+ All Categories
Home > Documents > High-throughput combinatorial cell co-culture using microfluidics

High-throughput combinatorial cell co-culture using microfluidics

Date post: 09-Dec-2016
Category:
Upload: eugenia
View: 212 times
Download: 0 times
Share this document with a friend
11
ISSN 1757-9694 1757-9694(2011)3:6;1-R www.rsc.org/ibiology Volume 3 | Number 6 | June 2011 | Pages 595–708 PAPER Kumacheva et al. High-throughput combinatorial cell co-culture using microfluidics Volume 3 | Number 6 | 2011 Integrative Biology Pages 595–708 Quantitative biosciences from nano to macro Indexed in MEDLINE!
Transcript

ISSN 1757-9694

1757-9694(2011)3:6;1-R

www.rsc.org/ibiology Volume 3 | Number 6 | June 2011 | Pages 595–708

PAPERKumacheva et al.High-throughput combinatorial cell co-culture using microfl uidics

V

olu

me

3

| N

um

be

r 6

| 2

01

1

Inte

grative B

iolo

gy

Pa

ge

s 5

95

–7

08

Quantitative biosciences from nano to macro

Indexed in

MED

LINE!RSC Advances

An international journal to further the chemical sciences

RSC Advances is a new peer-reviewed journal covering all the chemical sciences, including interdisciplinary fields. Published articles will report high quality, well-conducted research that adds to the development of the field.

Go online today

Submissions now open – first issue mid-2011

An expert editorial team led by Professor Mike Ward, University of Sheffield, UK

Free access to all content throughout 2011 and 2012

Free colour, no page charges

Published online only

RSC Advances

ISSN 2046-2069

www.rsc.org/advances Volume 1 | Number 1 | 2011

www.rsc.org/advances Registered Charity Number 207890

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online / Journal Homepage / Table of Contents for this issue

This journal is c The Royal Society of Chemistry 2011 Integr. Biol., 2011, 3, 653–662 653

Cite this: Integr. Biol., 2011, 3, 653–662

High-throughput combinatorial cell co-culture using microfluidicsw

Ethan Tumarkin,aLsan Tzadu,

aElizabeth Csaszar,

bdMinseok Seo,

aHong Zhang,

a

Anna Lee,aRaheem Peerani,

abcKelly Purpura,

bcdPeter W. Zandstra*

bcdand

Eugenia Kumacheva*ab

Received 9th January 2011, Accepted 23rd March 2011

DOI: 10.1039/c1ib00002k

Co-culture strategies are foundational in cell biology. These systems, which serve as mimics of

in vivo tissue niches, are typically poorly defined in terms of cell ratios, local cues and supportive

cell–cell interactions. In the stem cell niche, the ability to screen cell–cell interactions and identify

local supportive microenvironments has a broad range of applications in transplantation, tissue

engineering and wound healing. We present a microfluidic platform for the high-throughput

generation of hydrogel microbeads for cell co-culture. Encapsulation of different cell populations

in microgels was achieved by introducing in a microfluidic device two streams of distinct cell

suspensions, emulsifying the mixed suspension, and gelling the precursor droplets. The cellular

composition in the microgels was controlled by varying the volumetric flow rates of the

corresponding streams. We demonstrate one of the applications of the microfluidic method by

co-encapsulating factor-dependent and responsive blood progenitor cell lines (MBA2 and M07e

cells, respectively) at varying ratios, and show that in-bead paracrine secretion can modulate the

viability of the factor dependent cells. Furthermore, we show the application of the method as a

tool to screen the impact of specific growth factors on a primary human heterogeneous cell

population. Co-encapsulation of IL-3 secreting MBA2 cells with umbilical cord blood cells revealed

differential sub-population responsiveness to paracrine signals (CD14+ cells were particularly

responsive to locally delivered IL-3). This microfluidic co-culture platform should enable high

throughput screening of cell co-culture conditions, leading to new strategies to manipulate cell fate.

Introduction

Local and systemic cell–cell interaction networks modulate

development and homeostasis and underlie ageing and disease

progression.1 In vivo, stem cell (SC) fate may be regulated by

tissue-specific niches such as the bone marrow, vascular, and

hepatic microenvironments.2 Stem cell fate in these niches is

regulated by chemical and biological microenvironments via

the release of cytokines, by cellular contacts via adhesion

molecules, and by interactions with components of the extra-

cellular matrix.3 One of the key functions of SC niches is to

aDepartment of Chemistry, University of Toronto,80 Saint George Street, Toronto, Ontario, M5S 3H6, Canada.E-mail: [email protected]

b Institute of Biomaterials & Biomedical Engineering, University ofToronto, 164 College Street, Toronto, Ontario, M5S 3G9, Canada

c Terrence Donnelly Centre for Cellular and Biomolecular Research,University of Toronto, 160 College Street, Toronto, Ontario,M5S 3E1, Canada

dDepartment of Chemical Engineering and Applied Chemistry,University of Toronto, 200 College Street, Toronto, Ontario,M5S 3E5, Canada

w Electronic supplementary information (ESI) available: Fig. S1–S4,Table S1–S2. See DOI: 10.1039/c1ib00002k

Insight, innovation, integration

This paper reports a microfluidic approach to the fast-

throughput generation of microenvironments for cell

co-culture. The strategy provides a straightforward opera-

tional platform for the co-encapsulation of a controlled

number of cells of different lineages in well-defined

microgel environments. Cells co-encapsulated in each indi-

vidual microgel are separated from cells encapsulated in

different microgels, which enables interrogation of interactions

between two types of cells encapsulated in the same microgel.

The microfluidic approach enables understanding of how cell

fate is affected by their interactions in microenvironments.

One of the applications of the method is demonstrated by

co-encapsulating factor-dependent and responsive blood

progenitor cell lines. We also utilized the method to screen

the impact of specific growth factors on a primary human

heterogeneous cell population.

Integrative Biology Dynamic Article Links

www.rsc.org/ibiology PAPER

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

654 Integr. Biol., 2011, 3, 653–662 This journal is c The Royal Society of Chemistry 2011

maintain tissue homeostasis through direct and indirect

cell–cell interactions. This is achieved by balancing the

proportions of quiescent and activated cells, and the signals

they produce. Stem cells may be activated by specific cues

within the niche to proliferate, self-renew, or differentiate to

regenerate damaged tissue.4 In addition to gaining funda-

mental understanding of cellular interactions, guiding cell fate

by co-culturing specific cell lineages has potential applications

in wound healing,5 tissue engineering,6 and cancer research.7

Cellular co-culture is also important in SC research both in

establishing and maintaining embryonic SC cultures from

mouse or human sources8 and in guiding SC fate. For

instance, co-culture of OP9 stromal cells (derived from

M-CSF deficient mice) with embryonic SCs leads to

primitive and definitive hematopoetic development of the

SCs in vitro, in a manner similar to that occurring in murine

ontogenesis.9

The development of efficient methods which would provide

rapid generation of massive libraries of co-cultured cells has

fundamental and practical importance. These methods should

enable (i) the controlled generation of interactive cellular

microenvironments that closely mimic the native cellular

microenvironment in architecture and chemical and mecha-

nical properties; (ii) the spatio-temporal control over the

relative number of co-cultured cells in the microenvironment;

and (iii) the ability to carry out high-throughput screening

integrated with function and relevant cell outputs. Since

interactions between cells can occur through direct contact

or through soluble factors via paracrine signalling, it is

important to maximize the interaction between neighbouring

cell types by creating microenvironments with an inter-cellular

distance on the scale of cells.

Currently, micropatterning of planar substrates does not

provide the ability to co-culture SCs in 3D environments.10

Microfabrication of scaffolds11 or the generation of sphere-

shaped cell colonies12 produce 3D microenvironments for

co-cultured cells. For example, mixing of cell-laden and cell-

free microgels at varying ratios was utilized to produce

microscale constructs with a controllable ratio of the two

microgels.13 Yet, these approaches have a limited ability to

vary the relative ratios of the co-cultured species in a

continuous high-throughput manner, which counteracts the

capability to create libraries for rapid screening of paracrine

signalling mediated cell–cell interactions.

Emulsification of single cells or cell aggregates14,15 and more

recently, microfluidic (MF) emulsification of suspensions

of cells has offered a new operational platform in cell

biology.16–18 Rapid MF generation of thousands of cell-laden

droplets with precisely controlled dimensions and compositions

provided the ability to create a large number of well-defined

3Dmicroenvironments for studies of cell growth and viability,19,20

gene expression,21 and enzymatic activity.22–24 Encapsulation

of cells in hydrogel microbeads (microgels) derived from

precursor droplets, yielded biocompatible microenvironments,

which allowed exchange of nutrients and oxygen between the

encapsulated cells and the surrounding medium.25–28 Micro-

fluidic encapsulation of cells in droplets and microgels paved

the way for high-throughput generation of combinatorial

libraries of assays,29–31 however, microenvironments for

co-cultured cells with a controlled ratio of the number of cells

of different lineages has not been demonstrated.

Herein we describe a MF strategy for rapid (40 s�1)

generation of 3D microenvironments—agarose microgels—

for cell co-culture. Agarose is a neutral polysaccharide that

undergoes gelation upon cooling.32 Agarose is generally bio-

inert, it exhibits low adsorptivity to proteins and cells, it can be

readily functionalized with cell adhesion proteins, and its

mechanical properties can be tuned by varying the agarose

concentration in the gel.33–36 We demonstrate the practical

advantage of the MF method by (i) achieving a 100%

encapsulation efficiency; (ii) by growing pluripotent cell

derived aggregates in the microgels, (iii) by controlling the

extent of paracrine signalling between M07e human mega-

karyoblastic leukemia cells and MBA2 support cells, which

were encapsulated in varying ratios, and ultimately affecting

the survival of M07e cells37 and (iv) co-encapsulating MBA2

cells in combination with heterogeneous umbilical cord blood

(UCB) cells to discover and determine the rescue effect of

localized IL-3 on the various hematopoietic sub-populations.

The impact of the MBA2 :M07e and the MBA2 :UCB case

studies demonstrates a novel approach to the controlled

generation and characterization of paracrine signalling

amongst cell-populations in an in vitro setting. In comparison

with existing methods used for cell encapsulation, this

approach allows for the continuous high-throughput

generation of microenvironments with a tuneable ratio of

encapsulated species.

Materials and methods

Generation of cell-laden agarose microgels

Two distinct cell suspensions were prepared in 2 wt%

solutions of ultra-low gelling temperature agarose (SeaPrep,

Lonza, Switzerland) in a PBS (Gibco-BRL, Rockville, MD)

buffer pre-heated to 37 1C. The concentration of cells varied in

the range from 2 � 106 to 8 � 106 cells mL�1. The suspensions

with equal concentrations of cells were supplied to two

channels of the MF T-junction droplet generator (Fig. 1a,

Fig. S1w) using two independently controlled syringe pumps

(Harvard Apparatus 33 Dual Syringe Pump, USA). The

temperature of the suspensions was maintained at 37 1C in a

temperature controlled incubator, in order to prevent agarose

gelation. The suspensions were mixed in a serpentine channel

with a length of 250 mm. Mineral oil containing 3 wt% of the

non-ionic surfactant Span 80 was supplied to the horizontal

channel of the MF device, as shown in Fig. 1a.

Droplets compartmentalizing the cells were generated at a

T-junction.38 Control over the encapsulation ratio of different

cells was achieved by tuning the ratio of flow rates of the

individual cell suspensions, while maintaining constant total

flow rate of the suspensions.

The cell-laden droplets travelled towards the outlet of the

MF device and entered the outlet tubing (HPFA tubing,

Upchurch Scientific, USA). The tubing was jacketed with a

hose connected to a water circulator cooled to 2 1C with a 1 : 4

vol glycerol : water mixture. The droplets passed through the

tubing within 5 min and were collected in a 15 mL centrifuge

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

This journal is c The Royal Society of Chemistry 2011 Integr. Biol., 2011, 3, 653–662 655

tube containing Hank’s Buffered Saline Solution (HBSS,

Gibco-BRL). The centrifuge tube was placed in an ice bath

to complete the gelation of the microgels. After the microgels

were maintained for 45 min in the bath, the suspension was

centrifuged, the microgels were separated, washed twice with

HBSS buffer, transferred into HBSS buffer containing 2% v/v

FBS and centrifuged at 1000 rpm for 5 min at 4 1C. The

centrifuged microgels were washed with HBSS buffer before

being transferred to cell culture media.

Cell culture

Cells were cultured under sterile conditions and maintained in

a 5% CO2 humidified incubator at 37 1C. Mouse embryonic

stem (mES) cells, R1 and YC5-YFP-NEO,39 were maintained

in cell culture media composed of Dulbecco’s Modified Eagle

Media (DMEM, Gibco-BRL) supplemented with 15% (v/v)

Fetal bovine serum (FBS, Gibco-BRL), 2 mM L-glutamine,

0.1 mM b-mercaptoethanol (BME, Sigma, St. Louise, MO),

0.1 mM non-essential amino acids (NEAA, Gibco-BRL),

1 mM sodium pyruvate (Gibco-BRL), 50 mg mL�1 penicillin

(Gibco-BRL), 50 mg mL�1 streptomycin (Gibco-BRL), and

500 pM Leukemia inhibitory factor (LIF, Chemicon,

Temecula, CA). Confluent dishes of mES cells were fed and

passaged on alternate days. M07e and MBA2 human mega-

karyoblastic leukemia cells were maintained in cell culture

media composed of Iscoves Modified Dulbecco’s Media

(IMDM, Gibco-BRL), 10% (v/v) FBS, 50 mg mL�1 penicillin,

50 mg mL�1 streptomycin, 2 mM L-glutamine, and 0.1 mM

BME. Cells were passaged every three days (1 : 6 split ratio)

and M07e cells were supplemented with 10 ng mL�1 of IL-3.

Embryoid body formation was induced by transfer of the

microgels laden with mES cells to the culture media comprising

DMEM supplemented with 15% (v/v) FBS, 2 mM L-glutamine,

0.1 mM BME, 0.1 mM NEAA, 50 mg mL�1 penicillin,

50 mg mL�1 streptomycin. The samples were maintained under

sterile conditions in a 5% CO2 humidified incubator at 37 1C

for 4.5 days.

Umbilical cord blood cells were obtained from consenting

donors according to procedures accepted by the ethics board

of Mt. Sinai hospital (Toronto, ON, Canada). Mononuclear

cells were processed as reported elsewhere.40 Progenitor cells

were isolated from the mononuclear cell fraction using the

EasySep human hematopoietic progenitor cell enrichment kit

(Stem Cell Technologies, Vancouver, BC, Canada), following

the manufacturer’s protocol. Lineage-depleted cells (lin-) were

cultured in serum-free conditions, as previously described40 for

six days to achieve required cell numbers.

Fabrication of microfluidic devices

Photolithographic masters were prepared from SU-8 50

photoresist (MicroChem, USA) in base-relief on silicon wafers.

The MF devices were fabricated in poly(dimethyl siloxane)

(Sylgard 184, Dow Corning, USA), using a standard soft

lithography procedure.41 After fabrication, the devices were

post-hydrophobized by maintaining the sealed MF device for

12 h in an oven at 140 1C and subsequently, by exposing it to

the vapor of 1,1,1,3,3,3-hexamethyldisilazane. To achieve

silanization, N2 gas was bubbled into a sealed 25 mL vial

containing 1,1,1,3,3,3-hexamethyldisilazane (99.9% pure,

Sigma, USA). During silanization, the MF devices were placed

on a hot plate at 80 1C for 1 h. The vial was sealed with a sleeve

style rubber stopper (Wheaton Science Products, USA).

Bubbling of N2 gas resulted in the vaporization of the

1,1,1,3,3,3-hexamethyldisilazane. Polyethylene tubing (Small

Parts, USA) was placed into the rubber stopper to supply the

1,1,1,3,3,3-hexamethyldisilazane gas into the MF device.

Characterization of cell-laden agarose microgels

Distributions of the dimensions of precursor agarose droplets

and the corresponding microgels were determined by

analyzing optical microscopy images of 300 droplets or micro-

gels using Image Pro 5.0 (Media Cybernetics, USA) software.

Fluorescence microscopy images of encapsulated fluores-

cently-labelled cells were captured using a Zeiss Microscope

(Axio Observer D1, USA) coupled with a digital camera

(Axio Cam HRm, Zeiss, USA). Ratios between different

dye-labelled encapsulated cells were determined by analyzing

images obtained by optical fluorescence microscopy. Approxi-

mately 8000 cell-laden microgels were included in the analysis,

in order to determine the optimized conditions for the

generation of cell-laden agarose microgels.

Flow Cytometry Analysis of cell-laden agarose microgels

Encapsulation ratios of R1 mES cells were monitored using a

Biosort flow cytometer (Union Biometrica, USA) with a bore

diameter of 250 mm. The cytometer was equipped with 488 and

561 nm excitation lasers. Prior to encapsulation, the cells were

labelled with CFDA SE (green) or CMTMR (red) dyes.

Microgel samples were introduced into the flow cytometer

following the manufacturer’s instructions (Union Biometrica).

Settings for sorting of cell-laden microgels

The sorting process was performed according to the manu-

facturer’s instructions (Union Biometrica). The cell-laden

microgels were gated to exclude 1.5% of cell-free microgels

and cell debris by using the gating parameters Time of Flight vs.

Extinction. Time of Flight provided the time required for an

object to pass through the instrument detector, that is, the

length of an object. Extinction reflected the internal complexity

Fig. 1 Microfluidic encapsulation of cells in agarose microgels.

(a) Schematic of MF cell co-encapsulation system. A serpentine

downstream channel with a length of 250 mm is not shown in the

figure. (b) The distribution of sizes of cell-laden droplets of 2 wt%

agarose solution (dotted curve) and of the corresponding agarose

microgel (solid curve).

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

656 Integr. Biol., 2011, 3, 653–662 This journal is c The Royal Society of Chemistry 2011

of an object similar to side scatter in conventional flow

cytometry instruments. Microgels were then sorted based on

fluorescence intensities, that is, Red Peak Height vs. Green

Peak Height. We set the parameters for the fluorescence signal

amplification as follows: Full scale: Tof 256, Ext 256, RedPH

65536, GreenPH 65536; Gains (Signal): Ext 3, Green 2, Red 2;

Trigger: Ext Thresholds: Signal 30, TOF Minimum: 10; PMT

Control: Green 500, Red 600.

Characterization of cell viability as a function of encapsulation

ratio

M07e cells labelled with CFDA SE were monitored by

fluorescence-activated cell sorting (FACS). Following 4.5 days

incubation of the cell-laden microgels in a co-culture environ-

ment, agarose microgels were digested by adding a 0.1% v/v

aqueous solution of agarase enzyme (Sigma) and maintaining

the system for 30 min at 37 1C. In the control experiments

conducted with non-encapsulated M07e cells it was established

that no significant change in cell viability occurred before and

after the digestion of agarose with agarase (82.3 � 1.5% and

81.8 � 1.1%, respectively). The released cells were washed

with HBSS buffer and centrifuged at 1000 rpm for 5 min at r.t.

to remove excess agarose. 7-Aminoactinomycin D (7-AAD)

dye (Molecular Probes, USA) was added at 1 mg ml�1 to cell

suspensions to label non-viable cells. The dispersion of

cells was transferred to an appropriate tube for FACS analysis

and the viability of M07e cells was determined as a function

of the encapsulation ratio of the MBA2 and M07e cell

populations.

Characterization of % rescue of specific lineage of UCB cells

Flow cytometry analysis was performed on cells prior to their

encapsulation in microgels, using a FACSCanto flow

cytometer (BD Biosciences, San Jose, CA, USA), to assess

for cell type frequencies. Cells were stained with the following

conjugated antibodies: CD34-APC (BD Biosciences),

CD133-PE (Miltenyi Biotec, Auburn, CA, USA), CD14-PE

(BD Biosciences), GyA-PE (BD Biosciences), CD41-PE

(BD biosciences), and CD11b-APC (BD Biosciences).

7-AAD dye was added to assess cell viability and isolate live

cells for quantification. UCB cells were then encapsulated in

microgels either in the absence of MBA2 cells (1 : 0 ratio), or in

the presence of MBA2 cells in a 1 : 1 ratio. MBA2 cells were

labelled with CFDA SE dye prior to encapsulation, in order to

distinguish MBA2 cells from UCB cells in subsequent analysis.

The cells encapsulated within the microgels were cultured in

Iscove’s Modified Dulbecco’s Medium (IMDM, Gibco) with

no addition of growth factors or serum substitutes. Following

three days of co-culture incubation, the agarose was digested,

counts of viable cells were performed, and flow cytometry

analysis was repeated. Cells labelled positive in the FITC

channel were excluded to remove all MBA2 cells from

analysis, and viable hematopoietic cell sub-populations were

quantified as

%Rescue¼ # Viable cellsð3 days post-encapsulationÞ# Viable cellsðpre-encapsulationÞ �100%:

ð1Þ

Results and discussion

A schematic of the MF method for producing 3D micro-

environments for cell co-culture is shown in Fig. 1a. Agarose

solutions carrying two distinct populations of cells were

supplied at 37 1C to the MF droplet generator (streams R

and G) at the total constant flow rate of 0.09 mL h�1. Prior to

reaching the T-junction, mixing of streams R and G was

enhanced by passing them through a serpentine channel.

A stream of mineral oil (viscosity 30 cP) was supplied to the

MF device perpendicular to the stream of the mixed cell

suspensions. At the T-junction, the shear stress imposed by

the stream of the mineral oil onto the stream of the mixed cell

suspension led to the generation of droplets of cell suspension.

During the encapsulation, no cells were released into the

continuous phase.

The dimensions of droplets were tuned from 70 to 110 mmby varying the flow rate of the continuous oil phase, Qc, from

to 1.2 to 2.5 mL h�1. The relative standard deviation in droplet

dimensions did not exceed 2.5%. At the end of the down-

stream channel the mineral oil was cooled to 2 1C. Gelation of

the agarose solution led to the transformation of droplets into

agarose particles (microgels). The microgels were collected in

15 mL centrifuge tubes containing HBSS buffer cooled to 4 1C.

Uniform and fast gelation of the droplets yielded microgels

with polydispersity below 3% (Fig. 1b). The final size of the

microgels in the HBSS buffer was approximately 10% smaller

than the diameter of the precursor droplets, owing to droplet

shrinkage upon gelation. Cell-laden microgels were centri-

fuged, washed with the HBSS buffer and transferred to the

appropriate cell media culture (described in Materials and

methods).

To achieve a 100% encapsulation efficiency of cells within

droplets, we utilized the Poisson distribution for the compart-

mentalized cells as in ref. 42

PðxÞ ¼ e�llx

x!ð2Þ

where P(x) is the fraction of droplets expected to contain x

cells, and l is the average number of cells per droplet. The

theoretical value of l was determined by the ratio between the

concentration of cells in the droplet phase and the average

volume of droplets (Fig. S2w). By setting the concentration of

cells to e.g., 8�106 cells mL�1 we used eqn (2) to establish that

for 110 mm-diameter droplets the encapsulation efficiency was

99.7%, sufficiently close to a 100% encapsulation rate.

In the encapsulation experiments, the concentration of cells

in the feed suspension varied from 2�106 to 8�106 cells mL�1.

We controlled the encapsulation efficiency by tuning the

dimensions of droplets from 70 to 110 mm. The theoretical

and experimental results for the total number of cells per

droplet—without differentiating between the different types of

cells—were in excellent agreement. For example, for

70 mm-diameter precursor droplets and cell concentration of

8�106 cells mL�1, we achieved 78.0% encapsulation efficiency

vs. the predicted value of 76.2%, demonstrating concordance

between theory and experiment. An encapsulation efficiency

of 98.5% was achieved for 110 mm-diameter droplets at a

cell concentration of 8�106 cells mL�1 (Fig. 2). At a cell

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

This journal is c The Royal Society of Chemistry 2011 Integr. Biol., 2011, 3, 653–662 657

concentration of 2�106 cells mL�1, a maximum encapsulation

efficiency of 74.6%, in agreement with a theoretical encapsula-

tion efficiency of 75.2%. We attained excellent correlation

between experimental results and theoretical values of Poisson

distribution for each discrete number of cells per droplet in

experiments conducted at a cell concentration of 4�106 and

8�106 cells mL�1 (Fig. S3 and Table S1 and S2w).Co-encapsulation of cells was tested for mES (murine

embryonic stem) cells labelled with Vybrant Cell Tracer

(CFDA SE ‘‘green’’ (Invitrogen)) and CellTracker Orange

(CMTMR ‘‘red’’ (Invitrogen)). Two suspensions, each con-

taining red or green mES cells (streams R and G, respectively),

at a concentration of cells of 8�106 cells mL�1, were supplied

to the MF device at equal volumetric flow rates QR = QG =

0.045 mL h�1. Fig. 3a shows a representative image of red and

green mES cells encapsulated in agarose microgels. Following

the encapsulation of the cell in the microgels and their

subsequent transfer to culture media, cell viability was

79.6 � 2.5% and embryoid bodies were observed 4.5 days

after encapsulation (Fig. 3b). We note that the viability of

individual mESC cells is low (typically, in the range from 20 to

50%) and is dependent on the cell line used.43

In co-encapsulation experiments, the concentration of cells

in each stream was 8�106 cells mL�1 and the total flow rate of

the R and G streams was 0.09 mL h�1. The ratios of flow rates

of the R-to-G streams were 1 : 0, 4 : 1, 1 : 1, 1 : 4, 0 : 1, where the

flow rate ratios of 1 : 0 and 0 : 1 corresponded to the generation

of droplets from a single stream (R or G, respectively), thereby

encapsulating a single cell population. We note that when

the two cell suspensions were mixed, the concentration of the

individual cell populations reduced in proportion to the

relative flow rates. The effect of dilution on the encapsulation

concentration of each cell population was described by the

Poisson distribution expected for 110 mm-diameter droplets

and the corresponding cell concentration in the mixed suspen-

sion (Fig. S4w).The encapsulation of red and green mES cells in the micro-

gels was analyzed using optical fluorescence microscopy

(Zeiss Axio Observer D1) and flow cytometry (Biosort, Union

Biometrica, 250 mm-diameter bore size) (Fig. 4a–e). The

utilization of flow cytometry provided a high-throughput

approach to analyzing cell-laden microgels. In Fig. 4 encap-

sulation of red and green mES cells at various ratios is

reflected in flow cytometry plots with populations in three

separate gating regions, which were determined by the ratio of

the red-to-green cells in the entire microgel population. As

expected, the samples comprising microgels laden solely with

green (R :G= 0 : 1) or solely with red mES cells (R :G= 1 : 0)

(Fig. 4a and e, respectively) were characterized by signals

appearing exclusively in their respective gating regions. Signals

appearing in the third region (double positive for both red and

green cells) were a direct result of a red cell and a green cell

residing in the same path length of the flow cytometer excita-

tion laser. This effect would result in the instrument detector

registering a signal comprised of both a red and green

component. Signals in the double positive region could not

be quantitatively assessed, due to the uncertainty in the

number of red or green cells which contributed to the signal.

To quantitatively compare the results of flow cytometry

experiments, only signals appearing exclusively in the red

region or exclusively in the green region were compared

(Flowjo 7.5). Signals collected exclusively in the red-gated

and the green-gated regions were used for analysis, and signals

outside of either the red or green gate were excluded from the

analysis. The microgels generated at the 1 : 4 flow rate ratio of

the R-to-G streams (Fig. 4b) contained B80% of the cell

population in the green gating region and B20% in the red-

gating region. The microgels produced at equal flow rates of

the R and G streams (Fig. 4c) had statistically equal ratios of

red-to-green cells of 47.7% and 52.3%, respectively, appearing

in the pure red and pure green regions. The microgels formed

at the R-to-G flow rate ratios of 4 : 1 (Fig. 4d) showed

an encapsulation trend that was inverse to that shown in

Fig. 4b, with populations of red and green cells of 79.7%

Fig. 2 Encapsulation efficiency at varying cell concentration and

droplet diameter. Experimental (blue) and theoretical (grey) fractions

of droplets encapsulating at least one cell are plotted for 70 and

110 mm-diameter droplets at cell concentrations of 2�106 and

8�106 cells mL�1. The theoretical fraction of cell-laden droplets was

calculated using eqn (2).

Fig. 3 Cell-laden agarose microgels. (a) Fluorescence optical micro-

scopy image of 100 mm-diameter agarose microgels encapsulating mES

cells. Encapsulation occurred at a total cell concentration of

8�106 cells mL�1 in the droplet phase. The cells appearing to reside

outside the microgel are out of the focus plane. (b) Optical microscopy

image of embryoid bodies formed from mES cells encapsulated in

agarose microgels after 4.5 days after encapsulation. The flow rate of

the oil and aqueous phases were 1.2 and 0.09 mL h�1, respectively.

Scale bar is 100 mm.

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

658 Integr. Biol., 2011, 3, 653–662 This journal is c The Royal Society of Chemistry 2011

and 20.3%, respectively. The results of the analysis of cell

co-encapsulation by flow cytometry and optical microscopy

images were in excellent agreement (Fig. 4f).

One utility of the MF cell co-culture method was demon-

strated by uncovering cellular interaction networks. We

encapsulated in the microgels in different ratios the MBA2

cell line (secreting IL-3) and the IL-3-dependent M07e human

megakaryoblastic leukemia cell line.37 The M07e cells are a

sub-line of the M07 human megakaryoblastic leukemia cell

lines that require interleukin-3 (IL-3) for proliferation and

survival. MBA2, a cell line transfected with the P210 gene,

secretes IL-3. We hypothesized that the survival and prolifera-

tion of encapsulated M07e cells could be controlled by varying

the relative number of MBA2 cells in the microgels (Fig. 5a).

The IL-3 secreted by the MBA2 cells would be used in both an

autocrine and paracrine fashion.

Experiments were carried out by co-encapsulating M07e

and MBA2 cells in the microgels at the flow rate ratios of the

corresponding suspensions of 0 : 1, 1 : 9, 1 : 4, 1 : 1, and 4 : 1 and

by determining the viability of M07e cells after 4.5 days

(Fig. 5b). As expected, low survival of M07e was observed

in the presence of a small relative fraction of MBA2 cells. With

increasing fractions of MBA2 cells in the microgels, M07e

viability increased, reaching a maximum of 65.9 � 8.4% at the

ratio of the encapsulated cell lines of 1 : 1. To verify that higher

M07e viability results from the higher relative number of

MBA2 cells secreting IL-3, we conducted control experiments

with non-encapsulated M07e cells in the presence of exogenous

IL-3. At the IL-3 concentration of 10 ng mL�1, the viability of

M07e cells was 76.3 � 3.5% after 4.5 days. A modestly higher

value than that achieved for a 1 : 1 ratio of encapsulated

MBA2 :M07e cells was presumably caused by the higher levels

of IL-3 in the control system.

Since M07e cells are a parental line to the MBA2 cells,

standard approaches to distinguish the two populations by

side scatter (SSC) vs. front scatter (FSC) could not be applied.

The SSC vs. FSC was utilized in distinguishing between viable

and non-viable M07e cells, based on the change in cell

Fig. 4 Co-encapsulation of mES cells in microgels achieved at varying flow rate ratios of the corresponding cell suspensions. (a)–(e) Left: Relative

fluorescence intensity plots of red and green channels for sorted microgels laden with R1 mES cells labelled with Vybrant CFDA or CellTrackert

Orange CMTMR (green and red cells, respectively). Right: Optical microscopy images of the corresponding microgels. The microgels were

produced at the respective flow rate ratios of the R and G streams: (a) QR :QG = 0 : 1, (b) QR :QG = 1 : 4, (c) QR :QG = 1 : 1. (d) QR :QG = 4 : 1,

(e) QR :QG = 1 : 0. Gating was determined by positive controls comprising cells labelled with only one dye (R :G = 1 : 0 and 0 : 1). (f) Fraction, a,of green and red cells encapsulated at different ratios of QR :QG. Light and dark green bars show the fraction of encapsulated green cells,

determined by image analysis and flow cytometry, respectively. Light and dark red bars represent the fraction of encapsulated red cells, determined

by image analysis and flow cytometry, respectively. Fluorescence intensity scale is defined by the sorting parameters (ESIw). Scale bar is 100 mm.

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

This journal is c The Royal Society of Chemistry 2011 Integr. Biol., 2011, 3, 653–662 659

morphology upon cell death. In order to differentiate between

the two cell lines, M07e cells were marked with the CFDA dye

prior to encapsulation in the microgels. The dye was retained

by the cells throughout cellular development, inherited by

daughter cells, and was not transferred to adjacent cells.44–46

Furthermore, to quantify the viability of M07e cells after their

encapsulation in microgels, 7-Aminoactinomycin D (7-AAD)

dye was added to all cell samples prior to FACS analysis.

7-AAD cannot readily pass through intact cell membranes

and is commonly used for DNA staining of dead cells. The

fluorescence intensity as a result of cellular uptake of 7-AAD

was determined by flow cytometry.

Typical histograms of the cellular uptake of 7-AAD dye by

CFDA-labelled cells at various co-culture ratios of M07e-to-

MBA2 cells are shown in Fig. 5c.

To verify that endogenously produced IL-3 resulted from

cell co-encapsulation within the microgels, rather than

diffusion of IL-3 through the medium, we carried out control

experiments. We encapsulated either MBA2, or M07e cells in

agarose microbeads and mixed cell-laden microgels in a 1 : 1

ratio in the media with varying volume. Encapsulation of each

cell line in the microgels was carried out at a cell concentration

of 8�106 cells mL�1 Fig. 6 shows that regardless of the volume

of the media, M07e cells had viability not exceeding 20%. We

attribute this effect to the limited diffusion of IL-3 from the

MBA2-laden microgels into the M07e-laden microgels, there-

by resulting in low concentrations of endogenously produced

IL-3 in the bulk media.

In order to explore another application of the proposed MF

platform, we investigated the impact of a growth factor on

Fig. 5 Variation in the viability of M07e human megakaryoblastic leukemia cells examined at varying encapsulation ratio of MBA2 to M07e

cells. (a) Schematic of co-culture of MBA2 and M07e cells in different number ratios. MBA2 cells secrete IL-3 which is required for the survival of

both cell populations. Increase in the relative number of MBA2 cells results in an increase in M07e survival rate, due to the local increase in IL-3.

(b) Viability of M07e cells plotted as a function of the number ratio of MBA2 cells-to-M07e cells encapsulated in 100 mm-diameter agarose

microgels and analyzed by flow cytometry. Asterisk denotes non-encapsulated M07e cells in the presence of exogenous IL-3. (c) Representative

flow cytometry histograms of cellular uptake of 7-AAD fluorescence intensity for varying M07e co-culture conditions 4.5 days after encapsulation.

X-axis represents fluorescence intensity of cells stained with 7-AAD. Numerical percentages represent the mean and standard error of M07e

cellular uptake of 7-AAD dye from at least three independent experiments.

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

660 Integr. Biol., 2011, 3, 653–662 This journal is c The Royal Society of Chemistry 2011

specific cell types within the primary human hematopoietic

SC culture system (Fig. 7a). Hematopoietic cell culture is

highly complex and heterogeneous. The heterogeneity is

compounded by system dynamics and paracrine signalling

cascades that act to regulate the system. A study of the role

of specific ligand-cell interactions within the context of this

system would enable a greater understanding of signalling

networks and provide a means of manipulating the culture

system, however, currently this task is biologically and tech-

nically challenging. Fig. 6 shows that cells encapsulated in

each microgel are isolated from each other. This feature

enables high throughput (parallel) interrogation of inter-

actions between two types of cells, when they are interacting

within the microgel, but are isolated from each other between

different microgels. An important aspect is the co-localization

of two types of cells within a small sample volume. In

comparison to bulk encapsulation in a macroscale agarose

gel, the probability of putting two types of cell populations in

close spatial proximity to each other would be very low and

highly variable, making it difficult to examine interactions and

their effects across a highly heterogeneous cell hierarchy in a

controlled manner.

In our work, after co-culturing the IL-3-secreting MBA2

cells with heterogeneous UCB cells, the UCB cells were

assessed to determine cell type specific effect of localized

IL-3 secretion on viability. By assessing the cell phenotype

of the viable UCB cells, the rescue (relative to control condi-

tions where the paracrine cell was not present) of specific

hematopoietic cell types was quantified. Fig. 7b shows poor

survival of individually encapsulated UCB cells when no

growth factor (IL-3) was added. Furthermore, when bulk

exogenous IL-3 was added to the culture, weak enhancement

of UCB cell survival was observed, likely due to the limited

IL-3 diffusion across the agarose microgel matrix to the UCB

cells. When the UCB cells were cultured in the presence of IL-3

secreting MBA2 cells, a dramatic increase in cell survival was

observed, indicating that the local (in-bead) secretion of IL-3

enabled survival and detection of responsive phenotypes from

a heterogeneous input cell population. Importantly, the

effectiveness of the rescue of the UCB cells varied among the

hematopoietic cell types that were present in the hetero-

geneous culture. Fig. 7b illustrates that monocytes (CD14+)

were most impacted by the co-culture, achieving 480%

rescue, which suggested the importance of IL-3 to the viability

of this cell type. Intermediate IL-3 dependent viability was

observed in the hematopoietic progenitors (CD34+, CD133+),

erythroid cells (GyA+), and granulocytes (CD11b+). Minimal

rescue (effect on viability) was observed in the megakaryocyte

population (CD41+), indicating that IL-3 did not play a role

Fig. 7 Co-encapsulation of MBA2 cells with UCB cells were used to determine the survival effect of IL-3 on sub-populations within a

hematopoietic culture system. (a) Schematic of co-culture of MBA2 cells and UCB cells. (i) Encapsulation of solely UCB cells without exogenous

IL-3 results in poor cell rescue. (ii) Addition of exogenous IL-3 (10 ng mL�1) results in low rescue of several UCB phenotypes.

(iii) Co-encapsulation of MBA2 cells and UCB cells results in low, moderate, and high rescue of specific UCB phenotypes. (b) % Rescue of specific

lineages of hematopoietic UCB cells co-encapsulated with MBA2 cells. % rescue of specific UCB phenotypes were measured 3 days after encapsulation

in a non-supportive media using eqn (1). This is a representative experiment conducted from at least two independent experiments (n = 2).

Fig. 6 Viability of encapsulated M07e cells at varying media volume.

Microgels containing only MBA2 cells and only M07e cells mixed in a

1 : 1 ratio at varying media volume. Asterisk denotes non-encapsulated

M07e cells in the presence of exogenous IL-3.

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

This journal is c The Royal Society of Chemistry 2011 Integr. Biol., 2011, 3, 653–662 661

in the survival of this cell type under these conditions. These

results show the differential effect of IL-3 among cell types

within a hematopoietic culture system and provides a high-

throughput means of screening signalling effects in stem

cell systems. This experiment also suggests that there are

qualitative differences in biological response seen by local

(cell mediated—in bead) delivery of soluble factors versus

global (media added) soluble factor delivery, in part due to

higher effective concentrations likely achievable through con-

tinuous local delivery.

Given the complex heterogeneity of hematopoietic cell

cultures, it has been difficult to discern the varying effects of

individual growth factors on specific cell types within the

heterogeneous culture. Paracrine signalling cascades and

culture dynamics further complicate traditional methods of

analysis. The co-encapsulation of the MBA2 cells and the

UCB cells allows one to determine the hematopoietic cell types

that can be rescued by the localized presence of IL-3. Thus, a

MF platform provides a strategy for intricate co-culture

studies, while minimizing the barriers of complex paracrine

interactions that are seen in bulk cultures and eliminating

difficulties in isolating individual cell types from a hetero-

geneous culture. By doing so, specific ligand-cell interactions

can be precisely interrogated and new biological hypotheses

can be developed and investigated.

Conclusions

The MF platform provides a fast, efficient, and easy to

implement method for the high-throughput generation of cell

co-culture compound libraries which, after incubation, can be

analyzed by optical microscopy and flow cytometry. The MF

strategy offers the speed and the control over the total number

of compartmentalized cells and the relative numbers of

different cell populations. Co-encapsulation and co-culture

of MBA2 and M07e cells at varying ratios demonstrated the

ability of the MF approach to modulate paracrine signalling

amongst cell populations in a well-defined microenvironment.

Furthermore, the MF method can be utilized as a tool to

investigate the impact of specific growth factors on a hetero-

geneous cell population. Co-encapsulation of MBA2 cells with

UCB cells was used to determine sub-populations of the

hematopoietic cell types that are largely IL-3 dependent. This

approach offers two important abilities: (1) the use of isolated

micro-environments provides the capability for studying

biological networks which otherwise could not be easily

isolated in a bulk environment. (2) The MF platform offers

the ability to control the ratio of the two cell populations

(a secreted factor population and a heterogeneous cell population).

As was demonstrated, the ratio of the encapsulated species is

important as it determines the dose of the signal given by the

secreting cells and received by the receptor cells (and its

consequent biological response). Our findings indicate that

direct co-culture had different biological effects from soluble

factor addition.

Further development and improvement of the MF platform

will increase the breath of its applications. The ability to screen

the phenotypic and functional effects of cell–cell interactions

in-bead, ideally with live cells, would prevent the need to

remove cells from the beads for analysis. Additionally,

co-encapsulation of various test cell populations combined

with microfluidically mediated variation in properties of the

microenvironment (i.e., growth factors, physical properties

(stiffness), or other ‘‘niche’’ parameters would add to the

range of conditions that can be queried. Given the rapid

development of microfluidic encapsulation technologies, these

next steps should emerge shortly.

Acknowledgements

The authors acknowledge financial support of this work by the

Natural Sciences and Engineering Research Council of

Canada and the Canadian Institutes for Health Research.

Biosort flow cytometry results were acquired by Dr Rock

Pulak (Union Biometrica).

References

1 D. E. Discher, D. J. Mooney and P. W. Zandstra, Science, 2009,324, 1673–1677.

2 J. Voog and D. L. Jones, Cell Stem Cell, 2010, 6, 103–115.3 R. Peerani and P. W. Zandstra, J. Clin. Invest., 2010, 120, 60–70.4 M. P. Lutolf, P. M. Gilbert and H. M. Blau, Nature, 2009, 462,433–441.

5 F. Grinnell, J. Cell Sci., 1992, 101, 1–5.6 A. F. Black, F. Berthod, N. L’Heureux, L. Germain andF. A. Auger, FASEB J., 1998, 12, 1331–1340.

7 J. L. Camps, S. M. Chang, T. C. Hsu, M. R. Freeman, S. J. Hong,H. E. Zhau, A. C. Voneschenbach and L. W. K. Chung, Proc.Natl. Acad. Sci. U. S. A., 1990, 87, 75–79.

8 H. J. Rippon and A. E. Bishop, Cell Proliferation, 2004, 37, 23–34.9 T. Nakano, H. Kodama and T. Honjo, Science, 1996, 272,722–724.

10 E. J. Lee, E. W. L. Chan and M. N. Yousaf, ChemBioChem, 2009,10, 1648–1653.

11 A. P. McGuigan and M. V. Sefton, Tissue Eng., 2007, 13,1079–1089.

12 A. Y. Hsiao, Y. S. Torisawa, Y. C. Tung, S. Sud, R. S. Taichman,K. J. Pienta and S. Takayama, Biomaterials, 2009, 30, 3020–3027.

13 Y. A. Du, E. Lo, S. Ali and A. Khademhosseini, Proc. Natl. Acad.Sci. U. S. A., 2008, 105, 9522–9527.

14 G. Karoubi, M. L. Ormiston, D. J. Stewart and D. W. Courtman,Biomaterials, 2009, 30, 5445–5455.

15 S. M. Dang, S. Gerecht-Nir, J. Chen, J. Itskovitz-Eldor andP. W. Zandstra, Stem Cells, 2004, 22, 275–282.

16 E. Tumarkin and E. Kumacheva, Chem. Soc. Rev., 2009, 38,2161–2168.

17 A. Huebner, S. Sharma, M. Srisa-Art, F. Hollfelder, J. B. Edel andA. J. deMello, Lab Chip, 2008, 8, 1244–1254.

18 M. E. Vincent, W. S. Liu, E. B. Haney and R. F. Ismagilov, Chem.Soc. Rev., 2010, 39, 974–984.

19 J. Clausell-Tormos, D. Lieber, J. C. Baret, A. El-Harrak,O. J. Miller, L. Frenz, J. Blouwolff, K. J. Humphry, S. Koster,H. Duan, C. Holtze, D. A. Weitz, A. D. Griffiths andC. A. Merten, Chem. Biol., 2008, 15, 427–437.

20 E. Brouzes, M. Medkova, N. Savenelli, D. Marran,M. Twardowski, J. B. Hutchison, J. M. Rothberg, D. R. Link,N. Perrimon and M. L. Samuels, Proc. Natl. Acad. Sci. U. S. A.,2009, 106, 14195–14200.

21 J. U. Shim, L. F. Olguin, G. Whyte, D. Scott, A. Babtie, C. Abell,W. T. S. Huck and F. Hollfelder, J. Am. Chem. Soc., 2009, 131,15251–15256.

22 A. Huebner, D. Bratton, G. Whyte, M. Yang, A. J. deMello,C. Abell and F. Hollfelder, Lab Chip, 2009, 9, 692–698.

23 A. Huebner, L. F. Olguin, D. Bratton, G. Whyte, W. T. S. Huck,A. J. deMello, J. B. Edel, C. Abell and F. Hollfelder, Anal. Chem.,2008, 80, 3890–3896.

24 S. Y. Teh, R. Lin, L. H. Hung and A. P. Lee, Lab Chip, 2008, 8,198–220.

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online

662 Integr. Biol., 2011, 3, 653–662 This journal is c The Royal Society of Chemistry 2011

25 S. Sugiura, T. Oda, Y. Izumida, Y. Aoyagi, M. Satake, A. Ochiai,N. Ohkohchi and M. Nakajima, Biomaterials, 2005, 26, 3327–3331.

26 W. H. Tan and S. Takeuchi, Adv. Mater., 2007, 19, 2696–2701.27 P. Panda, S. Ali, E. Lo, B. G. Chung, T. A. Hatton,

A. Khademhosseini and P. S. Doyle, Lab Chip, 2008, 8, 1056–1061.28 D. Chicheportiche, S. Darquy, J. Lepeintre, P. Chenard,

C. Khosrovaninejad, H. Briandet, J. Rouchette and G. Reach,Diabetologia, 1988, 31, A479–A479.

29 A. D. Griffiths and D. S. Tawfik, Trends Biotechnol., 2006, 24,395–402.

30 J. J. Agresti, E. Antipov, A. R. Abate, K. Ahn, A. C. Rowat,J. C. Baret, M. Marquez, A. M. Klibanov, A. D. Griffiths andD. A. Weitz, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 4004–4009.

31 S. Vyawahare, A. D. Griffiths and C. A. Merten, Chem. Biol., 2010,17, 1052–1065.

32 A. Hayashi and T. Kanzaki, Food Hydrocolloids, 1987, 1, 317–325.33 C. A. Agudelo, Y. Teramura and H. Iwata, Transplantation, 2009,

87, 29–34.34 D. Pelaez, C. Y. C. Huang and H. S. Cheung, Stem Cells Dev.,

2009, 18, 93–102.

35 A. Kumachev, J. Greener, E. Tumarkin, E. Eiser, P. W. Zandstraand E. Kumacheva, Biomaterials, 2011, 32, 1477–1483.

36 S. Lahooti and M. V. Sefton, Cell Transplant., 2000, 9, 785–796.37 C. Sirard, P. Laneuville and J. E. Dick, Blood, 1994, 83, 1575–1585.38 T. Thorsen, R. W. Roberts, F. H. Arnold and S. R. Quake, Phys.

Rev. Lett., 2001, 86, 4163–4166.39 A. K. Hadjantonakis and A. Nagy, Genesis, 2000, 27, 95–98.40 D. C. Kirouac, G. J. Madlambayan, M. Yu, E. A. Sykes, C. Ito

and P. W. Zandstra, Mol. Syst. Biol., 2009, 5, 293–313.41 Y. N. Xia and G. M. Whitesides, Annu. Rev. Mater. Sci., 1998, 28,

153–184.42 F. A. Haight,Handbook of the Poisson distribution, ed. F. A. Haight,

Wiley, New York, 1967.43 S. M. Dang, M. Kyba, R. Perlingeiro, G. Q. Daley and

P. W. Zandstra, Biotechnol. Bioeng., 2002, 78, 442–453.44 A. Nose and M. Takeichi, J. Cell Biol., 1986, 103, 2649–2658.45 A. B. Lyons and C. R. Parish, J. Immunol. Methods, 1994, 171,

131–137.46 P. D. Hodgkin, J. H. Lee and A. B. Lyons, J. Exp. Med., 1996, 184,

277–281.

Dow

nloa

ded

by D

alho

usie

Uni

vers

ity o

n 08

/05/

2013

21:

11:0

2.

Publ

ishe

d on

28

Apr

il 20

11 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C1I

B00

002K

View Article Online


Recommended