Biotechnology and bioengineering, 115(3): 536-544...

43
Posted at the Institutional Resources for Unique Collection and Academic Archives at Tokyo Dental College, Available from http://ir.tdc.ac.jp/ Title Preferential capture of EpCAM-expressing extracellular vesicles on solid surfaces coated with an aptamer-conjugated zwitterionic polymer. Author(s) Alternative Yoshida, M; Hibino, K; Yamamoto, S; Matsumura, S; Yajima, Y; Shiba, K Journal Biotechnology and bioengineering, 115(3): 536-544 URL http://hdl.handle.net/10130/4837 Right This is the pre-peer reviewed version of the following article: Biotechnol Bioeng. 2018 Mar;115(3):536-544, which has been published in final form at https://doi.org/10.1002/bit.26489. This article may be used for non-commercial purposes in accordance with Wiley Terms and Conditions for Use of Self-Archived Versions. Description

Transcript of Biotechnology and bioengineering, 115(3): 536-544...

Page 1: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

Posted at the Institutional Resources for Unique Collection and Academic Archives at Tokyo Dental College,

Available from http://ir.tdc.ac.jp/

Title

Preferential capture of EpCAM-expressing

extracellular vesicles on solid surfaces coated

with an aptamer-conjugated zwitterionic polymer.

Author(s)

Alternative

Yoshida, M; Hibino, K; Yamamoto, S; Matsumura, S;

Yajima, Y; Shiba, K

Journal Biotechnology and bioengineering, 115(3): 536-544

URL http://hdl.handle.net/10130/4837

Right

This is the pre-peer reviewed version of the

following article: Biotechnol Bioeng. 2018

Mar;115(3):536-544, which has been published in

final form at https://doi.org/10.1002/bit.26489.

This article may be used for non-commercial

purposes in accordance with Wiley Terms and

Conditions for Use of Self-Archived Versions.

Description

Page 2: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

Original Article

Preferential capture of EpCAM-expressing extracellular vesicles on solid surfaces

coated with an aptamer-conjugated zwitterionic polymer.

Mitsutaka Yoshida1, 2, 3

, Kazuhiro Hibino1, Satoshi Yamamoto

1, 2, 3, Sachiko Matsumura

1,

Yasutomo Yajima2, 3

, and Kiyotaka Shiba1, 3

1Division of Protein Engineering, Cancer Institute, Japanese Foundation for Cancer Research, Koto,

Tokyo Japan

2Department of Oral & Maxillofacial Implantology, and

3Division of Oral Implants Research and

Oral Health Science Center, Tokyo Dental College, Tokyo Japan

Correspondence and requests for materials should be addressed to K.S. (email: [email protected])

Page 3: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

1

ABSTRACT

Extracellular vesicles (EVs) collectively represent small vesicles that are secreted from

cells and carry biomolecules (e.g., miRNA, lncRNA, mRNA, proteins, lipids, metabolites,

etc.) that originate in those cells. Body fluids, such as blood and saliva, include large

numbers of EVs, making them potentially a rich source of diagnostic information.

However, these EVs are mixtures of vesicles released from diseased tissues as well as

from normal cells. This heterogeneous nature therefore blurs the clinical information

obtainable from EV-based diagnosis. Here, we synthesized an EpCAM-affinity coating

agent, which consists of a peptide aptamer for EpCAM and a zwitterionic MPC polymer,

and have shown that this conjugate endowed the surfaces of inorganic materials with the

preferential affinity to EpCAM-expressing EVs. This coating agent, designated as

EpiVeta, could be useful as a coating for various diagnostic devices to allow

concentration of cancer-related EVs from heterogeneous EV mixtures.

Page 4: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

2

Introduction

The term extracellular vesicles (EVs) has been coined to describe small vesicles found

in extracellular spaces (Lotvall et al. 2014) and are believed to be secreted or broken off

from cells. Depending on the possible mechanisms of their genesis, EVs are given more

specific names, including exosomes and microvesicles, among others (Gould and

Raposo 2013). EVs carry their parental cell’s miRNA (Valadi et al. 2007), mRNA

(Valadi et al. 2007), lncRNA (Sato-Kuwabara et al. 2015), proteins (Hegmans et al.

2004), lipids (Llorente et al. 2013), metabolites (Palomo et al. 2014) and DNA (Balaj et

al. 2011) within or on their membranous particles. They are released from almost all cell

types, and are present in considerable numbers in many body fluids, such as blood,

saliva (Marzesco et al. 2005), urine (Marzesco et al. 2005), mammary secretions (milk)

(Zonneveld et al. 2014), prostatic fluid (Ronquist and Hedstrom 1977), seminal fluid

(Marzesco et al. 2005), cerebrospinal fluid (Chiasserini et al. 2014), ascites (Press et al.

2012), pleural effusion (Roca et al. 2016), bronchoalveolar fluid (Torregrosa Paredes et

al. 2012), aqueous humor (Kang et al. 2014), and pericardial fluid (Beltrami et al. 2017)

and even exhaled breath (Sinha et al. 2013), and they may participate in various cellular

Page 5: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

3

activities, including cancer growth (Skog et al. 2008), cancer metastasis (Luga et al.

2012; Peinado et al. 2012), acquisition of drug resistance (Boelens et al. 2014),

infectious disease development (Regev-Rudzki et al. 2013), neurodegeneration (Ritchie

et al. 2013), and immune deficiency (Thery et al. 2002), among others.

The recognition of the prevalence and versatility of EVs now predicts the

emergence of novel medical fields involving EV-based diagnosis (Melo et al. 2015;

Skog et al. 2008), EV-based disease prevention (Bottero et al. 2013) and EV-based

therapeutics (Vader et al. 2014). However, body fluids contain mixtures of EVs and

analysis of these mixtures without further differentiation will therefore give results that

are the averaged features of the mixtures. Exploitation of the potential of EVs as

diagnostic and therapeutic agents will require methodologies that can differentiate EVs

into subgroups based on specific characteristics. Because of this situation, we set our

goal to establish a methodology that would enrich the cancer-related EVs from body

fluids. Here, we report the synthesis of an EpCAM-affinity coating agent, which

consists of a peptide aptamer for EpCAM (which has been used as a hallmark of

circulating tumor cells(Allard et al. 2004; Myung et al. 2010) and a zwitterionic

Page 6: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

4

poly-2-methacry loyloxyethyl phosphorylcholine(MPC) polymer. Coating of this

conjugate (designated as EpiVeta ) onto silica or polystyrene surfaces imparted an

affinity for EpCAM on the material’s surfaces. Atomic force microscopy (AFM)

analysis also revealed a preferential binding of EpCAM-positive EVs onto the surface

of an EpiVeta-coated polystyrene plate.

Materials and Methods

Synthesis of EpiVeta. The Ep114 peptide aptamer for EpCAM was conjugated to MPC

copolymer (Lipidure®-5903S, average molecular weight of 400,000, synthesized from a

mixture of 2-methacryloyloxyethyl phosphorylcholine, n-butyl methacrylate, and

2-methacryloyloxyethyl succinate, with mole ratio of 3:6:1 by NOF Corporation,

Tokyo) by a two-step reaction. First, an azido-dPEG7-amine (Quanta BioDesign, Ohio,

USA) was covalently linked to the MPC copolymer by condensation between the amino

group of the linker and the carboxyl group derived from 2-methacryloyloxyethyl

succinate unit of the polymer. This reaction was carried out with 0.5 ml of 0.5% (W/W)

Lipidure-5903S solution in ethanol, 10 equivalent amounts of azido-dPEG7-amine and

50 equivalent amounts of 4-(4,6-dimethoxy-1,3,5-triazin-2-yl)-4-methylmorpholinium

Page 7: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

5

chloride (DMT-MM) at ambient temperature. After incubation for approximately 12 h,

the mixture was dialyzed in a dialysis tube against methanol for 48 h (MWCO=8,000,

Spectra/Por, Spectrum Laboratories, Houston, USA). After dialysis, the volume of

reaction mixture was reduced to 0.5 mL by evaporation. The modified MPC copolymer

was then conjugated with the peptide [in which the extra C-terminal peptide of

dPEG2-Lys(FITC)-dPEG2-Gly(propargyl)-NH2, or

dPEG2-Ser-dPEG2-Gly(propargyl)-NH2 was appended to Ep114] in the presence of

CuSO4 and sodium ascorbate. The reaction was a click chemistry reaction (Rostovtsev

et al. 2002) between the azide group of the linker and the propargyl group at the

C-terminus of the peptides. For this purpose, the azido-dPEG7-linked MPC copolymer

was mixed with 10 equivalent amounts of peptide in 5 ml of 40% methanol. After

dissolving peptide by sonication, 0.5 M L-ascorbic acid sodium (Wako Pure Chemical

Industries, Osaka, Japan) was added until the reaction mixture turned yellow, followed

by the addition of 0.5 M CuSO4, changing the color to brown. For the click chemistry

reaction, the reaction tube was capped and heated at 50 °C for 30 min in a controlled

microwave synthesizer (Biotage Initiator+, Biotage AB, Uppsala, Sweden). The

Page 8: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

6

resultant conjugate was purified by either of the following methods. Method 1

(employed in the experiment shown in Fig. 2 A) consisted of adding the reaction

mixture to ethylenediaminetetraacetic acid (pH 8.0) to 0.5 M, and dialyzing for 1 week

in a dialysis tube with a MWCO=8,000 (Spectra/Por) against 3 L of 50% methanol

containing 0.5 M of ethylenediaminetetraacetic acid (pH 8.0). Further dialysis was

performed by changing the dialysis buffer to 50% methanol. Method 2 (employed in the

experiments shown in Fig. 2 B and 3) involved the use of 1 g of chelating fiber (Chelest

Fiber IRY-HW, Chelest, Osaka, Japan) to remove copper ions. After incubation for 1 h

at ambient temperature, the chelating fiber was removed by filtration through paper.

The conjugate was further purified and concentrated using an ultrafiltration membrane

(Amicon Ultra-15, Merck Millipore, Darmstadt, German) for 50 min and centrifuging at

4,000 × g at ambient temperature. The experiment shown in Fig. 2 A used EpiVeta with

FITC and Fig. 2 B and 3 were EpiVeta without FITC. The estimated efficiency of

peptide incorporation per a EpiVeta molecule was calculated to be 0.11 (Supporting

Information).

Preparation of EpiVeta-coated materials. Silica beads (mean diameter = 75 μm,

Page 9: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

7

30511-51, Nacalai Tesque, Kyoto, Japan) were coated with EpiVeta as follows. The

beads were first treated with oxygen plasma at 10 mA for 1 min (SEDE-GE, Meiwafosis,

Tokyo, Japan), and then rinsed once with 100 μL of methanol. The beads were

resuspended in 100 μL of EpiVeta solution, and incubated for 10 h at 4 °C with vigorous

shaking in a PetiSyzer. The suspensions were then transferred to Ultrafree Centrifugal

Filters (Low-binding hydrophilic PTFE membrane 0.45 μm UFC30LH00, Millipore

Corp, MA), and residual EpiVeta was removed by centrifugation at 16,000 x g for 30

seconds at 4 °C (Microcentrifuge 5415R, Eppendorf, Hamburg, German). The beads

were dried and resuspended in 300 μL of PBS by vigorous shaking for 10 h at 4 °C. The

beads were then recovered by centrifuging at 16,000 x g for 30 seconds at 4 °C, and

resuspended in 300 μL of PBS. Similarly, 40 μL of polystyrene bead suspension

(Copolymer Microsphere Suspensions 7545A, mean 42 μm in diameter Thermo Fisher

Scientific Inc. MA, USA) was diluted with 300 μL of ethanol in a Protein LoBIND

plastic tube (Eppendorf). The suspension was dispensed into 2 × 120 μL and the beads

were collected by centrifugation and resuspended in 300 μL of EpiVeta or

Lipidure-5903S solution. The suspensions were vigorously shaken for 1 h at 37 °C in a

Page 10: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

8

PetiSyzer and then transferred to Ultrafree Centrifugal Filters and centrifuged 16,000 x

g for 30 seconds at 4 °C (Microcentrifuge 5415R). The beads were then dried

(GCD-051X Ulvac, Miyazaki, Japan) and resuspended in 400 μL of PBS (137 mM

NaCl, 2.68 mM KCl, 8.10 mM Na2HPO4, pH 7.4 buffer ) by vigorous shaking for 1 h at

37 °C. The beads were washed three times with 400 μL of PBS, and the pellets were

resuspended in 400 μL of 1% BSA in PBS by vigorous shaking for 1 h at 37 °C. The

coated beads were incubated with EVs derived from HCT-15 (15 μL of 3.5 × 1010

particles/ mL) for 15 h at 4 °C with shaking. The beads were washed once or twice with

400 μL of PBS, suspended in SDS buffer, and subjected to SDS-PAGE for western blot

analyses.

For AFM experiments (Fig. 3), a 15 × 15 mm polystyrene plate (70090

Heat-Shrinking Pla-Plate, Tamiya, Shizuoka, Japan) was first sonicated in ethanol for 5

min. After removing residual ethanol by forced air blowing, the plate was coated with

40 μL of EpiVeta or Lipidure-5903S and incubated for 16 h at 4 °C. The residual

EpiVeta was removed with a spin-coater (MS-A100, MIKASA, Tokyo, Japan) run at

2,000 rpm for 30 sec. The plate was then air-dried for 30 min, followed by drying in

Page 11: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

9

vacuo (GCD-051X) for 2 h. The coated plate was equilibrated with 400 μL of PBS at

4°C for 16 h, followed by washing three times with 400 μL of Milli-Q water. The plate

was attached to a slide glass with glue (Quick 5, Konishi, Osaka, Japan) and air dried. A

10 μL volume of PBS was applied for 1 h at ambient temperature and removed. The

EVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 10

10 particles/ mL) or HT1080

(EpCAM-, 10 μL of 1.75× 10

11 particles/ mL) were applied for 15 h at 4 °C. After three

washes with PBS, the surface was observed by AFM as described below.

Preparations of EVs. EpCAM-positive (HT-29 and HCT-15) and negative (HEK-293T

and HT-1080) cells were obtained from American Type Culture Collection. HT-29 or

HEK-293T cells were propagated using the BelloCell cell culture system

(CES-BCB01000, Cesco Bioengineering, Taichung) as follows. First, 30 mL of cell

culture (2 × 108

cells) were inoculated into a bottle of BelloCell-500p filled with 120

mL of McCoy’s 5A medium supplemented with 10% fetal bovine serum and

penicillin-streptomycin solution (168-23191, Wako Chemicals, Tokyo, Japan). When

the concentration of the cells reached 1 × 109 cells / 500 mL (approximately 96 h), the

medium was replaced with DMEM/ F12 (B-27 supplement, 20 ng/ mL EGF, 20ng/ mL

Page 12: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

10

FGF, PS), and the cells were incubated for an additional 72 h. After the incubation, 125

mL of the supernatant was centrifuged 3,500 x g for 10 min at 4 °C in a JA-14 rotor

(Beckman Coulter, Brea) to remove the cells. The resultant supernatant was filtered

through a Stericup-GP (SCGPT02RE 250 mL 0.22 µm, Polyethersulfone, Merck

Millipore, Darmstadt) using a full Teflon diaphragm vacuum pump (FTP-18A, Iwaki,

Tokyo). The filtered supernatant was then centrifuged at 160,000 × g for 70 min at 4 °C

(L-90K and SW32Ti rotor, Beckman Coulter) to prepare the crude EVs. The sediment

was washed once with 30 mL PBS at 160,000 × g for 70 min at 4 °C, and then

resuspended in 3 mL of 2.5 M sucrose Hepes/NaOH, pH 7.2 buffer. This crude EV

fraction was layered on the top of a 0.25-2 M sucrose, 20 mM Hepes/NaOH, pH 7.2

gradient prepared in a UC tube (Beckman Coulter) with a cylindrical type density

gradient former (4023, Sanplatec, Osaka). The tube was centrifuged at 100,000 × g for

17 h (L-90K, SW32Ti). Ten × 3.3 mL of the fractions were collected from the top of the

tube and their densities were determined with a refractometer (RX-5000α, Atago,

Tokyo). Each fraction was washed by centrifuging at 160,000 × g for 2 h (L-90K,

SW32Ti) in a PC tube (Beckman Coulter) after adding 27 mL of PBS. The pellets were

Page 13: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

11

suspended in 300μL (for HT-29) or 500 (for HEK-293T) μL of PBS and transferred to

1.5 mL plastic tubes (MS-4265M, Sumitomo Bakelite, Tokyo). The EVs fractions were

quick-frozen in liquid nitrogen and stored at −70 °C. Characterizations of

density-gradient fractionated EVs are shown in Supporting Information Fig. S1. Similar

EVs were prepared from HCT-15 or HT1080 cells, except that iodixanol was used

instead of sucrose as the gradient medium (Iwai et al. 2016).

Quantification of EVs by NanoSight. The numbers of EVs were determined by

NanoSight (LM10, Malvern Instruments, Worcestershire, UK) (Oosthuyzen et al. 2013).

The instrument was calibrated with silica microspheres 0.10±0.03 μm in diameter (#

24041, Polysciences, Inc., PA, USA). Samples were diluted with PBS buffer to a

concentration of 2–10 × 108 particles/mL, and were measured for 30 seconds at ambient

temperature.

Western blotting. Ten microliters of EV samples were mixed with 8 μL of reducing

sample buffer [1 M Tris-HCl (pH 6.8), 30% glycerol, 6% SDS, 3% 2-mercaptoethanol,

and 0.005% bromophenol blue] and incubated at 95 °C (ALB-121, Sansyo, Tokyo,

Japan) for 5 min. The proteins were then separated by electrophoresis in 10 or 15%

Page 14: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

12

polyacrylamide gels (Long Life GEL 10 or 15%, Oriental Instruments, Kanagawa,

Japan) at 1,000 V, 40 mA for 40 min. The separated proteins were then transferred onto

a polyvinylidene difluoride (PVDF) membrane using the iBlot Dry Blotting System

(Invitrogen, CA, USA). Nonspecific binding sites were blocked by incubating the

membrane in Blocking One (03953-95, Nacalai Tesque, Kyoto, Japan) for 1 h, followed

by washing with TBS-T (10 mM Tris-HCl, pH7.4, 150 mM NaCl, 0.05% Tween-20)

three times for 5 min each. The membrane was incubated with anti-EpCAM/VU-1D9

(1:1,000 dilution. GTX-11294, Gene Tex, Irvine, USA) for 1 h in Can Get Signal

Solution 1 (Toyobo, Osaka, Japan), followed by incubation with the secondary antibody

(goat anti-mouse IgG-HRP conjugate, Bio-Rad; 1:2000 dilution) in Can Get Signal

Solution 2 for 30 min in darkness. After three 5-min washes with TBS-T, the signals

were detected using ECL and a ChemiDoc system (Bio-Rad, Hercules, CA, USA).

Immunofluorescence microscopy observations of EpiVeta coated beads. The

immunostaining experiments (Figure 2) were conducted by incubating an aliquot of

beads with 100 μL of 1% BSA in PBS for 60 min at ambient temperature, followed by

three washes with 500 μL of PBS. The beads were then incubated with 1 μg of

Page 15: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

13

anti-Ep114 rabbit antibody (prepared by Immuno-Biological Laboratories, Gunma,

Japan, and purified from the serum with a peptide affinity column) or normal rabbit IgG

(sc-3888, Santa Cruz Biotechnology, CA, USA) for 30 min with mild shaking on a

mixer (M-36, Taitec, Koshigaya, Japan). After three washes with 100 μL of PBS, the

rabbit IgGs were visualized by incubating 5 μL of Zenon rabbit IgG labeling reagent

(Component A) for 5 min at ambient temperature in the dark, followed by incubation

with 5 μL of Zenon blocking reagent (Component B) for 5 min at ambient temperature

(Z-25307, Life Technologies). After three washes with 500 μL of PBS, the beads were

observed with a confocal laser scanning microscope (FV1000, Olympus Corporation,

Tokyo, Japan). For the EpCAM binding assay, purified EpCAM (5 μL of 140 ng/ μL)

was incubated with the beads for 60 min at ambient temperature, and then washed three

times with 500 μL of PBS. The bound EpCAM was visualized with 1 μg of

anti-EpCAM antibody (GTX-11294, Gene Tex, Irvine, USA). For the EV binding assay,

the EVs derived from HT-29 (7.7 μL of 6.4× 1011

particles/ mL) and HEK-293T (13.9

μL of 3.5× 1011

particles/ mL) were incubated with the beads for 60 min at ambient

temperature, followed by three washes with 500 μL of PBS. The bound EVs were

Page 16: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

14

visualized and with the Zenon labeling system after adding 1 μg of anti-EpCAM

antibody or anti-CD63 antibody (ab8219, Abcam, Cambridge, UK).

AFM observations. MFP-3D (Oxford Instruments, Santa Barbara, USA) was employed

for the observations of EVs on material surfaces under aqueous conditions. Silicon

probes (BL-AC40TS, Olympus, Tokyo, Japan) were used in the tapping mode (Hardij

et al. 2013). Topographic height and phase images of three randomly selected areas

were recorded at 5 × 5 μm2, 512 × 256 pixels, at a scan rate of 0.2 Hz (1 μm/ s). The

SPIP image analyzing software (Image Metrology, Hørsholm, Denmark) was used to

calculate the numbers of particles in the images, with thresholds of 15 nm in height and

2.0 in aspect ratio.

Results and Discussion

Synthesis of EpiVeta. MPC, 2-methacryloyloxyethyl phosphorylcholine has been

designed as a membrane-mimicking polymer unit consisting of methacrylate containing

a phosphoryl-choline polar group (Ishihara et al. 1990; Kadoma et al. 1978). Until now,

various types of MPC-based derivatives have been designed and used as coating agents

that reduce non-specific binding of proteins onto material surfaces (Ishihara et al. 1991;

Page 17: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

15

Sawada et al. 2003; Yoneyama et al. 1998). These coating agents usually contain a

zwitterionic MPC unit co-polymerized at various ratios with a hydrophobic polymer

unit, such as n-butyl methacrylate (BMA), to make MPC copolymer. The hydrophobic

unit is responsible for the physical interaction with the surfaces of inorganic materials,

and MPC displaying phosphoryl-choline suppresses non-specific binding of proteins or

cells to various devices (Ishihara et al. 1998; Sawada et al. 2003; Sibarani et al. 2007;

Watanabe and Ishihara 2008; Xu et al. 2010). A unit having a reactive group, such as

p-nitrophenyloxycarbonyl poly(oxyethylene)methacrylate (MEONP) or

2-methacryloyloxyethyl succinate (HOMS), can be doped along with MPC/BMA units,

to allow conjugation of types of biomolecules, such as enzymes and antibodies, to MPC

copolymer via a reactive group (Kim et al. 2012; Sakai-Kato et al. 2004). These

biomolecule-conjugated MPC copolymers can be used for biological functionalization

of the surfaces of medical devices (Watanabe and Ishihara 2008; Xu et al. 2010). In this

study, we aimed to make a coating agent that can endow the surfaces of inorganic

materials with an affinity for EpCAM (Litvinov et al. 1994), one of epithelial markers

expressed on EVs (Taylor and Gercel-Taylor 2008). For this purpose, as shown in

Page 18: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

16

Figure 1, we conjugated a MPC copolymer consisting of a 3:6:1 (mole ratio) of

hydrophilic (MPC), hydrophobic (BMA), and reactive (HOMS) units (developed by

NOF Corporation as Lipidure®-5903S), with the peptidic aptamer, Ep114, via a short

polyethylene glycol (PEG) linker, as described in the experimental section. Ep114 is a

12-mer peptide with the sequence Lys-His-Leu-Gln-Cys-Val-Arg-Asn-Ile-Cys-Trp-Ser.

This aptamer has been created as a binder for a recombinant EpCAM molecule using an

in vitro evolution system (the details of the creation of Ep114 will be published

elsewhere), and it can discriminate EpCAM positive cell from negative cell in a

complexed media (an example was demonstrated in Supporting Information Fig. S2).

For conjugation reactions, as well as for tracing the peptide, the Ep114 peptide was

appended with a C-terminal extension, dPEG2-Lys(FITC)-dPEG2-Gly(propargyl)-NH2,

consisting of a fluorescent tag and a propargyl moiety to serve as a reactive group for

the Huisgen reaction(Rostovtsev et al. 2002). Other studies have revealed that flexibility

of the displayed ligands would improve their functionality(Lahiri et al. 1999), so we

also introduced a polyethylene linker. We designated the resulting conjugate as

EpiVeta.

Page 19: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

17

Affinity coating of silica and polystyrene beads with EpiVeta. We demonstrated the

capacity of EpiVeta to coat the surfaces of inorganic materials and endow them with an

affinity for EpCAM by coating commercially available silica beads with EpiVeta, as

described in the Methods, and then characterizing the surfaces by immunostaining

experiments (Fig. 2A). Incubation of the EpiVeta-coated beads with anti-Ep114

antibody, raised by immunizing a rabbit by Ep114 peptide (our unpublished results),

revealed retention of the antibodies after washing, whereas control antibodies did not

adhere (Fig.2A a – f), indicating that Ep114 moieties are immobilized and displayed on

the surface of the beads and recognized by the anti-Ep114 antibody. Furthermore, the

coated beads retained purified EpCAM after washing, demonstrating that the beads

have an affinity for EpCAM (Fig. 2A g - i). The beads were then incubated with

EpCAM-positive and EpCAM-negative EVs isolated from HT-29 and HEK-293T cells,

respectively. Beads incubated with EpCAM-positive EVs bound the EVs, as confirmed

by binding of the anti-EpCAM antibody (Fig. 2A j - l). By contrast, beads incubated

with EpCAM-negative EVs generated very little binding signal (Fig. 2A m - o).

Confirmation that the observed binding was from EpCAM expressed on EVs and not

Page 20: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

18

free EpCAM molecules in solution was obtained by reacting the EV-incubated beads

with anti-CD63 antibody (as both EpCAM positive and negative EVs express CD63).

Stronger CD63 signals were observed when the beads were incubated with

EpCAM-positive EVs (Fig. 2A p - u), supporting the binding of the EpCAM-positive

EVs to the EpiVeta-coated beads. These experiments were semi-quantitative (the

washing steps after antibody incubation in the EV binding experiments hindered the

quantitative assessments), but the data obtained here confirmed the binding ability of

EpiVeta-coated beads for Ep114. Fluorescence microscopy observations also showed

that the EpiVeta polymer evenly coated the beads (Fig 2A a, d, g, j, m, p and s).

The specificity of the interaction between EpCAM positive EVs and the

EpiVeta-coated surface was confirmed by coating polystyrene beads with EpiVeta or

MPC copolymer, as described in Methods. These beads were then incubated with

EpCAM positive EVs and washed (once or twice), and then the amounts of bound EVs

were evaluated by western blotting using the anti-EpCAM antibody. As shown in

Figure 2B, EpiVeta-coated polystyrene beads retained substantial amounts of EpCAM

after two washings, whereas no bound EpCAM was detected on the MPC-coated beads,

Page 21: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

19

even after only one washing. This result confirmed that the EpiVeta-coating was

responsible for the specific interaction of EpCAM-positive EVs with the beads.

AFM evaluation of the interaction between EpiVeta-coated polystyrene and EVs.

The interaction between EpCAM-positive EVs and the EpiVeta-coated surface was

assessed more quantitatively using AFM to count the numbers of bound EVs. In these

experiments, polystyrene plates were coated with EpiVeta or MPC copolymer, as

described in the Methods, followed by incubation with EpCAM-positive or

EpCAM-negative EVs. After washing, the numbers of bound particles were determined

by AFM measurements. The adhesive nature of the material surfaces for biological

molecules allowed substantial numbers of EVs to be observed on non-coated (i.e.,

pristine) polystyrene plates, irrespective of the expression of EpCAM molecules on EVs

(Fig. 3 b and c). This non-specific binding of EVs was markedly reduced by the MPC

coating, indicating that that the zwitterionic polymer can suppress the non-specific

binding of EVs onto inorganic materials, as shown in the case of cells (Holmlin et al.

2001) (Fig. 3 e and f). Coating of the polystyrene plates with EpiVeta resulted in

binding of the EpCAM-positive EVs onto the surface (Fig. 3 h). By contrast,

Page 22: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

20

EpCAM-negative EVs did not bind to the surface (Fig. 3 i), validating the EpiVeta

coating as an EpCAM-specific affinity surface. Table 1 summarizes the quantitative

analyses of these AFM observations.

Affinity surfaces have been created on inorganic materials using antibodies conjugated

to self-assembling monolayers (SAMs) (Chen et al. 2003; Herrwerth et al. 2003) as well

as to polymers (Kim et al. 2012). The advantages of peptide aptamers over antibodies as

the source of affinity molecules include the ease of preparation and the versatility in

conjugation to foreign molecules (Shiba 2010). Current technologies allow the chemical

synthesis of large quantities of peptides of 100-residue sizes in a far shorter time than is

needed to obtain antibodies (which must be prepared in biological systems).

Furthermore, various modifications, including conjugations with foreign molecules, can

be incorporated during or after peptide synthesis. If necessary, the affinity of a peptide

can be decreased by rational alteration of the amino acid residues, or it can be increased

by secondary in vitro evolution (Smith and Yu 1996), a process that is generally

difficult to conduct with antibodies. A general disadvantage of peptide aptamers is the

weakness of the affinity when compared to antibodies. However, the “weak but specific”

Page 23: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

21

interactions of peptidic aptamers is sufficient for fabricating devices that are required to

continuously separate biomolecules depending on their interactions with material

surfaces, as has been successfully exploited in various types of separation columns. The

aim of this study was to establish a platform system for “a programmable bio-surface”

that would capitalize on the “weak but specific” interactions of peptidic aptamers

toward certain molecules to create affinity surfaces on various materials (Figure S3).

The ability to create peptidic aptamers de novo for specific target molecules by

employing in vitro evolution systems means that bio-surfaces can be created with a

specific affinity for a certain molecule using this method, giving this system a

“programmable” nature.

Acknowledgments

We thank T. Minamisawa, K. Suga, and K. Iwai for the preparations and

characterizations of EVs, and NOF Corporation for providing the MPC polymers.

Author Contributions Statement

Page 24: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

22

M. Y. and K. S. wrote the main manuscript text and K. S. and Y. Y. designed overall

experiments. K. H. and S. M. synthesized EpiVeta and M. Y. and S. Y. analyzed the

interaction between EpiVeta and EVs. This work was supported by ‘‘Project for Private

Universities: matching fund subsidy’’ from Ministry of Education, Culture, Sports,

Science and Technology to Y. Y. and a grant from Vehicle Racing Commemorative

Foundation to K. S.

Competing financial interest

K. S. received a joint research funding from NOF Corporation.

Page 25: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

23

References

Allard WJ, Matera J, Miller MC, Repollet M, Connelly MC, Rao C, Tibbe AG, Uhr JW, Terstappen LW.

2004. Tumor cells circulate in the peripheral blood of all major carcinomas but not in healthy

subjects or patients with nonmalignant diseases. Clin Cancer Res 10(20):6897-904.

Balaj L, Lessard R, Dai L, Cho YJ, Pomeroy SL, Breakefield XO, Skog J. 2011. Tumour microvesicles

contain retrotransposon elements and amplified oncogene sequences. Nat Commun 2:180.

Beltrami C, Besnier M, Shantikumar S, Shearn AI, Rajakaruna C, Laftah A, Sessa F, Spinetti G, Petretto

E, Angelini GD and others. 2017. Human Pericardial Fluid Contains Exosomes Enriched with

Cardiovascular-Expressed MicroRNAs and Promotes Therapeutic Angiogenesis. Mol Ther

25(3):679-693.

Boelens MC, Wu TJ, Nabet BY, Xu B, Qiu Y, Yoon T, Azzam DJ, Twyman-Saint Victor C, Wiemann BZ,

Ishwaran H and others. 2014. Exosome transfer from stromal to breast cancer cells regulates

therapy resistance pathways. Cell 159(3):499-513.

Bottero D, Gaillard ME, Errea A, Moreno G, Zurita E, Pianciola L, Rumbo M, Hozbor D. 2013. Outer

membrane vesicles derived from Bordetella parapertussis as an acellular vaccine against

Bordetella parapertussis and Bordetella pertussis infection. Vaccine 31(45):5262-8.

Chen S, Liu L, Zhou J, Jiang S. 2003. Controlling Antibody Orientation on Charged Self-Assembled

Monolayers. Langmuir 19(7):2859-2864.

Chiasserini D, van Weering JR, Piersma SR, Pham TV, Malekzadeh A, Teunissen CE, de Wit H, Jimenez

CR. 2014. Proteomic analysis of cerebrospinal fluid extracellular vesicles: A comprehensive

dataset. J Proteomics 106C:191-204.

Gould SJ, Raposo G. 2013. As we wait: coping with an imperfect nomenclature for extracellular vesicles.

J Extracell Vesicles 2.

Hardij J, Cecchet F, Berquand A, Gheldof D, Chatelain C, Mullier F, Chatelain B, Dogne JM. 2013.

Characterisation of tissue factor-bearing extracellular vesicles with AFM: comparison of

air-tapping-mode AFM and liquid Peak Force AFM. J Extracell Vesicles 2.

Hegmans JP, Bard MP, Hemmes A, Luider TM, Kleijmeer MJ, Prins JB, Zitvogel L, Burgers SA,

Hoogsteden HC, Lambrecht BN. 2004. Proteomic analysis of exosomes secreted by human

mesothelioma cells. Am J Pathol 164(5):1807-15.

Herrwerth S, Rosendahl T, Feng C, Fick J, Eck W, Himmelhaus M, Dahint R, Grunze M. 2003. Covalent

Coupling of Antibodies to Self-Assembled Monolayers of Carboxy-Functionalized

Page 26: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

24

Poly(ethylene glycol):  Protein Resistance and Specific Binding of Biomolecules. Langmuir

19(5):1880-1887.

Holmlin RE, Chen X, Chapman RG, Takayama S, Whitesides GM. 2001. Zwitterionic SAMs that Resist

Nonspecific Adsorption of Protein from Aqueous Buffer. Langmuir 17(9):2841-2850.

Ishihara K, Nomura H, Mihara T, Kurita K, Iwasaki Y, Nakabayashi N. 1998. Why do phospholipid

polymers reduce protein adsorption? J Biomed Mater Res 39(2):323-30.

Ishihara K, Ueda T, Nakabayashi N. 1990. Preparation of Phospholipid Polylners and Their Properties as

Polymer Hydrogel Membranes. Polym J 22(5):355-360.

Ishihara K, Ziats NP, Tierney BP, Nakabayashi N, Anderson JM. 1991. Protein adsorption from human

plasma is reduced on phospholipid polymers. J Biomed Mater Res 25(11):1397-407.

Iwai K, Minamisawa T, Suga K, Yajima Y, Shiba K. 2016. Isolation of human salivary extracellular

vesicles by iodixanol density gradient ultracentrifugation and their characterizations. J Extracell

Vesicles 5:30829.

Kadoma Y, Nakabayashi E, Masuhara J, Yamauchi E. 1978. Synthesis and hemolysis test of the polymer

containing phosphorylcholine groups. Kobunshi Ronbunshu 35(7):423-427.

Kang GY, Bang JY, Choi AJ, Yoon J, Lee WC, Choi S, Yoon S, Kim HC, Baek JH, Park HS and others.

2014. Exosomal proteins in the aqueous humor as novel biomarkers in patients with neovascular

age-related macular degeneration. J Proteome Res 13(2):581-95.

Kim G, Yoo CE, Kim M, Kang HJ, Park D, Lee M, Huh N. 2012. Noble Polymeric Surface Conjugated

with Zwitterionic Moieties and Antibodies for the Isolation of Exosomes from Human Serum.

Bioconjugate chemistry 23(10):2114-20.

Lahiri J, Isaacs L, Tien J, Whitesides GM. 1999. A strategy for the generation of surfaces presenting

ligands for studies of binding based on an active ester as a common reactive intermediate: a

surface plasmon resonance study. Anal Chem 71(4):777-90.

Litvinov SV, Velders MP, Bakker HA, Fleuren GJ, Warnaar SO. 1994. Ep-CAM: a human epithelial

antigen is a homophilic cell-cell adhesion molecule. The Journal of cell biology 125(2):437-46.

Llorente A, Skotland T, Sylvanne T, Kauhanen D, Rog T, Orlowski A, Vattulainen I, Ekroos K, Sandvig K.

2013. Molecular Lipidomics of Exosomes Released by PC-3 Prostate Cancer Cells. Biochim

Biophys Acta 1831(7):1302-9.

Lotvall J, Hill AF, Hochberg F, Buzas EI, Di Vizio D, Gardiner C, Gho YS, Kurochkin IV, Mathivanan S,

Quesenberry P and others. 2014. Minimal experimental requirements for definition of

extracellular vesicles and their functions: a position statement from the International Society for

Extracellular Vesicles. J Extracell Vesicles 3:26913.

Page 27: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

25

Luga V, Zhang L, Viloria-Petit AM, Ogunjimi AA, Inanlou MR, Chiu E, Buchanan M, Hosein AN, Basik

M, Wrana JL. 2012. Exosomes mediate stromal mobilization of autocrine Wnt-PCP signaling in

breast cancer cell migration. Cell 151(7):1542-56.

Marzesco AM, Janich P, Wilsch-Brauninger M, Dubreuil V, Langenfeld K, Corbeil D, Huttner WB. 2005.

Release of extracellular membrane particles carrying the stem cell marker prominin-1 (CD133)

from neural progenitors and other epithelial cells. J Cell Sci 118(Pt 13):2849-58.

Melo SA, Luecke LB, Kahlert C, Fernandez AF, Gammon ST, Kaye J, LeBleu VS, Mittendorf EA, Weitz

J, Rahbari N and others. 2015. Glypican-1 identifies cancer exosomes and detects early

pancreatic cancer. Nature 523(7559):177-82.

Myung JH, Launiere CA, Eddington DT, Hong S. 2010. Enhanced Tumor Cell Isolation by a Biomimetic

Combination of E-selectin and anti-EpCAM: Implications for the Effective Separation of

Circulating Tumor Cells (CTCs). Langmuir.

Oosthuyzen W, Sime NE, Ivy JR, Turtle EJ, Street JM, Pound J, Bath LE, Webb DJ, Gregory CD, Bailey

MA and others. 2013. Quantification of human urinary exosomes by nanoparticle tracking

analysis. J Physiol 591:5833-42.

Palomo L, Casal E, Royo F, Cabrera D, van-Liempd S, Falcon-Perez JM. 2014. Considerations for

applying metabolomics to the analysis of extracellular vesicles. Front Immunol 5:651.

Peinado H, Aleckovic M, Lavotshkin S, Matei I, Costa-Silva B, Moreno-Bueno G, Hergueta-Redondo M,

Williams C, Garcia-Santos G, Ghajar CM and others. 2012. Melanoma exosomes educate bone

marrow progenitor cells toward a pro-metastatic phenotype through MET. Nature medicine

18:883–891.

Press JZ, Reyes M, Pitteri SJ, Pennil C, Garcia R, Goff BA, Hanash SM, Swisher EM. 2012.

Microparticles From Ovarian Carcinomas Are Shed Into Ascites and Promote Cell Migration. Int

J Gynecol Cancer 22(4):546-52.

Regev-Rudzki N, Wilson DW, Carvalho TG, Sisquella X, Coleman BM, Rug M, Bursac D, Angrisano F,

Gee M, Hill AF and others. 2013. Cell-Cell Communication between Malaria-Infected Red

Blood Cells via Exosome-like Vesicles. Cell 153:1120-1133.

Ritchie AJ, Crawford DM, Ferguson DJ, Burthem J, Roberts DJ. 2013. Normal prion protein is expressed

on exosomes isolated from human plasma. Br J Haematol 163(5):678-80.

Roca E, Lacroix R, Judicone C, Laroumagne S, Robert S, Cointe S, Muller A, Kaspi E, Roll P, Brisson

AR and others. 2016. Detection of EpCAM-positive microparticles in pleural fluid: A new

approach to mini-invasively identify patients with malignant pleural effusions. Oncotarget

7(3):3357-66.

Page 28: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

26

Ronquist G, Hedstrom M. 1977. Restoration of detergent-inactivated adenosine triphosphatase activity of

human prostatic fluid with concanavalin A. Biochim Biophys Acta 483(2):483-6.

Rostovtsev VV, Green LG, Fokin VV, Sharpless KB. 2002. A Stepwise Huisgen Cycloaddition Process:

Copper(I)-Catalyzed Regioselective “Ligation” of Azides and Terminal Alkynes. Angew. Chem.

Int. Ed. 41(14):2596-2599.

Sakai-Kato K, Kato M, Ishihara K, Toyo'oka T. 2004. An enzyme-immobilization method for integration

of biofunctions on a microchip using a water-soluble amphiphilic phospholipid polymer having a

reacting group. Lab on a chip 4(1):4-6.

Sato-Kuwabara Y, Melo SA, Soares FA, Calin GA. 2015. The fusion of two worlds: Non-coding RNAs

and extracellular vesicles - diagnostic and therapeutic implications (Review). Int J Oncol

46(1):17-27.

Sawada S, Sakaki S, Iwasaki Y, Nakabayashi N, Ishihara K. 2003. Suppression of the inflammatory

response from adherent cells on phospholipid polymers. J Biomed Mater Res A 64(3):411-6.

Shiba K. 2010. Exploitation of peptide motif sequences and their use in nanobiotechnology. Curr Opin

Biotechnol 21(4):412-25.

Sibarani J, Takai M, Ishihara K. 2007. Surface modification on microfluidic devices with

2-methacryloyloxyethyl phosphorylcholine polymers for reducing unfavorable protein

adsorption. Colloids and surfaces. B, Biointerfaces 54(1):88-93.

Sinha A, Yadav AK, Chakraborty S, Kabra SK, Lodha R, Kumar M, Kulshreshtha A, Sethi T, Pandey R,

Malik G and others. 2013. Exosome-enclosed microRNAs in exhaled breath hold potential for

biomarker discovery in patients with pulmonary diseases. J Allergy Clin Immunol

132(1):219-22.

Skog J, Wurdinger T, van Rijn S, Meijer DH, Gainche L, Sena-Esteves M, Curry WT, Jr., Carter BS,

Krichevsky AM, Breakefield XO. 2008. Glioblastoma microvesicles transport RNA and proteins

that promote tumour growth and provide diagnostic biomarkers. Nature cell biology

10(12):1470-6.

Smith GP, Yu JN. 1996. In search of dark horses: Affinity maturation of phage-displaced ligands. Mol

Divers 2(1-2):2-4.

Taylor DD, Gercel-Taylor C. 2008. MicroRNA signatures of tumor-derived exosomes as diagnostic

biomarkers of ovarian cancer. Gynecol Oncol 110(1):13-21.

Thery C, Zitvogel L, Amigorena S. 2002. Exosomes: composition, biogenesis and function. Nature

reviews. Immunology 2(8):569-79.

Torregrosa Paredes P, Esser J, Admyre C, Nord M, Rahman QK, Lukic A, Radmark O, Gronneberg R,

Page 29: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

27

Grunewald J, Eklund A and others. 2012. Bronchoalveolar lavage fluid exosomes contribute to

cytokine and leukotriene production in allergic asthma. Allergy 67(7):911-9.

Vader P, Breakefield XO, Wood MJ. 2014. Extracellular vesicles: emerging targets for cancer therapy.

Trends Mol Med 20(7):385-393.

Valadi H, Ekstrom K, Bossios A, Sjostrand M, Lee JJ, Lotvall JO. 2007. Exosome-mediated transfer of

mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat Cell Biol

9(6):654-9.

Watanabe J, Ishihara K. 2008. Establishing ultimate biointerfaces covered with phosphorylcholine groups.

Colloids and surfaces. B, Biointerfaces 65(2):155-65.

Xu Y, Takai M, Ishihara K. 2010. Phospholipid Polymer Biointerfaces for Lab-on-a-Chip Devices. Ann

Biomed Eng 38(6):1938-53.

Yoneyama T, Ishihara K, Nakabayashi N, Ito M, Mishima Y. 1998. Short-term in vivo evaluation of

small-diameter vascular prosthesis composed of segmented

poly(etherurethane)/2-methacryloyloxyethyl phosphorylcholine polymer blend. J Biomed Mater

Res 43(1):15-20.

Zonneveld MI, Brisson AR, van Herwijnen MJ, Tan S, van de Lest CH, Redegeld FA, Garssen J, Wauben

MH, Nolte-'t Hoen EN. 2014. Recovery of extracellular vesicles from human breast milk is

influenced by sample collection and vesicle isolation procedures. J Extracell Vesicles 3:24215.

Page 30: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

28

Figure legends

Figure 1. Scheme for the synthesis of EpiVeta. MPC-polymer (Lipidure-5903S),

consisting of random zwitterionic (2-methacryloyloxyethyl phosphorylcholine, MPC),

hydrophobic (n-butyl methacrylate, BMA), and reactive (2-methacryloyloxyethyl

succinate, HOMS) units with the ratio of 3:6:1 (average molecular weight of 400,000),

was first ligated with a polyethylene glycol (PEG) linker, azido-dPEG7-amine, by

condensation between the amino group of the linker and the carboxyl group of the

HOMS unit. A peptide with the sequence

Lys-His-Leu-Gln-Cys-Val-Arg-Asn-Ile-Cys-Trp-Ser-dPEG2-Lys(FITC)-dPEG-Gly(pro

pargyl)-NH2 or

Lys-His-Leu-Gln-Cys-Val-Arg-Asn-Ile-Cys-Trp-Ser-dPEG2-Ser-dPEG-Gly(propargyl)-

NH2 at the C-terminus was then conjugated to the modified MPC copolymer by a click

chemistry reaction(Rostovtsev et al. 2002) to synthesize EpiVeta.

Figure 2. Immunological characterizations of EpiVeta-coated beads. (A) Silica

beads (a bright field microscopy image is shown on the left with bar of 100 μm) were

coated with EpiVeta, and their surfaces were observed by confocal laser scanning

Page 31: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

29

fluorescence microscopy, after reacting with antibodies or extracellular vesicles (EVs).

Reconstituted 3D images from the fluorescein isothiocyanate (FITC) channel

(representing EpiVeta), the Alexa 594 channel (representing antibodies), and their

merged images are shown. The beads were incubated with Anti-Ep114 antibody (a-c),

normal control antibodies (d-f), purified EpCAM + anti-EpCAM antibody (g-i),

EpCAM+ EVs + anti-EpCAM antibody (j-l), EpCAM

- EVs plus anti-EpCAM antibody

(m-o), EpCAM+ EVs plus anti-CD63 antibody (p-r) and EpCAM

- EVs plus anti-CD63

antibody (p-r). EVs of EpCAM+ and EpCAM

- were obtained from HT-29 and

HEK-293T cells, respectively. (B) Polystyrene beads were coated with EpiVeta (lanes

1-3) or MPC copolymer (lanes 4-6) and incubated with EpCAM+ EVs. After no

washing (lanes 1 and 4), or washing once (lanes 2 and 5) or twice (lanes 3 and 6), the

amounts of EpCAM (EpCAM+ EVs) molecules retained on the beads were evaluated by

western blotting. HCT-15 EVs were used for EpCAM+ EVs.

Figure 3. AFM observations of EVs on EpiVeta-coated polystyrene surfaces. The

panels show 3D images of the surfaces of non-coated (upper), MPC copolymer coated

(middle), and EpiVeta-coated (lower) polystyrene plates. The left, middle, and right

Page 32: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

30

panels show the results obtained from the surfaces incubated with no EVs or with

EpCAM+ EVs (HCT-15 EVs) and EpCAM

- EVs (HT-1080 EVs), respectively. The

colored bar represents the height of the particles. Representative images from three data

are shown for each sample.

Table 1 Summary of numbers on particles (EVs) observed by AFM on the surfaces of

pristine, MPC copolymer-coated, and EpiVeta-coated polystyrene (PS) plates.

Page 33: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

MPC copolymer EpiVetaMPC unit Ep114

FITC

azido-PEG7-amineEp114-PEG2-K(FITC) PEG2-G(propargyl)-NH2

BMA unit

HOMS unit

azideazide

Page 34: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

anti-Ep114 Ab(Alexa 594)

cont Ab(Alexa 594)

EpCAM+EpCAM Ab(Alexa 594)

Silica beadsEpiVeta (FITC)

a b c

d e f

g h i

j k l

m n o

p q r

s t u

EpiVeta (FITC) Abs (Alexa 594) Merged

Bar: 100 µm

EpCAM+ EVs+EpCAM Ab(Alexa 594)

EpCAM- EVs+EpCAM Ab(Alexa 594)

EpCAM- EVs+CD63 Ab(Alexa 594)

EpCAM+ EVs+CD63 Ab(Alexa 594)

A

B

1 2 3 4 5 6

0 1 2 0 1 2

Input

EpCA

M

EpiVeta MPC copolymerpolymerwash

Page 35: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

no EVs EpCAM+ EpCAM- no

neEp

iVet

aM

PC c

opol

ymer

30nm

15

0

-15

-30

Figure 3

a b c

d e f

g h i

Page 36: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

Supporting Information

Preferential capture of EpCAM-expressing extracellular vesicles on solid surfaces

coated with an aptamer-conjugated zwitterionic polymer.

Mitsutaka Yoshida1, 2

, Kazuhiro Hibino1, Satoshi Yamamoto

1, 2, Sachiko Matsumura

1,

Yasutomo Yajima2, and Kiyotaka Shiba

1

1Division of Protein Engineering, Cancer Institute, Japanese Foundation for Cancer Research, Koto,

Tokyo Japan

2Department of Oral & Maxillofacial Implantology, Division of Oral Implants Research and Oral

Health Science Center, Tokyo Dental College, Tokyo Japan

Page 37: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

Supporting Materials and Methods.

Estimation of incorporation efficiency of Ep114 in EpiVeta. Numbers of Ep114

incorporated in a EpiVeta molecule were calculated as follows. First, a standard

concentration curve of fluorescein isothiocyanate was obtained by measuring

absorbance at 495 nm, which was used to estimate numbers of FITC molecules in

EpiVeta solution. Similarly, a standard concentration curve of Lipidure®-5903S was

deduced by measuring absorbance at 800 nm, in which observed density, average

molecular weight, i.e., 400,000 and molecular ratio of HOMS, i.e., 0.1 of

Lipidure®

-5903S were used for calculation. The effeminacy of peptide incorporation of

0.11 was obtained from measuring absorbance at 495 nm and 800 nm of EpiVeta

solution.

Page 38: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

Supporting Results.

Figure S1. Characterizations of density-gradient fractionated extracellular vesicles

(EVs). EVs from the conditioned media from HT-29 and HEK-293T were separated

into 10 fractions by sucrose gradient centrifugation, as described in the Methods. A.

Fractions indicated in the left from HT-29 were observed by atomic force microscope

Page 39: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

(AFM, MFP-3D, Oxford Instruments, Santa Barbara, USA) in a buffer. Bars represent

500 nm. B. Results obtained from nanoparticle tracking analysis (NTA) analyses from

HT-29. The abscissa axis indicates diameters of particles and the vertical axis represent

the numbers of particles calculated. Numbers in the figures show the calculated

concentrations of total particles in the fraction. C. Western blotting and silver staining

(by SilverQuest, Invitrogen, CA, USA) of each fraction from HT-29 and HEK-293T

cells. Sucrose densities in each fraction are shown schematically at the top. Numbers on

the right indicate the molecular weight × 10-3

of standards. Antibodies used in these

experiments and their dilution ratios were mouse anti-EpCAM/TROP-1 (AF960, R&D

Systems, MN; 1:1,000), rabbit anti-CD44 (HPA005785, Sigma, Diegem; 1:1,000),

mouse anti-CD9 (ab124476, Abcam, Cambridge; 1:500), mouse anti-CD63 (ab8219,

Abcam; 1:1,000), mouse anti-Alix (634501, BioLegend, CA; 1:500), rabbit

anti-Annexin V (ab-48815, Abcam; 1:1600), goat anti-rabbit IgG (H+L)-HRP conjugate

(#170-6515, Bio-Rad, USA; 1:2000), and goat anti-mouse IgG (H+L)-HRP conjugate

(#170-6516, Bio-Rad; 1:2000).

Page 40: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

Figure S2. Cell staining by Ep114-FITC. A, EpCAM expressing HT-29 cells were

grown in RPMI 1640 (Thermo Fisher Scientific, Waltham, USA) supplemented with

10% of fetal bovine serum, in which Ep114-FITC was added to 0.5 µM. Observations

by fluorescence microscope (two views were shown) indicated the surface staining by

Ep114-FITC (green signals). DAPI was used to stain nuclei. Note that no washing step

was included in the experiment. B, Similar experiment was performed with

EpCAM-non-expressing cell line, HT1080.

Page 41: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

Figure S3. Surface coating of various materials with EpiVeta. A solution (50 μL) of

EpiVeta was applied to pieces of the various materials (10 x 10 mm) and incubated for

15 h at 4 °C. The residual EpiVeta was then removed on a spin-coater (MS-A100,

MIKASA, Tokyo, Japan) at 2,000 rpm for 30 sec. The coated materials were air dried

for 30 min, and then dried in vacuo (GCD-051X, Ulvac, Miyazaki, Japan) for 2 h. After

equilibration in PBS for 30 min, the Ep114 immobilized on the materials was detected

using rabbit anti-Ep114 antibody. Primary antibodies were probed with horse radish

Page 42: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

peroxidase (HRP)-conjugated secondary antibodies and detected using enhanced

chemiluminescence (ECL) and a ChemiDoc system (Bio-Rad, Hercules, CA, USA).

Control experiments were performed on non-coated (none) and poly-MPC

(Lipidure-5903S) pieces. Non-immune antibody was also used for control experiments.

All materials investigated produced fluorescence and stained with anti-Ep114 antibody

after EpiVeta coating, indicating that EpiVeta can immobilize Ep114 peptide onto the

surfaces of PDMS (polydimethylsiloxane), polystyrene (PS, styrene123-13), silicon

wafers (Si, Panasonic, Osaka, Japan), polyethylene terephthalate (PET, Acrysunday,

Kashiwa, Chiba), polypropylene (PP, PF-11 natural, Acrysunday), polycarbonate (PC,

PC3045 × 03clear), and polyvinyl chloride (PVC, Sawada Plastic, Japan). Under the

conditions used, only Teflon did not show any signal, suggesting it did not become

coated with EpiVeta (data not shown). For PDMS, the effect of a hydrophilic treatment

was also tested. For this purpose, pieces of PDMS were treated with oxygen plasma 20

Pa, 20 sccm, 75 W, for 5 sec using a reactive ion etching system (RIE-10NR, Samco

international, Kyoto, Japan). This oxygen plasma-treated PDMS is denoted as PDMS*

in the figure. In this experiment, EpiVeta having the Ep114 peptide with the C-terminal

Page 43: Biotechnology and bioengineering, 115(3): 536-544 URLir.tdc.ac.jp/irucaa/bitstream/10130/4837/1/bit.26489.pdfEVs derived from HCT-15 (EpCAM+, 10 μL of 3.5× 1010 particles/ mL) or

extension of GlyGly-Lys(FITC)-Gly-Gly(propargyl)-NH2 was used.