorescent labels degrade and fail to provide long-term tracking. Thus, the development of

approaches to make fluorescent probes water soluble and label cells efficiently is desirable for most biological

applications. Here, we report on the fabrication and characterization of self-assembled nanodots (SANDs) from

3-aminopropyltriethoxysilane (APTES) as a probe for protein labeling. We show that fluorescent SAND probes

exhibit both bright photoluminescence and biocompatibility in an aqueous environment. Selective in vitro

imaging using protein and carbohydrate labeling of hepatoma cell lines are demonstrated using biocompatible

SANDs conjugated with avidin and galactose, respectively. Cytotoxicity tests show that conjugated SAND

particles have negligible effects on cell proliferation. Unlike other synthetic systems that require multistep

treatments to achieve robust surface functionalization and to develop flexible bioconjugation strategies, our

results demonstrate the versatility of this one-step SAND fabrication method for creating multicolor fluorescent

probes with tailored functionalities, efficient emission, as well as excellent biocompatibility, requisite for broad

biological use.

KEYWORDS Atomic force microscopy, hydrophilicity, molecular imaging, self-assembly, silane

1. Introduction 1

Fluorescence imaging and biolabeling technologies 2

enable visualization of a vast array of biological 3

moieties in their native environment. Biomolecules, 4

proteins, and ions, for example, may be monitored 5

within cells, tissues, and organisms, in order to better 6

understand their function and role in biological 7

processes [1, 2]. 8

Organic dyes, such as FITC and fluorescent 9

proteins, have been used extensively in cell and 10

tissue bioimaging because they are biodegradable, 11

nontoxic, and target-specific. However, they are 12

vulnerable to chemical and metabolic degradation, 13

and rapid photobleaching, which limits their range 14

of use in bio-applications and biomolecule tracking 15

[3]. 16

Nano Res DOI (automatically inserted by the publisher)

Research Article

————————————

Address correspondence to Shuchen Hsieh, [email protected]

Page 4: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

3

Fluorescent nanoparticles such as semiconductor 17

quantum dots (QDs) have been shown to exhibit 18

strong photoluminescence (high quantum yields), 19

multiple fluorescence, and excellent photostability 20

[4]. Thus, QDs have been proposed as promising 21

candidates for molecular labeling as well as in vitro 22

and in vivo imaging. Most importantly for cancer 23

screening, QDs could deliver tens to hundreds of 24

cancer biomarkers simultaneously, making them 25

suitable as a clinical diagnostic tool for the early 26

detection and classification of cancer [5]. 27

However, for most biolabeling and biomedical 28

applications, QDs must undergo secondary chemical 29

processing to modify the outer surface. This serves 30

the dual purpose of providing a mechanism for 31

adding specific chemical functionalities as well as 32

isolating the toxic QD core materials from the 33

environment. For example, it is often desirable to 34

convert hydrophobic characteristics to hydrophilic 35

(water-soluble) [6] so that different functional 36

molecules or biomolecules can be attached. Further, 37

their potential cytotoxicity, caused by the leaching of 38

heavy metal ions, remains a critical concern for in 39

vivo applications [7]. 40

In our previous work, Lin et al. [8] reported on the 41

one-pot self-assembled SiOx nanodots via hydrolysis 42

and condensation reaction using alkylakoxysilanes 43

as molecular precursors. These “self-assembled 44

nanodots (SANDs)” had the benefits of high 45

quantum yield, wide band absorption, multicolor 46

fluorescence, and very low toxicity. This combination 47

of properties were exploited in intravital imaging 48

experiments using zebrafish [8], demonstrating the 49

potential advantages for applications in biological, 50

biomedical, and clinical research. 51

Among the currently available materials from 52

alkylalkoxysilanes, 3-aminopropyltriethoxysilane 53

(APTES) is one an amine containing silane widely 54

used as a coupling agent for the surface modification 55

of silica or mica by self-assembly to enhance the 56

adhesion probability of biomolecules [9]. For 57

instance, Chai et al. [10] have developed an 58

immunosensor, which was fabricated by the 59

cross-linked bond of antibody with the 60

self-assembled monolayer (SAM) of an APTES 61

treated aluminum substrate, for the detection of ricin, 62

a lethal biological toxin. 63

Here, we demonstrate that SANDs from APTES 64

can be formed in aqueous solution and are well 65

suited for biolabeling applications. The APTES 66

derived SAND particles were further amine 67

functionalized, yielding a hydrophilic coating which 68

facilitates optical applications for biological systems. 69

This coating also provides for specific conjugation 70

with biomolecules [11]. Our results demonstrate that 71

the SANDs exhibit low cytotoxicity, are highly 72

photostable, and can be utilized for protein- and 73

carbohydrate-labeling via bioconjugation, producing 74

SAND-avidin and SAND-galactose probes, 75

respectively. Because of their biocompatibility and 76

amine-based labeling in aqueous solution, the 77

functionalized SANDs can be widely used as an 78

imaging probe, biomarker, or drug carrier for the 79

development of nanomedicine. 80

81

2. Experimental 82

2.1 SANDs synthesis 83

3-Aminopropyltriethoxysilane (APTES, >99%) was 84

purchased from (Sigma-Aldrich, USA), and used 85

without further purification. The molar 86

concentration of APTES was 4.3 M and calculated 87

from known densities and molecular weights (0.946 88

g/ml and 221.4 g/mol). To prepare the SAND 89

particles, 2 ml of APTES was placed into a 90

thoroughly cleaned glass vial and heated at a 91

constant temperature of 200°C while stirring (1100 92

rpm) on a hot-plate/stirrer (Suntex Inc. Co., Taiwan, 93

SH-301) for 4 h under an ambient “air” atmosphere 94

and then cooled with continued stirring to room 95

temperature. The SAND liquids were then used 96

directly to prepare sample for analysis by PL, XPS, 97

FTIR, and Raman. Aliquots of these samples were 98

diluted in toluene at a 1:200 V/V ratio to prepare 99

stock solutions for analysis by AFM and TEM. 100

101

2.2 Characterization of SANDs 102

Photoluminescence (PL) spectra measurements 103

were acquired using a HITACHI F-7000 104

Fluorescence Spectrometer. The excitation 105

Page 5: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

4

wavelengths were 280 to 520 nm at 20 nm intervals 106

in the excitation domain. Emission spectra were 107

recorded from 200-800 nm. The UV–Vis absorption 108

spectrum was recorded with a UV–Vis 109

spectrophotometer (JASCO V-630, Japan) using a 110

1-cm path length quartz cell. In atomic force 111

microscope (AFM) imaging, 10 μl of SAND/toluene 112

solution was dropcast on a freshly-cleaved mica 113

substrate. An AFM (MFP-3D, Asylum Research, 114

USA), operating in intermittent contact mode under 115

ambient conditions, was used to visualize the 116

resulting surfaces to characterize the Au NP 117

distributions. A silicon cantilever (Olympus, 118

AC240TS) with a nominal spring constant of 2 Nm-1 119

was used for all images at a scan rate of 1.0 Hz and 120

image resolution of 512×512 pixels. Transmission 121

electron microscopy (TEM) images of SAND 122

samples were acquired using a JEOL JEM-2100 123

operated at 200 KV and at a vacuum of 2×10-5 Pa. 124

The TEM sample was prepared by depositing 0.5 μl 125

of SAND solution onto a TEM carbon grid with 126

copper 200 mesh (No. 01801, Ted Pella Inc, CA, USA) 127

and allowed to dry. Transmission Fourier transform 128

infrared (FTIR) spectra were acquired (8 cm-1 129

resolution 256 scans, at a sample compartment 130

vacuum pressure of 0.12 hPa) using a Bruker 66v/s 131

FTIR spectrometer. The sample was prepared by 132

depositing an aliquot of SAND solution to clean 133

double-side polished silicon (100) wafer substrates. 134

Spectra from a freshly plasma-cleaned silicon wafer 135

sample were collected before each measurement to 136

obtain the background spectrum. The X-ray 137

photoelectron spectroscopy (XPS) analysis sample 138

was prepared by placing 0.5 μl of the SAND 139

solution onto clean gold substrates at room 140

temperature. Elemental composition of the SAND 141

particles was determined by XPS using a JEOL JPS 142

9010 MX equipped with a monochromatic Mg Kα 143

X-ray radiation source. 144

145

2.3 Bicinchoninic acid (BCA) assays for the 146

determination of avidin 147

The measurements of protein concentrations by 148

BCA assay were performed with the Pierce® BCA 149

Protein Assay Kit (Thermo Fisher Scientific, 150

Germany). This assay estimates the formation of 151

Cu+ from Cu2+ by the Biuret complex in alkaline 152

solutions of protein (containing cysteine, cystine, 153

tryptophan and tyrosine or peptides containing 154

three or more amino acid residues) [12]. BCA reacts 155

with cuprous ion forming purple-colored reaction 156

product and measured at 562 nm in the UV–Vis 157

Spectrophotometer JASCO V-630. In this method, 10 158

μl of SAND-avidin solution (with different avidin 159

concentrations for 50, 25, 12.5, 6.5, 3.125, and 1.5 μg) 160

or 10 μl of SAND stock solution mixed with 190 μl 161

of deionized water were added to microtube. Then, 162

adding 0.8 ml of BCA in microtube at 65°C for 15 163

minutes and measuring at 562 nm in the UV/Vis 164

Spectrophotometer JASCO V-630. Blanks were 165

prepared by substituting SAND reacted with BCA. 166

167

2.4 The phenol-sulfuric acid spectrophotometric 168

method 169

The phenol-sulfuric acid spectrophotometric 170

method developed by Dubois [13] was adapted for 171

the determination of galactose binding on SAND. In 172

the method, 20 μl of SAND-galactose solution (with 173

different galactose concentrations for 2.5, 5, 10, 20, 174

40 and 80 mM) or 20 μl of SAND stock solution 175

mixed with 480 μl of 5% phenol solution were 176

added to glass vials, which were capped and 177

vortex-stirred. Then, using the concentrated sulfuric 178

acid (0.5 ml) to break down galactose (at room 179

temperature for 10 minutes) and to produce hexose 180

(6-carbon compounds) converting to 181

hydroxymethyl furfural. These compounds then 182

reacting with phenol to produce a yelloworange 183

color, which was measured at 490 nm in the UV/Vis 184

Spectrophotometer JASCO V-630. Blanks were 185

prepared by substituting SAND react with 186

phenol-sulfuric acid. The sulfuric acid (95-97%, ACS 187

reagent grade) was obtained from Scharlau 188

chemical and phenol (>98%) was purchased from 189

Shimakyu's Pure Chemical (Osaka, Japan). 190

191

2.5 Preparation of SAND solution and 192

functionalized SAND bioconjugates 193

We prepared a SAND solution by adding original 194

Page 6: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

5

SAND liquid (1 ml) into DMEM medium (7 ml; 195

Thermo Scientific) and adjusted the pH value to 196

7.0-7.3 by 20% acetic acid (2 ml). For preparation of 197

avidin-SAND conjugates and galactose-SAND 198

conjugates, 1.0 mg of avidin (Sigma) and galactose 199

(Sigma) were added respectively into 100 μl of 200

SAND solution (4.3 M), mixed on a Vortex-Mixer 201

(Scientific Industries, Inc.), and then left to stand for 202

15 min at room temperature. Subsequently, 300 μl 203

of 5% acetic acid was added to these two mixtures 204

to adjust the pH to 7.0-7.3 to obtain the stock 205

avidin-SAND (1.07 M) and galactose-SAND (1.07 M) 206

conjugates, respectively. The biomolecules avidin or 207

galactose conjugated with SANDs through either 208

physical adsorption, hydrogen bonding, or 209

electrostatic interactions, between amine groups of 210

the SAND and carboxyl groups of avidin [14] or 211

hydroxyl group of galactose [15] in aqueous 212

solution. 213

214

2.6 Pull-down analysis and biotin competition 215

assay 216

For isolation of avidin-SAND conjugates, a mixture 217

of avidin (Sigma) and SAND solution was pH 218

adjusted to 7.0-7.3 by 20% acetic acid. The 219

avidin-SAND conjugates solution was then filtered, 220

using a cut-off filtration, centricon centrifugation 221

and wash, whereby unconjugated SANDs and 222

avidin-SAND conjugates were separated. The 223

bright avidin-SAND conjugates which remained on 224

the upper column were collected. UV images and 225

protein absorbance (2 µ l conjugates or avidin 226

protein) demonstrated that the unconjugated SAND 227

particles were removed from the avidin-SAND 228

conjugates solution. For a biotin competition assay, 229

avidin-SAND conjugates (100 µ l) were co-incubated 230

with 2-iminobiotin agarose (100 µ l) for 10 min at 231

4˚C. After a series of centrifugation and basic buffer 232

rinsing, excess avidin-SAND conjugates were 233

removed and the 2-iminobiotin agarose pull-down 234

avidin-SAND conjugates were collected. Protein 235

quantification (2 µ l pull-down beads conjugates) 236

were recorded and analyzed. SK-Hep-1 cells were 237

incubated with the 2-iminobiotin-avidin-SAND 238

complex (50 μg/ml) for 48 h, and the uptake of 239

avidin-SAND conjugates and cell morphology were 240

analyzed by fluorescence microscopy. 241

242

2.7 Cell culture and in vitro cytotoxicity assay 243

SK-Hep-1 cells, HepG2 cells and HeLa cells were 244

maintained in DMEM medium containing 10% calf 245

serum (Hyclone) and supplemented with 100 IU/ml 246

penicillin and 100 g/ml streptomycin (Invitrogen) 247

in 5% CO2 at 37°C incubator temperature. For the 248

cytotoxicity assay, SK-Hep-1 cells were harvested 249

and seeded in a 96 well culture plate (1×104 cells/100 250

μl/well) and kept overnight at 37°C and 5% CO2 in a 251

humidified incubator. Cells were then treated with 252

various concentrations (0, 0.25, 1, 2, 5, 10, and 20 253

mM) of SAND for 48 h. Subsequently, the surviving 254

cells were analyzed by MTT 255

(3-(4,5-dimethy-thiazol-2-yl)-2,5-diphenyl 256

tetrazolium bromide) assay. 257

258

2.8 Fluorescence microscopy 259

SK-Hep-1 cells, HepG2 cells, and HeLa cells (4×104) 260

were seeded onto coverslips in 12-well culture plates 261

and were cultured at 37°C overnight. For protein 262

biolabling experiments, SK-Hep-1 cells were treated 263

with the SAND (20 mM), avidin (50 μg/ml), 264

avidin-FITC (50 μg/ml; Sigma) and avidin-SAND (20 265

mM) for 48 h. For saccharide sugar biolabling 266

experiments, HepG2 cells and HeLa cells were 267

treated with galactose-SAND conjugates (80 mM) for 268

2 h. Subsequently, these cell samples were washed 269

with PBS, fixed with 4% paraformaldehyde, and 270

perfused with 0.2% Triton X-100 for the inverted 271

fluorescence microscope (Leica, DMIL, Germany), 272

though the filters: DAPI, FITC and R-DIL. Image 273

processing and projections were performed using 274

Photoshop 5.5 (Adobe Systems, Mountain View, CA). 275

276

2.9 Flow cytometry 277

SK-Hep-1 cells (3×105) were seeded into 6 cm 278

culture plates overnight and then treated with 279

different concentrations (0, 2, 5, 10, and 20 mM) of 280

SAND for 48 h at 37°C. The cells were harvested, 281

washed twice with PBS, and fixed with 4% 282

paraformaldehyde, and perfused with 0.2% Triton 283

Page 7: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

6

X-100. Cells were suspended in the PBS and the 284

SAND uptake was analyzed by detection of the 285

fluorescence intensity in single-cell using flow 286

cytometry analysis (Beckman Coulter) with a 488 287

nm laser. At least 10,000 cells were counted. The 288

data is presented as mean fluorescence index (MFI) 289

and calculated as the ratio of the mean fluorescence 290

intensity of the sample and the control. Green and 291

red fluorescence were monitored using FITC and 292

PE channels, respectively. 293

294

3. Results and discussion 295

3.1 Structural and optical analysis to characterize 296

SANDs 297

Nanoparticles (NPs) such as inorganic 298

nanoparticles, quantum dots (QDs), magnetic NPs 299

(MPs), and multifunctional NPs, often have a 300

hydrophobic surfactant coating that limits their use 301

in biomedical applications. Therefore, methods to 302

functionalize NPs to increase water solubility prior 303

to use are desired [16, 17]. 304

The focus of this work is the fabrication and 305

characterization of a new type of SAND particle 306

derived from 3-aminopropyltriethoxysilanes 307

(APTES). The APTES organofunctional alkoxysilane 308

molecules self-assemble via attack and 309

displacement of the alkoxy groups, (usually the 310

ethoxy (–OCH2CH3) groups), forming a covalent 311

Si–O–Si bond [18, 19]. During SAND formation, the 312

hydrophobic alkoxysilane backbones are most 313

likely embedded inside the SAND particles while 314

the hydrophilic amine groups extend outside into 315

the aqueous environment. Therefore, APTES 316

SANDs may be highly water soluble and thus ideal 317

for a variety of biomedical applications. 318

Additionally, the terminal amine groups at the 319

outer surface of the SANDs would have a positive 320

surface potential [20] and could easily be 321

covalently-linked to other interesting chemicals or 322

biomolecules using established protocols. 323

Structural characterization of the SAND particles 324

was carried out by AFM and TEM. Fig. 1(a) shows 325

AFM topographical images of SANDs deposited on 326

a clean mica substrate. 327

Figure 1 Structural Characteristics of SAND particles. (a) AFM 328 topographic image of SANDs deposited on mica (1×1 μm2 329 scan). The actual AFM tip apex radius of curvature was on the 330 order of ~10-25 nm, so the apparent width of the particles is 331 larger due to the tip broadening effect. The measured height of 332 the particles, was used to determine the size, because the AFM 333 height measurement is not subject to this lateral tip-broadening 334 effect [8]. (b) Particle size distribution histogram of APTES 335 deposited on mica from (a). (c) TEM image with the lattice 336 fringes from HRTEM of the SAND particles. (d) PL spectra of 337 SAND solution recorded for progressively longer excitation 338 wavelengths from 280 to 520 nm in 20 nm increments. The 339 inset photographs show SAND solutions irradiated by visible 340 light and by an ultraviolet lamp (365 nm), respectively. 341

The SAND particles are spherical, monodisperse, 342

and well-separated on the mica surface. The 343

corresponding particle size histograms in Fig. 1(b) 344

show that the SAND particles have a narrow size 345

distribution with an average NP diameter of 2.5 ± 346

0.5 nm as determined from statistical analysis of 178 347

particles. The SAND particle size distribution was 348

further confirmed to be 2.4 ± 0.6 nm by the 349

high-resolution TEM image in Fig. 1(c). These 350

results show a homogeneous distribution and are 351

consistent with AFM results. The inset in Fig. 1(c) 352

shows characteristic lattice fringes of the SAND 353

particles measured at ~0.21 nm. Although we do not 354

know the exact structure of the carbon around the 355

SiOx core, this value is comparable with the (110) 356

planes of graphene [21]. 357

Surface chemistry and elemental composition of 358

the SAND particles was characterized by FTIR, 359

Raman and XPS. In FTIR spectra (Fig. S-1 in the 360

Electronic Supplementary Material (ESM)) we 361

observed peaks at 2935 and 2885 cm-1 that are 362

assigned to the symmetric and anti-symmetric 363

Page 8: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

7

–CH2– stretch, and indicate the presence of 364

anchored propyl groups. The vibrational mode 365

peaks at 1570 and 1485 cm-1 arise primarily from 366

surface amine groups [22]. Two peaks attributed to 367

the C–N stretching vibration were observed at 1125 368

and 1041 cm-1 [23]. Raman spectra were acquired 369

(Fig. S-2 in the ESM) with peaks at 2892 cm-1, 2936 370

cm-1 and 2980 cm-1, corresponding to the ν(CH2)sym, 371

ν(CH2)asym, and ν(CH3) vibrational modes. The 372

presence of these peaks confirm that alkyl groups 373

were still present in the SAND particles [24]. 374

Elemental analysis by XPS (Fig. S-3 in the ESM) 375

showed that SANDs were mainly composed of Si, O, 376

C, and N. The Si 2p peaks at 99.4 and 103.4 eV arise 377

from silicon and SiO2, respectively. The O 1s 378

spectrum at 533.2 eV indicates a Si–O–Si bonding 379

environment (Fig. S-3(a-b) in the ESM) [25]. The C 380

1s spectrum had one peak at 286.0 eV which was 381

attributed to C–N bonds [26], and the N 1s 382

spectrum showed a binding energy peak at 400.2 eV, 383

which was attributed to NH2 groups [19] (Fig. 384

S-3(c-d) in the ESM). These results confirm that bare 385

amine groups were present in the SAND and that 386

all expected vibrational modes of the coupled 387

ligand shell were present. Our elemental analysis 388

revealed that the SANDs were composed of SiO2 389

clusters with amine-alkyl groups. The sensitivity 390

factors for Si, O, C, and N, are 0.654, 2.29, 1, and 391

1.61, respectively [27]. The integrated peak areas 392

normalized with the sensitivity factors for the Si 2p : 393

O 1s : C 1s : N 1s ratios are 2.1 : 3.3 : 6.0 : 1.0. 394

Multicolor imaging is ideal for simultaneous 395

evaluation of multiple biomarkers within a complex 396

tissue for medical applications. Since cancer and 397

many diseases involve a large number of genes and 398

proteins, multicolor imaging could provide an 399

efficient and quantitative way to distinguish each 400

biomarker accurately. Nanoparticles exhibit bright 401

and multicolor photoluminescence [28], and are 402

thus an important tool for biomedical applications 403

such as biolabeling, biosensing and bioimaging [29, 404

30]. To characterize the optical properties of the 405

SAND particles, we performed absorption and 406

emission measurements. As demonstrated in Fig. 407

S-4 (in the ESM), the as-obtained SANDs show a 408

clear absorption feature centered at 365 nm which is 409

comparable to those previously reported for 410

fluorescent CDs [31, 32]. The PL spectra (Fig. 1(d)) 411

revealed that SANDs exhibit wideband and 412

multicolor fluorescence emission, the intensity of 413

which depends on the excitation wavelength. The 414

fluorescence emission spectra at several excitation 415

wavelengths varied from 280 to 520 nm. The PL 416

peak exhibited a fluorescence emission maximum 417

of λem = 502 nm at an excitation wavelength of 440 418

nm, above which the PL peak shifted to longer 419

wavelengths and its intensity decreased rapidly. 420

Further, it exhibited strong bright blue fluorescence 421

under excitation by an ultraviolet lamp at 365 nm 422

(Fig. 1(d) inset). 423

The quantum yield (QY) was measured using the 424

F-7000 fluorescence spectrophotometer combined 425

with a Hitachi 5J0-0148, 60-mm integrating sphere 426

with a coating of aluminum oxide (Hitachi Inc., 427

Japan). The QY was calculated by the built-in 428

quantum yield measurement system in the Hitachi 429

software using direct and indirect excitation. Direct 430

excitation is when the sample is directly facing the 431

excitation beam. Indirect excitation uses an 432

integrating sphere, where a portion of the emitted 433

photons from the sample are reflected from within 434

the integrating sphere and lead to further sample 435

excitation. The QY was obtained from direct and 436

indirect excitation, and the final QY values were 437

corrected by the following equation: 438

439

ϕ_QY is the corrected SAND quantum yield. ϕ_P1 440

is the internal QY using direct excitation, abs_P1 is 441

absorptance using direct excitation, and ϕ_P2 is the 442

internal QY using indirect excitation. The 443

measurement of QY for the progressive excitation 444

wavelengths from 320 to 520 nm in 20 nm 445

increments is shown in Table S-1 (in the ESM). The 446

SANDs had a maximum quantum yield of 29.5% at 447

an excitation wavelength of 440 nm, which may be 448

compared to the measured QY values for 449

silica-coated carbon dots (11-14.7%) [33], 450

laser-ablated graphite (4-10%) [34], and the thermal 451

oxidation of carbon dots (<5%) [35]. 452

Page 9: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

8

Figure 2 Dose-dependent cellular toxicity of SANDs to 453 SK-Hep-1 cells. Relative cell viabilities after 48 h treatment 454 with SANDs from MTT assay were plotted against SAND 455 concentration. No obvious toxicity was observed. Values are 456 expressed as mean ± S.D. of three individual experiments. 457

458

This result shows that SANDs have a high quantum 459

yield compared to the other types of NPs 460

mentioned above. 461

462

3.2 The biocompatibility and multifluorescence 463

of SANDs 464

To validate the performance of SANDs as 465

cell-imaging agents, cytotoxicity was evaluated 466

using human hepatoma SK-Hep-1 cells through the 467

MTT assay (Fig. 2). SK-Hep-1 cells were cultured in 468

a medium containing different concentrations of 469

SANDs (0, 0.25, 1, 2, 5, 10, and 20 mM) for 48 h and 470

resulted in cell viability values of 100 ± 2.3, 95 ± 3.24, 471

97 ± 3.76, 94 ± 3.11, 96 ± 3.18, 99 ± 1.56, and 98 ± 3.07 472

(% of control) respectively. This result indicated that 473

the SANDs had no significant cytotoxic effect on 474

SK-Hep-1 cells, and thus may be a good candidate 475

for in vitro labeling or cellular imaging. 476

We further studied the uptake efficiency and 477

multicolor fluorescence of SANDs by flow 478

cytometry (using FITC and PE filters to detect the 479

green and red fluorescence, respectively), because 480

cell fluorescence intensity correlates with the 481

number of internalized nanoparticles. In flow 482

cytometry dot plots, each dot represents an 483

individual cell. Cells appearing in the lower left 484

quadrant (I) exhibited no fluorescence. Cells 485

appearing in the upper left (II) and lower right 486

quadrant (IV) correspond to cells emitting either 487

red or green fluorescence, respectively. And finally, 488

cells in the upper right quadrant (III) are cells 489

emitting both green and red fluorescence. 490

After cells were treated with various 491

concentrations of SAND (0, 2, 5, 10, and 20 mM; Fig. 492

3(a-e)) for 48 h, the cells emitting both red and 493

green fluorescence were detected using two 494

fluorescent filters: FITC and PE, and revealed an 495

increasing SAND uptake efficiency (Fig. 3(a-e), 496

quadrant III). These results are shown in the 497

histogram plot (Fig. 3(f)) and indicate that ~19% of 498

detected cells took up SANDs at lower dose (2 mM) 499

and more than 95% of cells were observed with 500

high SANDs uptake at higher dose (20 mM). 501

Further, the separate detection of cells exhibiting 502

red (Fig. 3(a-e), quadrant II and quadrant III) or 503

green fluorescence (Fig. 3(a-e), quadrant III and 504

quadrant IV) were also analyzed, and histograms 505

plots (Figs. S-5 and S-6 in the ESM) showed that 506

these cells were elevated in a dose-dependent 507

manner. Conversely, the control cells, which 508

exhibited no fluorescence, were decreasing (Fig. 509

3(a-e), lower left quadrant I). 510

Figure 3 Fluorescence intensity monitored by flow cytometry. 511 SK-Hep-1 cells were treated with various concentrations of 512 SAND for 48 h. (a-e) Flow cytometry dot plot (Fluorescence of 513 x-axis and y-axis is FITC and PE, respectively). The number of 514 dots in each quadrant represents the fluorescence intensity of 515 the cells. The lower left-hand (I), upper left-hand (II), lower 516 right-hand (IV), and upper right-hand (III) quadrants indicate 517 non-fluorescent, red fluorescence, green fluorescence, and both 518 green and red fluorescence, respectively. (f) Bar graph 519 representation of the fluorescence intensity of cells from the dot 520 plot at various concentrations of SAND. Increased cell 521 fluorescence due to accumulation of fluorescent nanoparticles is 522 presented as mean fluorescence index (MFI). The results were 523 obtained from three separate experiments and the bar represents 524 mean ± S.D. of three individual experiments. *p < 0.01. 525

Page 10: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

9

Here, auto-fluorescence (2-4%) was occurring in 526

the control cells (Fig. 3(a), quadrant III and IV), but 527

it did not interfere with the multi-fluorescence 528

emission of SAND in SK-Hep-1 cells. This result 529

demonstrated the excellent biocompatible nature 530

and multi-fluorescence potential of SANDs. 531

Next, we explored in vitro internalization of 532

SANDs in SK-Hep-1 cells. Cells were cultured and 533

maintained in DMEM medium containing 20 mM of 534

SANDs for 48 h, with a detected concentration of 535

95% cellular SAND-internalization efficiency. The 536

cell fluorescence images (Fig. S-7 in the ESM) shows 537

bright field (a), blue (b), green (c), and red (d) 538

emission, respectively under fluorescence 539

microscopy. The multi-photoluminescence images 540

of SK-Hep-1 cells revealed that SANDs were 541

internalized and accumulated into the cytoplasm 542

region and that they surrounded, but did not enter 543

into the nucleus (Fig. S-7 in the ESM). Endocytosis 544

is the likely internalization mechanism of SANDs 545

from outside the cell membrane into the cytoplasm. 546

The cell membrane is a negatively charged 547

phospholipid bilayer which separates the 548

intracellular substance and organelles from the 549

extracellular space. SANDs have abundant surface 550

functional amine groups, which imparts to them 551

more cationic characteristics. They thereby 552

interacted electrostatically with the anionic cell 553

membrane and were further taken up through an 554

adsorptive endocytosis mechanism and internalized 555

by cells [36-38]. More detailed internalization 556

mechanisms of SANDs will require additional 557

experiments. 558

559

3.3 Biomolecule labeling and imaging with 560

SANDs 561

Luminescent NPs are very attractive because of 562

their potential for applications in biolabeling, 563

bioimaging [6] and biomedicine [4]. However, these 564

hydrophobic core NPs frequently needs complex 565

grafting to provide the necessary surface chemistry 566

[6]. Our SANDs are derived from APTES, which is 567

commonly used for surface functionalization to 568

prepare amine-terminated (–NH2) silica-based 569

materials, or to convert nanoparticles from 570

hydrophobic to hydrophilic character [39, 40]. The 571

presence of amine groups on the surface of targets 572

is important for biological applications, such as 573

biolabeling [20] and clinical immunosensor [41], as 574

it allows proteins or other biomolecules to be 575

attached in a simple manner [14, 42]. Thereby, 576

SANDs prepared using APTES as the precursor 577

possess the characteristics of APTES and can be 578

effective tools in the biological sciences for such 579

applications as biolabeling, cell imaging, biosensors, 580

and biomimetic membranes. 581

Before the cell-binding and cell-labeling 582

experiments, we used bioconjugate techniques for 583

the determination of whether a particular protein or 584

sugar is located on the SANDs. The bicinchoninic 585

acid protein assay was performed to measure the 586

total avidin content on SANDs (Fig. S-8 in the ESM), 587

and the phenol-sulfuric acid spectrophotometric 588

method was performed to monitor the 589

incorporation of galactose on SANDs (Fig. S-9 in the 590

ESM). Repeatability and reproducibility were 591

evaluated from the results obtained by two analysts 592

working on the same sample, which exhibit a linear 593

relationship between the absorbance and the 594

varying concentrations. 595

Avidin is a tetramic protein [43] and is bound to 596

biotin with high affinity and specificity [44]. It is 597

used in many biochemical assays, including 598

western blot, ELISA, and pull-down assays. To 599

demonstrate the feasibility of SANDs as a protein 600

labeling probe, we conjugated SANDs with avidin 601

(Fig. S-10 and S-11 in the ESM). To validate the 602

conjugated SAND interaction with biotin, a pull 603

down assay was executed using a commercial 604

2-iminobiotin agarose (Fig. S-12 in the ESM). Our 605

data showed that SANDs are successfully 606

conjugated to avidin and implied that avidin-SAND 607

conjugates could be taken up via a biomolecule 608

interaction-mediated uptake within the cell. We 609

chose SK-Hep-1 cells, which are a good source of 610

biotin [45, 46], for the cell targeting experiment. 611

SK-Hep-1 cells were cultured in a medium 612

containing SANDs, avidin, avidin-FITC, and 613

SAND-avidin conjugates for 48 h, and the particle 614

uptake efficiency was then analyzed using 615

Page 11: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

10

fluorescence microscopy. Fig. 4 shows that the cells 616

internalized the avidin-SAND conjugates over the 617

SANDs and avidin-FITC, thus demonstrating their 618

potential for protein labeling. 619

The SAND uptake was monitored using bright 620

field (Fig. 4(a)) and DAPI, FITC, and R-DIL filters 621

under fluorescence microscopy. SANDs were 622

internalized in the cytoplasm and surrounded the 623

cell nucleus, as shown in the blue, green and red 624

fluorescence data shown in Fig. 4(b-d). This result is 625

consistent with the data of Fig. S-7 in the ESM. Cells 626

that were incubated with avidin (Fig. 4(e)), as a 627

fluorescence negative control exhibited no 628

fluorescence signal (Fig. 4 (f-h)). 629

SK-Hep-1 cells treated with the avidin-FITC (Fig. 630

4(i)), which is a well-known organic dye and used 631

as a fluorescence positive control, showed bright 632

green fluorescence. Here, avidin-FITC internalized 633

in the cytoplasm (Fig. 4(k)), but the fluorescence 634

quickly decayed (Fig. S-13(d-f) in the ESM) using 635

excitation comparable to the SAND treated group 636

(Fig. S-13(a-c) in the ESM). 637

Figure 4 Comparison of the uptake of SANDs, avidin, 638 avidin-FITC conjugates, and SAND-avidin conjugates by 639 human hepatoma cell line. SK-Hep-1 cells were treated with 640 SAND (20 mM) (a-d), avidin (50 μg/ml) (e-h), avidin-FITC (50 641 μg/ml) (i-l) and SAND-avidin conjugates (20 mM) (m-p), 642 respectively for 48 h. Avidin was used as a fluorescent negative 643 control and avidin-FITC as a fluorescent internalized positive 644 control. The uptake of nanodots and cell morphology were 645 analyzed by fluorescence microscopy. The surface NH2 groups 646 of the SAND particles interacted with avidin and thus the 647 avidin-SAND conjugates were more internalized into the 648 cytoplasm than SANDs or avidin-FITC treated cells. The 649 cellular multi-photoluminescence images were monitored using 650 DAPI, FITC and R-DIL channels, respectively. Original 651 magnification was 400X. 652

Not surprisingly, avidin-FITC treatment showed 653

no blue or red signals (Fig. 4(j-l)). These results 654

show that SANDs were more photostable than the 655

organic dye-FITC, which suffered from significant 656

photobleaching. Further, we tested the SAND 657

solution stability at different pH values (0, 2, 4, 6, 8, 658

10, 12, 14) and in phosphate buffered saline (PBS) 659

and sodium chloride (NaCl) to mimic various 660

normal and extreme physiological conditions 661

(Figure S-14 and S-15 in the ESM). The PL results 662

from these experiments show that the SANDs were 663

stable under a variety of harsh solution conditions 664

(Table S-2 in the ESM). 665

Finally, we treated cells with avidin-SAND 666

conjugates (Fig. 4(n-p)) and the results revealed not 667

only the characteristic multi-color fluorescence 668

signals consistent with the SAND treated group 669

(Fig. 4(b-d)), but also showed agglomerated 670

fluorescent particles unlike in the SANDs and 671

avidin-FITC treated groups (Fig. 4(k)). In addition, 672

the agglomerated avidin-SAND conjugates showed 673

good distribution in the cytoplasm and around the 674

nucleus. The avidin-SAND conjugates (Fig. 4(n-p)) 675

were more accessible and accumulated in the cells 676

more efficiently than SAND (Fig. 4(a-d)) or 677

avidin-FITC (Fig. 4(k)) during an equivalent 678

incubation time. Moreover, a competition assay was 679

also executed. After avidin-SANDs conjugates were 680

pre-precipitated by 2-iminobiotin agarose and then 681

treated on SK-Hep-1 cells for 48 h. The cells 682

revealed that no fluorescence could be observed 683

(Fig. S-16 in the ESM) when compared to 684

avidin-SANDs conjugates and SANDs groups. This 685

result suggests that exogenous biotin interacts with 686

avidin-SAND conjugates and further interfered 687

with the internalization of avidin-SAND conjugates. 688

The bright fluorescence from avidin-SAND 689

conjugates revealed the high degree of avidin-biotin 690

interaction in SK-Hep-1 cells. Although the 691

avidin-FITC group was expected to exhibit a high 692

degree of avidin-biotin interaction, as in the 693

avidin-SAND conjugate group, both SANDs and 694

the avidin-SAND conjugates exhibited much more 695

stable fluorescence than the avidin-FITC group 696

when cells were exposed to continuous excitation 697

Page 12: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

11

(Fig. S-13 in the ESM). This indicates that SANDs 698

are more photostable and photobleaching-resistant 699

than avidin-FITC. We note also that the rapid 700

photobleaching of organic dyes has been reported 701

in the literature and is a significant limitation on 702

their use in bio-applications [3]. 703

Another example demonstrates targeted cell 704

imaging through SANDs conjugated with 705

carbohydrate-galactose to examine their feasibility 706

as a biomolecule probe. Carbohydrate 707

functionalized nanoparticles have been successfully 708

administered as nanoprobes for cellular targeting or 709

bioimaging applications and drug delivery in vivo 710

and in vitro [47, 48]. For example, an anti-cancer 711

drug-paclitaxel was loaded into a galactose 712

conjugated nanoparticle (NP-Gal-PTX) and 713

significantly inhibited HepG2 cell viability through 714

the interaction between galactose and 715

asialoglycoprotein receptor (ASGP-R) [47]. Here, 716

galactose-SAND conjugates were formed through 717

hydrogen bonding between the hydroxyl end of the 718

galactose and the primary amine group of the 719

nanoparticles [49]. Two cell lines, human 720

hepatocellular carcinoma cells (HepG2) and human 721

cervical cancer cell line (HeLa) were specifically 722

chosen for cell targeting experiments because 723

HepG2 cells express ASGP-R on the cell surface (the 724

galactose binding protein), whereas HeLa cells 725

without ASGP-R expression were chosen as the 726

negative control [50] (Fig. 5). 727

Both HepG2 and HeLa cells were incubated in 728

medium containing galactose-SAND conjugates (80 729

mM) for 2 h and then the cell targeting efficiency 730

was analyzed using fluorescence microscopy. Fig. 731

5(a) shows a healthy HepG2 cell without survival 732

stress from the higher concentration SAND 733

treatment. The strong signals of blue, green, and red 734

multicolor fluorescence demonstrate that the 735

galactose-SAND conjugate interacts strongly with 736

ASGP-R-positive HepG2 cells (Fig. 5(b-d)). Our 737

galactose-SAND conjugate can specifically target 738

the cell surface and then is internalized by HepG2 739

cells through the receptor-mediated endocytosis 740

process [51]. 741

742

Figure 5 The receptor mediated uptake of SAND-galactose 743 conjugates in hepatoma cell line-HepG2 cells. Images of 744 HepG2 cells (a-d) and HeLa cells (e-h) cells treated with 745 SAND-galactose conjugates (80 mM) for 2 h and their 746 corresponding optical images. The SAND-galactose conjugates 747 exhibited specific binding and uptake by HepG2 cells but not 748 by HeLa cells. The cellular multi-photoluminescence images 749 were monitored using DAPI, FITC and R-DIL channels, 750 respectively. Original magnification was 400X. 751

752

On the other hand, the ASGP-R-negative HeLa cells 753

(Fig. 5(e)) revealed that the high dose of SAND had 754

an insignificant effect on cell survival. As expected, 755

the negative ASGP-R expressed-HeLa cells showed 756

very weak fluorescence (Fig. 5(f-h)). This 757

observation demonstrates that functionalized 758

SANDs can be modified with specific ligands for 759

targeting biomolecule detection, such as 760

carbohydrates, and have potential for further 761

biomedical and biosensing applications. 762

In this present study, we have demonstrated that 763

bright and substrate-independent 764

functionalized-SANDs are an effective labeling 765

probe for proteins and carbohydrates. Further, 766

peptides, drugs, or biomolecules (i.e. DNA) can be 767

directly modified or conjugated with the APTES 768

derived SANDs via the amine-terminated (–NH2) 769

functional groups. This study demonstrates a new 770

and novel application for alkoxysilane 771

nanoparticles in bioimaging and biolabeling. 772

773

Page 13: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

12

4. Conclusion 774

In summary, our study offers a novel systematic 775

study of a new nanomaterial for bioimaging and 776

biolabeling. The APTES-derived SANDs have 777

several advantages including 778

substrate-independent functionalization, high water 779

solubility, biocompatibility, negligible cytotoxicity, 780

strong photoluminescence, and high photostability. 781

Moreover, we demonstrate the feasibility of SANDs 782

as a biomolecular labeling probe through direct 783

functionalization to obtain surface coatings for 784

specific biomolecular conjugates without secondary 785

chemical multi-modification or difficult 786

functionality processing. Hence, this new type of 787

nanodot has strong potential for the development of 788

in vitro biochemical assays and could be used 789

advantageously for in vivo imaging, disease 790

detection, and cancer diagnosis. 791

792

Acknowledgements 793

M. L. Kung and P. Y. Lin contributed equally to this 794

work. The authors would like to thank the Ministry 795

of Science and Technology (NSC 796

101-2113-M-110-013-MY3) and (NSC 797

101-2811-M-110-034 and NSC 102-2811-M-110-008) of 798

Taiwan and the National Sun Yat-sen University 799

Center for Nanoscience and Nanotechnology for 800

financial support of this work. Prof. Hsieh also 801

thanks Dr. Ming-Hong Tai for providing the 802

fluorescent microscope and SK-Hep-1 cell line, and 803

Dr. David Beck for helpful discussions. 804

805

Electronic Supplementary Material: Supplementary 806

material (Additional SANDs characterization data 807

including FTIR, Raman, and XPS, as well as 808

biolabeling and bioimaging applications of SANDs.) 809

is available in the online version of this article at 810

http://dx.doi.org/10.1007/s12274-***-****-* 811

812

References 813

[1] Bailey, R. E.; Smith, A. M.; Nie, S. M. Quantum dots in 814

biology and medicine. Physica E 2004, 25, 1-12. 815

[2] Jaiswal, J. K.; Goldman, E. R.; Mattoussi, H.; Simon, S. 816

M. Use of quantum dots for live cell imaging. Nat. 817

Methods 2004, 1, 73-78. 818

[3] Resch-Genger, U.; Grabolle, M.; Cavaliere-Jaricot, S.; 819

Nitschke, R.; Nann, T. Quantum dots versus organic dyes 820

as fluorescent labels. Nat. Methods 2008, 5, 763-775. 821

[4] Michalet, X.; Pinaud, F. F.; Bentolila, L. A.; Tsay, J. M.; 822

Doose, S.; Li, J. J.; Sundaresan, G.; Wu, A. M.; Gambhir, 823

S. S.; Weiss, S. Quantum dots for live cells, in vivo 824

imaging, and diagnostics. Science 2005, 307, 538-544. 825

[5] Smith, A. M.; Dave, S.; Nie, S.; True, L. Gao, X. 826

Multicolor quantum dots for molecular diagnostics of 827

cancer. Expert Rev. Mol. Diagn. 2006, 6, 231-244 828

[6] Selvan, S. T.; Tan, T. T. Y.; Yi, D. K.; Jana, N. R. 829

Functional and multifunctional nanoparticles for 830

bioimaging and biosensing. Langmuir 2010, 26, 831

11631-11641. 832

[7] King-Heiden, T. C.; Wiecinski, P. N.; Mangham, A. N.; 833

Metz, K. M.; Nesbit, D.; Pedersen, J. A.; Hamers, R. J.; 834

Heideman, W.; Peterson, R. E. Quantum dot nanotoxicity 835

assessment using the Zebrafish embryo. Environ. Sci. 836

Technol. 2009, 43, 1605-1611. 837

[8] Lin, P. Y.; Hsieh, C. W.; Kung, M. L.; Hsieh, S. 838

Substrate-free self-assembled SiOx-core nanodots from 839

alkylalkoxysilane as a multicolor photoluminescence 840

source for intravital imaging. Sci. Rep. 2013, 3, 1703. 841

[9] Sun, X.; Wei, W. Electrostatic-assembly-driven formation 842

of micrometer-scale supramolecular sheets of 843

(3-aminopropyl)triethoxysilane(APTES)-HAuCl4 and 844

their subsequent transformation into stable APTES 845

bilayer-capped gold nanoparticles through a thermal 846

process. Langmuir 2010, 26, 6133-6135. 847

[10] Chai, C.; Lee, J.; Takhistov, P. Direct detection of the 848

biological toxin in acidic environment by electrochemical 849

impedimetric immunosensor. Sensors 2010, 10, 850

11414-11427. 851

[11] Faucheux, N.; Schweiss, R.; Lützow, K.; Werner, C.; 852

Groth, T. Self-assembled monolayers with different 853

terminating groups as model substrates for cell adhesion 854

studies. Biomaterials 2004, 25, 2721-2730. 855

[12] Smith, P. K.; Krohn, R. I.; Hermanson, G. T.; Mallia, A. 856

K.; Gartner, F. H.; Provenzano, M. D.; Fujimoto, E. K.; 857

Goeke, N. M.; Olson, B. J.; Klenk, D. C. Measurement of 858

Page 14: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

13

protein using bicinchoninic acid. Anal. Biochem. 1985, 859

150, 76-85. 860

[13] DuBois, M.; Gilles, K. A.; Hamilton, J. K.; Rebers, P. A.; 861

Smith, F. Colorimetric method for determination of 862

Sugars and related substances. Anal. Chem. 1956, 28, 863

350-356. 864

[14] Aissaoui, N.; Bergaoui, L.; Landoulsi, J.; Lambert, J. F.; 865

Boujday, S. Silane layers on silicon surfaces: Mechanism 866

of interaction, stability, and influence on protein 867

adsorption. Langmuir 2012, 28, 656-665. 868

[15] Llewellyn, N. M.; Spencer, J. B. Chemoenzymatic 869

acylation of aminoglycoside antibiotics. Chem. Commun. 870

2008, 32, 3786-3788. 871

[16] Gerion, D.; Pinaud, F.; Williams, S. C.; Parak, W. J.; 872

Zanchet, D.; Weiss, S.; Alivisatos, A. P. Synthesis and 873

properties of biocompatible water-soluble silica-coated 874

CdSe/ZnS semiconductor quantum dots. J. Phys. Chem. B 875

2001, 105, 8861-8871. 876

[17] Pellegrino, T.; Manna, L.; Kudera, S.; Liedl, T.; Koktysh, 877

D.; Rogach, A. L.; Keller, S.; Radler, J.; Natile, G.; Parak, 878

W. J. Hydrophobic nanocrystals coated with an 879

amphiphilic polymer shell: A general route to water 880

soluble nanocrystals. Nano Lett. 2004, 4, 703-707. 881

[18] Chen, P. C.; Chen, Y. N.; Hsu, P. C.; Shih, C. C.; Chang, 882

H. T. Photoluminescent organosilane-functionalized 883

carbon dots as temperature probes. Chem. Commun. 2013, 884

49, 1639-1641. 885

[19] Wang, F.; Xie, Z.; Zhang, H.; Liu, C. Y.; Zhang, Y. G. 886

Highly luminescent organosilane-functionalized carbon 887

dots. Adv. Funct. Mater. 2011, 21, 1027-1031. 888

[20] Cheang, T. Y.; Tang, B.; Xu, A. W.; Chang, G. Q.; Hu, Z. 889

J.; He, W. L.; Xing, Z. H.; Xu, J. B.; Wang, M.; Wang, S. 890

M. Promising plasmid DNA vector based on 891

APTES-modified silica nanoparticles. Int. J. Nanomed. 892

2012, 7, 1061-1067. 893

[21] Lu, G.; Mao, S.; Park, S.; Ruoff, R.; Chen, J. Facile, 894

noncovalent decoration of graphene oxide sheets with 895

nanocrystals. Nano Res. 2009, 2, 192-200. 896

[22] Kim, J.; Seidler, P.; Wan, L. S.; Fill, C. Formation, 897

structure, and reactivity of amino-terminated organic 898

films on silicon substrates. J. Colloid Interf. Sci. 2009, 899

329, 114-119. 900

[23] Evans, D. The systematic identification of organic 901

compounds. J. Chem. Educ. 1999, 76, 1069. 902

[24] Bistričić, L.; Volovšek, V.; Dananić, V. Conformational 903

and vibrational analysis of 904

gamma-aminopropyltriethoxysilane. J. Mol. Struct. 2007, 905

834–836, 355-363. 906

[25] Vandenberg, E. T.; Bertilsson, L.; Liedberg, B.; Uvdal, K.; 907

Erlandsson, R.; Elwing, H.; Lundstrom, I. Structure of 908

3-aminopropyl triethoxy silane on silicon-oxide. J. 909

Colloid Interf. Sci. 1991, 147, 103-118. 910

[26] Léandri, C.; Oughaddou, H.; Aufray, B.; Gay, J. M.; Le 911

Lay, G.; Ranguis, A.; Garreau, Y. Growth of Si 912

nanostructures on Ag(001). Surf. Sci. 2007, 601, 262-267. 913

[27] Seah, M. P.; Gilmore, I. S.; Spencer, S. J. Quantitative 914

XPS: I. Analysis of X-ray photoelectron intensities from 915

elemental data in a digital photoelectron database. J. 916

Electron Spectrosc. 2001, 120, 93-111. 917

[28] Sharma, P.; Brown, S.; Walter, G.; Santra, S.; Moudgil, B. 918

Nanoparticles for bioimaging. Adv. Colloid Interfac. 2006, 919

123-126, 471-485. 920

[29] Eck, W.; Nicholson, A. I.; Zentgraf, H.; Semmler, W.; 921

Bartling, S. Anti-CD4-targeted gold nanoparticles induce 922

specific contrast enhancement of peripheral lymph nodes 923

in X-ray computed tomography of live mice. Nano Lett. 924

2010, 10, 2318-2322. 925

[30] Zhuo, Y.; Chai, Y. Q.; Yuan, R.; Mao, L.; Yuan, Y. L.; Han, 926

J. Glucose oxidase and ferrocene labels immobilized at 927

Au/TiO2 nanocomposites with high load amount and 928

activity for sensitive immunoelectrochemical 929

measurement of ProGRP biomarker. Biosens. Bioelectron. 930

2011, 26, 3838-3844. 931

[31] Qiao, Z. A.; Wang, Y.; Gao, Y.; Li, H.; Dai, T.; Liu, Y. 932

Huo, Q. Commercially activated carbon as the source for 933

producing multicolor photoluminescent carbon dots by 934

chemical oxidation. Chem. Commun. 2010, 46, 935

8812-8814. 936

[32] Li, Y.; Hu, Y.; Zhao, Y.; Shi, G.; Deng, L.; Hou, Y. Qu, L. 937

An Electrochemical avenue to green-luminescent 938

Page 15: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

14

graphene quantum dots as potential electron-acceptors for 939

photovoltaics. Adv. Mater. 2011, 23, 776-780 940

[33] Liu, R.; Wu, D.; Liu, S.; Koynov, K.; Knoll, W.; Li, Q. An 941

aqueous route to multicolor photoluminescent carbon dots 942

using silica spheres as carriers. Angew. Chem. Int. Ed. 943

Engl. 2009, 48, 4598-4601. 944

[34] Sun, Y. P.; Zhou, B.; Lin, Y.; Wang, W.; Fernando, K. A. 945

S.; Pathak, P.; Meziani, M. J.; Harruff, B. A.; Wang, X.; 946

Wang, H., et al. Quantum-sized carbon dots for bright and 947

colorful photoluminescence. J. Am. Chem. Soc. 2006, 128, 948

7756-7757. 949

[35] Liu, H.; Ye, T. Mao, C. Fluorescent carbon nanoparticles 950

derived from candle soot. Angew. Chem. Int. Ed. Engl. 951

2007, 46, 6473-6475. 952

[36] Doherty, G. J.; McMahon, H. T. Mechanisms of 953

endocytosis. Annu. Rev. Biochem. 2009, 78, 857-902. 954

[37] Hillaireau, H.; Couvreur, P. Nanocarriers' entry into the 955

cell: relevance to drug delivery. Cell. Mol. Life Sci. 2009, 956

66, 2873-2896. 957

[38] Perumal, O. P.; Inapagolla, R.; Kannan, S.; Kannan, R. M. 958

The effect of surface functionality on cellular trafficking 959

of dendrimers. Biomaterials 2008, 29, 3469-3476. 960

[39] Liu, T.; Wang, S.; Chen, G. Immobilization of trypsin on 961

silica-coated fiberglass core in microchip for highly 962

efficient proteolysis. Talanta 2009, 77, 1767-1773. 963

[40] Jang, L. S.; Liu, H. J. Fabrication of protein chips based 964

on 3-aminopropyltriethoxysilane as a monolayer. Biomed. 965

Microdevices 2009, 11, 331-338. 966

[41] Zhong, Z. Y.; Shan, J. L.; Zhang, Z. M.; Qing, Y.; Wang, 967

D. The signal-enhanced label-free immunosensor based 968

on assembly of prussian blue-SiO2 nanocomposite for 969

amperometric measurement of neuron-specific enolase. 970

Electroanalysis 2010, 22, 2569-2575. 971

[42] Chuang, Y. H.; Chang, Y. T.; Liu, K. L.; Chang, H. Y.; 972

Yew, T. R. Electrical impedimetric biosensors for liver 973

function detection. Biosens. Bioelectron. 2011, 28, 974

368-372. 975

[43] Livnah, O.; Bayer, E. A.; Wilchek, M.; Sussman, J. L. 976

Three-dimensional structures of avidin and the 977

avidin-biotin complex. Proc. Natl. Acad. Sci. U. S. A. 978

1993, 90, 5076-5080. 979

[44] Hiller, Y.; Gershoni, J. M.; Bayer, E. A.; Wilchek, M. 980

Biotin binding to avidin. Oligosaccharide side chain not 981

required for ligand association. Biochem. J. 1987, 248, 982

167-171. 983

[45] Said, H. M.; Ma, T. Y.; Kamanna, V. S. Uptake of biotin 984

by human hepatoma cell line, Hep G2: a carrier-mediated 985

process similar to that of normal liver. J. Cell. Physiol. 986

1994, 161, 483-489. 987

[46] Zempleni, J.; Mock, D. M. Biotin biochemistry and 988

human requirements. J. Nutr. Biochem. 1999, 10, 989

128-138. 990

[47] Wang, Y. C.; Liu, X. Q.; Sun, T. M.; Xiong, M. H.; Wang, 991

J. Functionalized micelles from block copolymer of 992

polyphosphoester and poly(epsilon-caprolactone) for 993

receptor-mediated drug delivery. J. Control. Release 994

2008, 128, 32-40. 995

[48] David, S.; Passirani, C.; Carmoy, N.; Morille, M.; Mevel, 996

M.; Chatin, B.; Benoit, J. P.; Montier, T.; Pitard, B. DNA 997

nanocarriers for systemic administration: characterization 998

and in vivo bioimaging in healthy mice. Mol. Ther. 999

Nucleic Acids 2013, 2, e64. 1000

[49] Nagabhushan, T. L.; Cooper, A. B.; Turner, W. N.; Tsai, 1001

H.; McCombie, S.; Mallams, A. K.; Rane, D.; Wright, J. 1002

J.; Reichert, P. Interaction of vicinal and nonvicinal 1003

amino-hydroxy group pairs in 1004

aminoglycoside-aminocyclitol antibiotics with transition 1005

metal cations. Selective N protection. J. Am. Chem. Soc. 1006

1978, 100, 5253-5254. 1007

[50] Basiruddin, S. K.; Ranjan Maity, A.; Jana, N. R. 1008

Glucose/galactose/dextran-functionalized quantum dots, 1009

iron oxide and doped semiconductor nanoparticles with 1010

<100 nm hydrodynamic diameter. RSC Advances 2012, 2, 1011

11915-11921. 1012

[51] Khorev, O.; Stokmaier, D.; Schwardt, O.; Cutting, B.; 1013

Ernst, B. Trivalent, Gal/GalNAc-containing ligands 1014

designed for the asialoglycoprotein receptor. Biorg. Med. 1015

Chem. 2008, 16, 5216-523. 1016

.1017

Page 16: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S1

Electronic Supplementary Material

Correct Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling

Mei-Lang Kung,1 Pei-Ying Lin,

1 Chiung-Wen Hsieh,

1 and Shuchen Hsieh

1,2 ()

1Department of Chemistry and Center for Nanoscience and Nanotechnology, “National Sun Yat-sen University”, Kaohsiung, 80424 2School of Pharmacy, College of Pharmacy, Kaohsiung Medical University, Kaohsiung, 80707

Supporting information to DOI 10.1007/s12274-****-****-* (automatically inserted by the publisher)

Page 17: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S2

Supplementary Figures

Figure S1 FTIR spectra of SAND particles deposited on a silicon substrate. The presence of the anchored propyl

group was confirmed by the presence of symmetric and anti-symmetric –CH2– stretching peaks at 2935 and 2885

cm-1

. The major contribution of the vibrational modes at 1570 and 1485 cm-1

arises from surface amine groups.

Peaks corresponding to C–N stretching vibrations were observed at 1125 and 1041 cm-1

.

Page 18: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S3

Figure S2 Raman spectra of SAND particles deposited on a silicon substrate. The peak at 669 cm-1

and 961 cm-1

correspond to a Si–O stretch. The peak at 2892 cm-1

, 2936 cm-1

and 2980 cm-1

correspond to the ν(CH2)sym,

ν(CH2)asym, and ν(CH3) stretching peaks, indicating the presence of alkyl groups.

Page 19: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S4

Figure S3 Near-surface elemental analysis of the SAND using XPS. XPS data of SANDs deposited on gold

substrate, showing (a) Si 2p, (b) O 1s, (c) C 1s, and (d) N 1s binding energy peaks. The Si 2p peak at 103.4 eV

corresponds to SiO2, and the lower binding energy peak at 99.4 eV indicates pure silicon. Figure (b) shows the O

1s spectra at 533.2 eV which indicates a Si–O–Si bonding environment. Figure (c) shows the C 1s spectrum with a

binding energy peak at 286.0 eV. This corresponds well to C–N carbon and confirms that the alkyl chains from the

precursor are preserved in the NP product. Figure (d) shows the N 1s spectra with a peak at 400.2 eV which is

attributed to NH2 groups.

Page 20: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S5

Figure S4 UV-Vis absorption spectra of SANDs.

Page 21: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S6

Table S1 Summary for the QY for SAND particle solution at different emission wavelengths.

Page 22: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S7

Figure S5 Red fluorescence monitored by flow cytometry. SK-Hep-1 cells were treated with different

concentrations of SANDs for 48 h. (a-e) In these histogram plots, grey filled regions indicate control normal cells;

red unfilled regions indicate SAND-treated cells. (f) Bar graph representation of the red fluorescence intensity at

various concentrations of SANDs. *p < 0.01.

Page 23: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S8

Figure S6 Green fluorescence monitored by flow cytometry. SK-Hep-1 cells were treated with different

concentrations of SANDs for 48 h. (a-e) In these histogram plots, grey filled regions indicate control normal cells;

green unfilled regions indicate SAND-treated cells. (f) Bar graph representation of the green fluorescence intensity

at various concentrations of SANDs. *p < 0.01.

Page 24: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S9

Figure S7 Uptake of SANDs nanodots by SK-Hep-1 cells. SK-Hep-1 cells were treated with SANDs (20 mM) for

48 h. The cells were fixed and then imaged at bright field (a). The blue, green, and red fluorescence (b-d) from

SAND were recorded using DAPI, FITC and R-DIL filters, respectively. Multicolor fluorescent emission as a

function of excitation wavelength was observed.

Page 25: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S10

Figure S8 The concentrations of avidin conjugation with SAND were determined by bicinchoninic acid-based

protein assay. Each experiment was repeated for three times. (Cavidin: avidin conjugated with SAND at different

concentrations)

Page 26: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S11

Figure S9 The contents of galactose binding to SAND were analyzed by the phenol-sulfuric acid method. Each

experiment was repeated for three times. (Cgalactose: galactose binding to SAND at different concentrations)

Page 27: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S12

Figure S10 Isolation of Avidin-SAND conjugates. Mixture of avidin (Sigma) and SAND solution was pH adjusted

to 7.0-7.3 using 20% acetic acid. After the avidin-SAND conjugate solution was filtered, centrifuged, and rinsed,

unconjugated SANDs and avidin-SAND conjugates were separated. The bright avidin-SAND conjugates which

remain on the upper column were collected. (a) The UV image and (b) protein absorbance (2 µl conjugates or

avidin protein) demonstrated that the unconjugated SAND particles were removed from avidin-SAND the

conjugate solution.

Page 28: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S13

Figure S11 FTIR spectra of SAND, avidin, and conjugated SAND-avidin solutions.

Page 29: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S14

Figure S12 A biotin pull-down experiment using biotin agarose to precipitate Avidin-SAND conjugates.

Avidin-SAND conjugates (100 µl) were co-incubated with 2-iminobiotin agarose (100 µl) for 10 min at 4˚C. After

a series of centrifugation and basic buffer rinsing, excess avidin-SAND conjugates were removed and the

2-iminobiotin agarose pull-down avidin-SAND conjugates were collected. (a) The UV image and (b) protein

absorbance (2 µl pull-down beads conjugates) were recorded and analyzed.

Page 30: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S15

Figure S13 SAND photostability compared to FITC. Consecutive images of SAND (a-c), avidin-FITC (d-f) and

avidin-SAND (g-i) expressed in SK-Hep-1 cells. These images represent the SAND, avidin-FITC and

avidin-SAND fluorescence intensity respectively through a 5 min period during which three fluorophores were

exposed to continuous excitation.

Page 31: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S16

Figure S14 Photographs of SAND solutions under different pH conditions illuminated with ambient light (a) and

365 nm UV radiation (b).

Page 32: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S17

Figure S15 Photographs of SAND solutions in two kinds of salt solutions under ambient light (a) and under 365

nm UV illumination (b).

Page 33: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S18

Table S2 Summary of SAND emission wavelengths under different pH conditions and under different salts

environments measured in the 280-520 nm range.

Page 34: Aqueous Self-Assembly and Surface-Functionalized Nanodots ... · 2 Aqueous Self-Assembly and Surface-Functionalized Nanodots for Live Cell Imaging and Labeling Mei-Lang Kung,1 Pei-Ying

S19

Figure S16 Biotin competition assay. After the 2-iminobiotin agarose pull-down avidin-SAND

conjugates were collected, we incubated SK-Hep-1 cells with SAND (a-d), avidin-SAND (e-h) and

the 2-iminobiotin-avidin-SAND complex (50 μg/ml) (i-l) for 48 h. The nanoparticle uptake and cell

morphology were analyzed by fluorescence microscopy. SANDs were taken up and surrounded the

cellular nucleus. The avidin-SAND conjugates were much more internalized into the cytoplasm than

SANDs. Moreover, following the 2-iminobiotin-avidin-SAND complex treatment, cells exhibited

no fluorescence. This result shows that exogenous biotin interfered with avidin-SAND conjugate

cellular uptake, and supports our assertion that the avidin and SAND were conjugated. The cellular

multi-photoluminescence images were monitored using DAPI, FITC and R-DIL channels,

respectively. Original magnification was 400X.