Analysis of Calcium Transients and Uniaxial ... - ACS Publications

Sep 19, 2016 - calcium transients and micropillar deflection induced by a single-cell uniaxial contraction force. ..... trend, we analyzed the calcium...
0 downloads 0 Views 8MB Size
Subscriber access provided by Eastern Michigan University | Bruce T. Halle Library

Article

Analysis of calcium transients and uni-axial contraction force in single human embryonic stem cell derived-cardiomyocytes on microstructured elastic substrate with spatially controlled surface chemistries Eleonora Grespan, Sebastian Martewicz, Elena Serena, Vincent Le Houérou, Jürgen Rühe, and Nicola Elvassore Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.6b03138 • Publication Date (Web): 19 Sep 2016 Downloaded from http://pubs.acs.org on September 29, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Langmuir is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 39

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

Langmuir

1

Analysis of calcium transients and uni-axial contraction force

2

in single human embryonic stem cells derived-cardiomyocytes

3

on microstructured elastic substrate with spatially controlled

4

surface chemistries

5 6

E. Grespan1, S. Martewicz2,3, E. Serena2,3, V. Le Houerou4, J. Rühe5, N.Elvassore2,3*

7

1.

CNR Institute of Neuroscience, Corso Stati Uniti 4, 35127, Padova, Italy;

8

2.

Department of Industrial Engineering (DII), University of Padova, Via Marzolo 9, 35131,

9

Padova, Italy;

10

3.

Venetian Institute of Molecular Medicine (VIMM), Via Orus 2, 35129, Padua, Italy

11

4.

Institute Charles Sadron (ICS), University of Strasbourg, 23 rue du Loess, 84047, Strasbourg,

12 13 14

France; 5.

Institute of Microsystems Engineering (IMTEK), Department of Chemistry and Physics of Interfaces, University of Freiburg, Georges-Köhler Allee 103, 79110, Freiburg, Germany

15 16 17 18 19 20 21 22 23 24 25

* Corresponding author:

26

Prof. Nicola Elvassore

27

Department of Industrial Engineering (DII),

28

University of Padova

29

Via Marzolo 9, 35129

30

Padova, Italy

31

Tel: +39-049-827-5469

32

Fax: +39-049-827-5461

33 34

Email: [email protected]

1

ACS Paragon Plus Environment

Langmuir

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

Page 2 of 39

1

Abstract

2

The mechanical activity of cardiomyocytes is the result of a process called excitation contraction

3

coupling (ECC). A membrane depolarization wave induces a transient cytosolic calcium

4

concentration increase that triggers activation calcium-sensitive contractile proteins leading to cell

5

contraction and force generation. An experimental setup capable of acquiring simultaneously all

6

ECC features would have an enormous impact on cardiac drug development and disease study.

7

In this work, we develop a micro-engineered elastomeric substrate with tailor-made surface

8

chemistry to measure simultaneously the uni-axial contraction force and the calcium transients

9

generated by single human cardiomyocytes in vitro. Micro-replication followed by photocuring is

10

used to generate an array consisting of elastomeric micropillars. A second photochemical process

11

is employed to spatially control the surface chemistry of the elastomeric pillar. As result, human

12

embryonic stem cell derived cardiomyocytes (hESC-CMs) can be confined in rectangular cell-

13

adhesive areas, which induce cell elongation and promote suspended cell anchoring between two

14

adjacent micropillars. In this end-to-end conformation, confocal fluorescence microscopy allows

15

simultaneous detection of calcium transients and micropillar deflection induced by single cell uni-

16

axial contraction force. Computational finite elements model (FEM) and 3D reconstruction of cell-

17

pillar interface allow force quantification. The platform is used to follow calcium dynamics and

18

contraction force evolution in hESC-CMs cultures over the course of several weeks. Our results

19

show how a biomaterial-based platform can be a versatile tool for in vitro assaying cardiac

20

functional properties of single-cell human cardiomyocytes, with applications in both in vitro

21

developmental studies and drug screening on cardiac cultures.

22 23 24

Keywords

25

Human embryonic stem cells derived cardiomyocytes

26

Micropillar

27

Photopattern

28

Force generation

29

Single cell

30

2

ACS Paragon Plus Environment

Page 3 of 39

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

Langmuir

1

1. Introduction

2

Cardiac output is mainly determined by the functional performance of its elementary contractile

3

unit: the single cardiomyocyte (CM) [1-3].

4

At single-cell level, the events leading to force generation are comprehensively termed excitation-

5

contraction coupling (ECC). This process involves voltage-gated ion channels opening, a rapid

6

increase in intracellular calcium concentration and actin–myosin system activation by displacement

7

of Ca2+-sensitive troponin complexes to generate cell shortening. Calcium acts as the second

8

messenger linking excitation to contraction by entering the myocyte through voltage-activated L-

9

type channels and triggering additional calcium release from intracellular stores via calcium-

10

induced calcium-release (CICR) by ryanodine receptors (RyR) on the sarcoplasmic reticulum (SR).

11

The initial conditions and mechanical relaxation are then restored by fast calcium clearance from

12

the cytosol mainly by two routes: reuptake into the SR by sarcoendoplasmic reticulum Ca2+-

13

ATPase (SERCA pumps) and extrusion through the sarcolemma by the Na+/Ca2+ exchanger

14

(NCX) [4]. Thus, the centerpiece of the entire ECC is represented by the momentary calcium

15

concentration change in the cytoplasm, called Ca2+ transient [5]. Deranged contractions of

16

cardiomyocytes can cause dysfunctions in cardiac pumping activity and cardiac failure, which is

17

currently the leading cause of mortality and morbidity in Western Countries [6]. For this reasons,

18

detailed assessment of human cardiac function at the single cardiomyocyte level can help to

19

elucidate the pathophysiological mechanisms of heart failure and facilitate development of novel

20

therapeutic interventions.

21

In recent years, human cardiomyocytes derived from pluripotent stem cells (hPSC-CMs) have

22

arisen as one of the most promising cellular models for human myocardium for in vitro studies.

23

Despite their immaturity and early phenotype, these cells are currently being employed in

24

numerous studies as models for human cardiomyocyte function. Single hPSC-CMs have been

25

widely assayed and characterized for their functional performance in electrical activity and calcium

26

handling (7, 8). To assess both features, a plethora of methods have been employed for either

27

high-content or high-throughput analysis, such as patch-clamp or micro-arrayed electrodes (MEA)

28

for action potential (10, 11) and imaging techniques for calcium cycling (12, 13, 14). Nevertheless,

29

fully comprehensive functional assays should not omit the evaluation of the end-point cardiac

30

features along with the other elements of the ECC: force generation and contraction

31

characteristics.

32

Despite several experimental methods have been employed to quantify cellular forces, such as

33

AFM [14-16], flexible cantilever [17], carbon fiber-based approaches [18], flat elastic substrates

34

(traction force microscopy) [19-23], strain-gauges [24] and arrays of microposts [25-34], none of

35

these provides a truly high-throughput method. Moreover, they are bound by experimental

36

constrains resulting in difficult comparisons with in vivo conditions especially because in most

3

ACS Paragon Plus Environment

Langmuir

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

Page 4 of 39

1

cases isolated CMs are analyzed in a flat environment. In 2D cultures, the recorded tension forces

2

are not generated by end-to-end cell configurations that characterize the functional syncytium of

3

the cardiac muscle, but rather tensions are transduced from a nearly flat cell to the specific

4

substrate of choice through adhesion of the whole cell membrane.

5

Most reported studies are based on primary animal-derived cell models. In order to overcome the

6

intrinsic differences between animal and human physiologies, a human model would be more

7

desirable. In this perspective, the generation of human pluripotent stem cell-derived

8

cardiomyocytes (hPSC-CMs) provides an excellent tool for studying in vitro cardiomyocytes of

9

human origin in physiological conditions as well as in genetically or environmentally pathological

10

ones.

11

In force-measurement studies specifically involving hPSC-CMs, mainly two approaches have been

12

used: traction force microscopy [35-37] and dense micropost arrays [38, 39]. In both cases,

13

cardiomyocytes are cultured on top of a compliant substrate and the total force produced by a

14

single cell is determined by summing the absolute magnitudes of the forces measured at each

15

point of interest, evaluated by physical displacement of microbeads embedded in an hydrogel or

16

the edge-movement of elastic microposts. These values represent tangential traction force of the

17

cardiomyocyte on a widespread flat surface, thus the resulting sums can be scarcely approximated

18

to two dipoles and are not equivalent to the gold standard measurement of axial traction [40].

19

In order to overcome the 2D limitations, engineered human cardiac tissues hanging between two

20

micropillars have been proposed as solution [41-43]. In these studies, the uni-axial traction force is

21

measured in dense cellular agglomerates without single-cell resolution.

22

A single-cell uni-axial approach has been recently proposed by using a thermoresponsive

23

sacrificial support layer in conjunction with an array of widely separated elastomeric microposts

24

[40]. Through this method, it was possible to achieve culture of immature cardiomyocytes,

25

including human embryonic stem cell derived (hESC) cardiomyocytes, which randomly anchor

26

between two or multiple pillars. However, the force was quantified only for neonatal rat

27

cardiomyocytes by phase contrast imaging, without taking into account the relevant calcium

28

dynamics data associated with each contraction and critical for the cardiac functionality

29

assessment [5].

30

In this work, we aimed at developing an in vitro platform for simultaneous acquisition of calcium

31

transients and uni-axial contraction force at a single human cardiomyocyte level. In particular, we

32

developed a photo-patterning technique on elastic micropillared substrate in order to guide the

33

cells into a physiological end-to-end cell configuration for long-term functional evaluation of

34

hESCs-CMs performances. Previous studies on cells in contact with micropillared surfaces [38-43]

35

were based on polydimethylsiloxane (PDMS). A limit of PDMS substrates is its challenging surface

36

chemistry, which hardly allows permanent chemical modification. To adapt to the specific biological

4

ACS Paragon Plus Environment

Page 5 of 39

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

Langmuir

1

aim, our platform required a novel polymer, which characteristics allow easy microstructuring and

2

stable chemical modification, in order to finely tune both topological and chemical properties of the

3

culture substrate. For substrate generation, we used a biocompatible elastomeric substrate

4

consisting of n-butylacrylate (n-BA) and methacryloxybenzophenone (MABP) as a photo-active

5

group. Microstructured surfaces composed of micropillar arrays were fabricated by replica molding

6

and photocuring. This was followed by spatially controlled photochemical polymerization of linear

7

polyacrylamide to define cell-adhesive and cell-repellent areas on the microstructured substrates

8

with micrometric resolution. Fluorescence confocal microscopy together with computational

9

mechanical modeling of three dimensional cell-pillar interface reconstructions provided

10

quantification of calcium transient parameters (Ca2+ Rel and Ca2+ Up) and uni-axial contraction

11

force (Figure 1).

12

13 14

Figure 1: Functional characterization of single hESCs-CM. hESCs-CMs are seeded on elastomeric

15

microstructured substrates that have been previously photopatterned to obtain cell-adhesive and

16

cell-repellent areas on the surface.

17

contraction. Confocal analysis allows the acquisition of calcium transients, recording of micropillar

18

deflection and visualization of focal adhesion points, and are repeated after 1 week and after 5

19

weeks of culture. The use of a proper finite element model (FEM) allows the quantification of

20

contraction forces.

Cells anchor to the micropillars and bend them upon

21 22

2. Materials and methods

23

2.1 Substrate fabrication

24

The elastomeric flat and microstructured substrates have been fabricated by a cell-adhesive

25

photopatternable material consisting of n-butylacrylate (n-BA) and methacryloxybenzophenone

26

(MABP). In particular, P(nBA-4%MABP) was obtained through free radical polymerization as

27

described in the Supplementary Information and in previous publications [44-46]. To tune cell

28

growth only in well defined regions, a thin layer of cell-repellent polyacrylamide (PAA) has been

29

photoattached by UV-light exposure to the P(nBA-co-4%MABP) surface, which prevented protein

5

ACS Paragon Plus Environment

Langmuir

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

Page 6 of 39

1

adsorption and cell adhesion everywhere, except than in certain areas which were kept cell

2

adhesive.

3 4

2.1.1 Fabrication of PDMS stamps

5

As first step, microstructured PDMS stamps have been realized by replica molding from micro-

6

structured Si-wafers obtained through standard photolitography technique. Briefly, a Si-wafer was

7

spin coated with a thin layer of negative photoresist (AZ-1518, MicroChemicals, Germany). The

8

photoresist was illuminated by UV-light (λ=365 nm) through a chrome mask (Delta Mask, The

9

Netherlands) that presented the desired pattern. Uncross-linked resist was removed rinsing the

10

wafer in the proper AZ-1518 developer (MicroChemicals, Germany) for few seconds. The wafer

11

was then etched to obtain the final micropillared stamp. The microstructured Si-wafer was

12

subsequently silanized by exposure to the vapor of (tridecafluoro-1,1,2,2-tetrahydrooctyl)-1-

13

trichlorosilane in vacuum for 30 minutes, to facilitate PDMS removal during the replica molding

14

step. PDMS (Sylgard 184, Dow-Corning. Midland, MI) elastomer was thoroughly mixed with the

15

silicone elastomer curing agent in a 10:1 ratio, poured over the microstructured Si-master and kept

16

under vacuum for 1 h to allow the complete filling of the pattern and air bubbles removal. The

17

sample was then cured at 80°C for 3 hours and subsequently peeled off the Si-wafers. These

18

samples were used as stamps for the subsequent fabrication of P(nBA-co-4%MABP)

19

microstructured substrates.

20 21

2.1.2 Fabrication of P(nBA-co-4%MABP) substrates

22

The fabrication of microstructured P(nBA-co-4%MABP) samples is described in figure 2: P(nBA-

23

co-4%MABP) polymer was dissolved in toluene (300 mg/ml) and a 60 µl drop of the resulting

24

solution was poured on the appropriate PDMS stamp. A silanized glass slide was then gently

25

pressed on the top of this drop, allowing the still not cross-linked polymer to completely fill the

26

pattern of the stamp. A covalent bond between the glass slide and the polymer was achieved

27

thanks to the use of triethoxybenzophenone (3EBP) silane, as described in the Supplementary

28

Information. This sandwich formed by PDMS stamp, P(nBA-co-4%MABP) and silanized glass was

29

exposed to UV-light (λ = 365 nm, P = 6.5 mW ) for 20 minutes. The cross-linked microstructured

30

samples covalently bonded to the glass slide were then peeled-off from the PDMS stamp.

31

To obtain flat P(nBA-co-4%MABP) samples, the solution was spin-coated on a silanized glass at a

32

velocity of 600 rpm for 1 minute and then irradiated by UV-light.

33 34

2.1.3 Photopatterning of flat and microstructured P(nBA-4%MABP) substrates with PAA

35

Acrylamide solution was prepared diluting acrylamide (Sigma Aldrich, Germany) in deionized water

36

obtaining a final concentration of 8%. PAA was supplemented with 30 mg/ml of photoinitiator

6

ACS Paragon Plus Environment

Page 7 of 39

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

Langmuir

1

(Irgacure, BASF, Germany) previously diluted in 100 µl methanol. Figure 2B shows the steps

2

necessaries to photopattern the P(nBA-4%MABP) substrates: A 70 µl drop of acrylamide/Irgacure

3

solution was poured on a glass coverslip, the P(nBA-co-%MABP) sample was turned upside down

4

and brought into contact with the polyacrylamide solution, and a photomask was gently placed on

5

the glass that supported the P(nBA-co-%MABP) sample.

6

This sandwich-like sample was then irradiated for 90 s with UV light (λ = 365 nm). As a result of

7

this process in the irradiated areas a surface-attached polymer network was generated. After UV

8

light exposure the whole elastomeric sample was covered by a thin layer of cell-repellent PAA,

9

except those areas that have not been shaded by the photomask and thus have not been exposed

10

to UV light. The sample was rinsed with deionized water for 3 minutes to remove all the residual

11

not cross-linked PAA. The same procedure has been applied both on flat elastomeric samples and

12

on microstructured elastomeric samples.

13 14

2.2 Mechanical characterization of the substrate

15

2.2.1 Dynamic Mechanical Thermal Analysis

16

The elastic properties of P(nBA-co-4%MABP) were tested through Dynamic Mechanical Thermal

17

Analysis (DMTA, developed on Instron 4502 Tensile Machine). To perform the DMTA test that

18

allowed to quantify the Young’s modulus of P(nBA-co-4%MABP) at different temperatures, specific

19

samples have been fabricated: the samples needed to be at least 3 mm thick and their length had

20

to be at least twice bigger than their width. For these reasons samples were fabricated by casting.

21

A solution of P(nBA-co-4%MABP) in toluene was prepared (300 mg/ml) and 2 ml of this solution

22

were poured in a rectangular Teflon stamp (2x5 cm). The solution was then exposed to UV-light for

23

30 minutes. The sample was peeled-off, turned upside-down, and exposed to UV-light for further

24

30 minutes until the film was totally cross-linked. The film thickness ranged from 3 to 3.5 mm. A

25

preload of 0.002 MPa was applied to each rectangular sample to assure the same tensile condition

26

for all the different samples before starting the test and a tensile loading all along the test. A

27

sinusoidal stress with a frequency of 1 Hz was applied deforming the samples of 1% maximum.

28

The frequency was fixed at 1 Hz because this frequency is comparable to the frequency of

29

deformation induced by the cells during contraction. The experiment was repeated for 5 samples.

30

2.1.2 Nanoindention tests

31

A nanoindenter (Nanoindenter CSM instruments, Switzerland) was used to experimentally

32

reproduce the process of bending of the micropillars when a tangential force is applied to the pillar

33

head. To this, a glass sphere was brought into contact with the microstructures of the specimen

34

and dragged laterally to induce bending of the micropillars while the tangential force was recorded.

35

When a specific tangential force is applied, the displacement of the top of the micropillars depends

36

on the material properties (Young’s modulus) and on the geometry of the micropillar array. An in-

7

ACS Paragon Plus Environment

Langmuir

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

Page 8 of 39

1

situ camera that recorded the contact through the transparent specimen during the test allowed to

2

simultaneously acquire the tangential force applied by the instrument and the micropillar deflection

3

(Figure 3B). More details on the nanoindenter test are provided in the Supplementary Information.

4 5

2.3 Finite element modeling

6

Finite element model (FEM) of the substrate was realized using the MSC MARC software. The FE

7

model of the substrate was developed according to the effective geometry configuration. Because

8

of the simple tensile response of P(nBA-co-4%MABP) in the range of moderate strains, a Hookean

9

constitutive model was adopted to model the material. Since the investigated cardiomyocytes act

10

between single couples of pillars, it was possible to estimate contractile forces of cells through an

11

indirect method, by applying assigned values of deflection (deduced experimentally) to the pillars

12

and estimating the corresponding forces. Deflections of the pillar were applied as imposed

13

displacements in the FEM. The force can be applied at different heights, as it happened in the

14

case of cell measurements.

15 16

2.4 Cell culture

17

The cardiac differentiation protocol has been adapted from Lian et al., [50]. The hESCs were

18

initially cultured on 0.5% Matrigel-coated (Corning) plates in mTeSR1 medium (StemCell

19

Technologies) until confluence. Differentiation was induced by medium change to RPMI 1640/B-27

20

w/o insulin (Invitrogen) supplemented with 12 µM CHIR99021 (Miltenyi). After 24 h, the medium

21

was changed to RPMI 1640/B27 w/o insulin for another 2 days. The medium was then changed to

22

RPMI 1640/B27 supplement w/o insulin and 10 µM IWP-4 (Stemgent). At day 5, the culture

23

medium was switched again to RPMI 1640/B27 w/o insulin for 2 days and then the cultures were

24

maintained for the entire culture time in RPMI 1640/B27. Medium was changed every 4 days. For

25

disaggregation, a digestion mix solution was prepared with collagenase I (2 mg/ml), collagenase IV

26

(1 mg/ml), DNase I (2 µl/ml) and 10 µM Y-27632 in PBS with Ca2+ and Mg2+. Cells were incubated

27

for 25 minutes at 37°C in this solution, washed in PBS w/o Ca2+ and Mg2+ and digested to single-

28

cell suspension with TryPLE Select reagent (Life technologies).

29

Single cell suspensions were plated on the substrates coated for 1 hour with 20 µg/ml laminin (BD,

30

Bioscience, USA) at a 4000 cells/mm2 density. Cells begun to contract after 3-4 days of seeding

31

and could be kept viable up to 8 weeks, changing the medium every 4-5 days.

32 33

2.5 Immunohistochemistry

34

Cells were fixed with PBS containing 2% paraformaldehyde (Sigma-Aldrich) for 7 minutes,

35

permeabilized with PBS containing 0.5% Triton X-100 (Sigma-Aldrich), and blocked in PBS

36

containing 2% horse serum for 45 minutes, at room temperature. Primary antibodies were applied

8

ACS Paragon Plus Environment

Page 9 of 39

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

Langmuir

1

for 1 hour at 37 °C. Cells were washed in PBS (Life Technologies) and incubated with

2

fluorescence-conjugated secondary antibodies against mouse for 45 minutes at 37 °C. Finally,

3

nuclei were counterstained with DAPI (Sigma Aldrich), and samples were viewed under Leica TCS

4

SP5 fluorescence confocal microscope (Leica Microsystems, Wetzlar, Germany). Primary

5

antibodies used were the following: mouse monoclonal anti-cTnT (Sigma Aldrich; 1:100 dilution),

6

mouse monoclonal anti-vinculin (Sigma-Aldrich; 1:100 dilution). Secondary antibody used was goat

7

anti-mouse (Invitrogen; 1:200 dilution). All antibodies were diluted in 3% bovine serum albumin

8

(Sigma Aldrich).

9 10

2.6 Calcium transients and micropillar deflection acquisition through confocal analysis

11

hESCs-CMs were loaded in serum-free Dulbecco’s modified Eagle medium (Invitrogen)

12

supplemented with 2.5 µmol/l of fluorescent calcium dye Fluo-4 AM (Invitrogen) for 20 minutes at

13

37°C in the presence of 2 µmol/l of Pluronic F-127 (Life Technologies) and 20 µmol/l of

14

sulfinpyrazone (Sigma-Aldrich), then incubated for additional 10 minutes at 37°C without Fluo-4

15

AM, and added with 0.2 µmol/l of di-8- ANEPPS (Invitrogen). Cell dynamics were obtained in

16

recording solution: NaCl, 125 mmol/l; KCl, 5 mmol/l; Na3PO4, 1 mmol/l; MgSO4, 1 mmol/l; CaCl2,

17

2 mmol/l; and glucose, 5.5 mmol/l, to pH 7.4 with NaOH. Line scans were acquired with a Leica

18

TCS SP5 fluorescence confocal microscope using a 63x oil immersion objective, with 488 nm Ar

19

laser line as an excitation source and 400 Hz acquisition frequency. Line scans were then

20

analyzed using Matlab (MathWorks) software to obtain calcium transient profile and quantify

21

micropillar deflection. For the calcium release phase (Ca2+ Rel), the time to peak value was

22

calculated considering the time from baseline to the 95% of the maximum value of fluo-4 intensity

23

recorded by the confocal microscope. For evaluating the calcium reuptake rate after contraction

24

(Ca2+ Up), the half-life of the calcium decay was considered. More information on the calculation of

25

Ca2+ Rel and Ca2+ Up are provided in the Supplementary Information (figure S1). The total time of

26

the calcium transients (calcium transient duration) has been calculated adding the time of calcium

27

release phase to two times the calcium reuptake phase: Ca2+ Tot = Ca2+ Rel + 2Ca2+ Up. The

28

deflection of the micropillars was recorded by following the fluorescence of di-8-ANEPPS, which

29

labeled both cell membranes and the surface of the micropillars, thanks to its lipophilic properties.

30 31

2.7 Statistical analysis

32

Data are presented as means +/- standard deviation. Data pairs were compared by non-directional

33

Student’s t-test. Correlation coefficients were calculated through the Pearson’s method. All data

34

manipulation and computation was performed with the Matlab (MathWorks) software.

35 36

3. Results

9

ACS Paragon Plus Environment

Langmuir

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

Page 10 of 39

1

3.1. Sample preparation

2

The micropillar structures (size 10µm x10µm, height 20µm, distance 25µm) are generated through

3

a standard micro-replication process. However, instead of using a standard polymer we employed

4

a polymer containing benzophenone units incorporated into the polymer. Upon UV-irradiation

5

these groups become activated into a biradicaloid triplet state, which lead to linking of two polymer

6

chains through (formal) C-H insertion reactions [44-47]. To locally modify the surfaces a photopoly-

7

merization reaction of acrylic acid (AA) was carried out during brief UV-exposure, where a

8

photomask prevented photoreaction to occur in the shaded areas. Transfer to polymer and

9

recombination lead a crosslinking of the forming polymer chains and thus to the formation of a PAA

10

network. The forming PAA network become attached to the surface presumably in two ways: on

11

the one hand grafting could occur through benzophenone units located at the surface, which had

12

been “surviving” the photocuring of the pillar material, could become activated and bind to the PAA

13

network. On the other hand, macroradicals could attach to the polymer substrate in a conventional

14

transfer reaction, where the macroradicals reacted with the substrate polymer. The thickness of the

15

PAA layer was determined by atomic force microscopy (AFM) and measured to be 250 ± 43 nm

16

(data not shown). The unmodified P(n-BA-4%-MABP) network surface is strongly non-polar and

17

thus allowed the adsorption of proteins such a laminin from solution rendering it cell adhesive. The

18

PAA-coated areas, however, represent a neutral, strongly swollen hydrogel, which prevented

19

protein adsorption and consequently cellular adhesion through entropic shielding/size exclusion

20

[47]. Thus cells brought into contact with such surface were strongly confined to that areas, which

21

were not covered by the hydrogel coating.

22 23 24

10

ACS Paragon Plus Environment

Page 11 of 39

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

Langmuir

1 2

Figure 2: Substrate fabrication. A) Fabrication of elastomeric microstructured substrates of P(nBA-

3

co-4%MABP): the polymer is crosslinked through UV light exposure. Upon illumination

4

benzophenone forms a biradicaloid triplet state, which abstracts a hydrogen atom from a

5

neighbouring C-H group of a polymer. The obtained closely adjacent two carbon radicals can

6

recombine and establish a covalent bond between the polymer chains and eventually crosslinking.

7

B) Photopatterning of PAA on P(nBA-co-4%MABP): PAA is exposed to UV light through a

8

photomask and photoattaches to P(nBA-co-4%MABP) surface forming a surface-attached network

9

only on the areas exposed to light. C) Laminin adsorption: laminin solution is incubated for 1 hour

10

on the substrate and the protein adsorbs on the P(nBA-co-4%MABP) surface.

11 12

3.2 Mechanical characterization

13

In order to mathematically derive the contraction force generated by the cells anchored on

14

micropillars, the determination of the mechanical properties of the elastomer employed in terms of

15

the modulus of the material is essential. Young’s modulus was measured by DMTA across an

16

interval of temperatures ranging from -40°C to 40°C (Figure 3).

17

applications such as cell cultures, temperatures of interest are above 25°C and the elastomer

In particular, for biological

11

ACS Paragon Plus Environment

Langmuir

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

Page 12 of 39

1

P(nBA-4%MABP) displayed a Young’s modulus of 1.5 ± 0.3 MPa at 36.5°C, corresponding to the

2

cell culture maintenance temperature (Figure 3C).

3

The bending tests performed on microstructured P(nBA-co-4%MABP) samples were instrumental

4

for developing a finite element model through which shear force applied to the micropillars and the

5

pillar deflection can be connected with each other. The samples consisted of a lattice of square

6

micropillars having a height o 10 µm, a width of 10 µm and an interaxial distance of 20 µm. A

7

round glass tip with a radius of 500 µm was put in contact to the top of the microstructured

8

substrate, without applying a normal load to avoid compression of the micropillars. Then the tip

9

was slightly moved laterally, and the tangential force produced between the tip and the bended top

10

of the elastomeric micropillars was recorded by the instrument while the camera showed the

11

displacement of the top of the micropillars. Assuming an equally distributed force among all the

12

micropillars in contact since the radius of the glass tip is much bigger than the micropillar width, the

13

tangential force ft acting on a single pillar could be calculated from the tangential force recorded by

14

the instrument (Ft) divided by the number of micropillars in contact. For the determination of the

15

deflection, the algorithm employed calculates the position of the centroid of the top of each pillar in

16

contact both in the undeformed (cundeformed ) and deformed configuration (cdeformed ). For each pillar

17

the displacement ∆xi is calculated as ∆xi = cdeformed − cundeformed. The final value of the displacement

18

d is given by the mean value of the displacements calculated for each micropillar: d = ∑ ∆xi / 9 for

19

i=1,..,9 .

20

The FE model has been validated by predicting the tangential forces which are must applied to the

21

top of the micropillars in order to achieve a certain deflection. Experimental data show a linear

22

elastic behavior of the pillar material for displacements of the top of the micropillars (∆x) of up to

23

2.5 µm. Within this range of displacement good agreement between experimental data and FEM

24

can be observed (Figure 3C). No permanent deformation was recorded once the force was

25

removed. Based on literature data [33,37], the expected contraction forces exerted by

26

cardiomyocytes on the micropillars fall within the linear response range of the plot, thus a Hookean

27

model for micropillar bending could be assumed.

28

12

ACS Paragon Plus Environment

Page 13 of 39

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

Langmuir

1 2

Figure 3: Mechanical characterization of P(nBA-co-4%MABP). A) Scheme of the bending test on

3

microstructured P(nBA-co-4%MABP) substrates. The glass sphere of a nanoindenter applies a

4

tangential force on the top of the elastomeric micropillars, bending them laterally (A.1) while

5

micropillar deflection is visualized through an in situ camera (A.2). Scale bar is 20 µm. B)

6

Tangential force vs micropillar deflection according to the experimental data obtained from the

7

bending test (circles) and the finite element model (line). C) Results of the DMTA test on P(nBA-

8

co-4%MABP) performed at a frequency of 1Hz. The Young’s modulus at 36.5°C was found to be

9

1.5 ± 0.3 Mpa.

10 11

3.3 Microscopy of the microstructures

12

The substrates presented in this work have been fabricated with the cell-adhesive elastomeric

13

polymer P(nBA-co-4%MABP), with a cell-repellent layer of polyacrylamide (PAA) photoattached to

14

the P(nBA-co-4%MABP) surface by photopolymerization with UV light.

15

patterns of PAA have been achieved by interposition between the sample and the light source a

16

chrome photomask, in order to cover and protect from PAA polymerization area desired to be cell-

17

adhesive (Figure 4A).

18

The cell-adhesive areas have been designed as rectangles with dimensions of 30x15 µm providing

19

a 2:1 ratio for cell adhesion and spreading. Moreover, as the micropillars have a width of 10 µm, a

20

height of 20 µm and an interaxial distance between pillars of 25 µm, these dimensions allowed to

21

coat such that the surfaces of two micropillars and the area between them remained cell adhesive,

Specific geometrical

13

ACS Paragon Plus Environment

Langmuir

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

Page 14 of 39

1

while the surrounding become strongly cell repellent (Figure 4B). The presence of the surface-

2

attached hydrogel in Figure 4B can be nicely seen through some wrinkling of the hydrogel occuring

3

during drying of the sample, which produces a clearly visible surface pattern.

4 5

6 7

Figure 4: Optical micrographs of surface-modified P(nBA-co-%MABP) substrates. Bright field

8

images of (A) a flat substrate with locally generated PAA layers and (B) of the same structures as

9

in (A) but on a microstructured substrate. Cell adhesive areas are marked by white dashed lines.

10

Scale bar is 30 µm. Details of the layer generation are given in the text.

11 12

3.4 Cell and substrate integration

13

hESCs-CMs were seeded on flat and on microstructured substrates after photopatterning with

14

linear PAA and functionalization with laminin solution. As the protein cannot adsorb to the PAA

15

coated areas, selective hESCs-CMs adhesion to the non-illuminated areas could be observed on

16

both types of substrates (Figure 5). hESCs-CMS could adhere to the cell-attractive areas of the

17

surface and after 4 days of culture the cardiomyocytes displayed spontaneous contracting activity,

18

demonstrating that laminin coated P(nBA-co-4%MABP)-surface does not impair cell function. The

19

cardiac culture could be maintained viable and contracting up to 8 weeks.

20

To evaluate the sarcomeric organization of hESCs-CMs, cells were fixed at day 20 to assay for

21

cardiac troponin-T (cTnT) by immunofluorescence staining (Figure 5B). Cardiomyocytes

22

constrained within the rectangular cell-adhesive areas spread assuming a rectangular shape and

23

displaying myofibril organization along the long axis of the cell (Figure 5A). When seeded on

24

microstructured and photopatterned substrates, hESCs-CMs anchored between two micropillars,

25

bending them during contraction (Figure 5B.1 and Figure 5B.2). The sarcomeric organization

26

perpendicular to the micropillars was assessed by immunofluorescence (Figure 5B.3), providing a

14

ACS Paragon Plus Environment

Page 15 of 39

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

Langmuir

1

clear proof of the generated force direction. Moreover, imaging of focal adhesion points by vinculin

2

staining (Figure 5B.4) allowed to determine at which height of the micropillars the force was

3

applied, in order to produce more accurate FEM data.

4 5

Figure 5: Cell and substrate integration. A) hESCs-CMs seeded on photopatterned substrates.

6

Cells selectively adhere to cell-adhesive areas (A.1) and assume a rectangular shape (A.2)

7

showing a sarcomeric organization comparable to that of embryonic cardiomyocytes, as indicated

8

by immunofluorescence analysis of cTnT (A.3). B) hESCs-CMs seeded on microstructured and

9

photopatterned substrates. Cells selectively adhere to cell adhesive areas extending between two

10

elastomeric micropillars (B.1) and deflecting them of the value ∆x upon contraction (B.2). B.3)

15

ACS Paragon Plus Environment

Langmuir

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

Page 16 of 39

1

Immunostaining of cTnT of a single hESCs-CM anchored between two micropillars. B.4) Focal

2

adhesion points indicated by vinculin immunostaining on a single hESCs-CM anchored between

3

two micropillars.

4 5

3.5 Contraction force and calcium transients along cell maturation

6

Fluorescence confocal microscopy allowed live imaging of fast dynamics of simultaneous events

7

by taking advantage of the spectral properties of the employed fluorescent dyes. The

8

cardiomyocytes were loaded with fluo-4 green fluorescent dye used to assay calcium dynamics,

9

while the red lipophilic dye di-8-ANEPPS was employed for micropillar visualization (Figure 6).

10

Line-scan mode was employed to achieve high recording speed (frequency of acquisition: 400 Hz)

11

and providing experimental data of calcium transient dynamics through fluo-4 fluorescence

12

intensity and micropillar edge deflection acquired against time (Figure 6A). Prior to data analyses,

13

we evaluated the effect of fluo-4 loading on calcium dynamics parameters. Comparison of Ca2+ Rel

14

and Ca2+ Up values for different loading conditions resulted in no significant differences when the

15

dye concentration was varied keeping the incubation time constant (Figure 2SA). When calcium

16

transient parameters were evaluated keeping the dye concentration constant and varying the

17

incubation time (Figure 2SB), we found a statistically significant deviation only at longer dye

18

incubation times than the 20 minutes used here.

19

Taking advantage of the grid of the microstructured substrate, the position of each individual

20

analyzed cardiomyocyte could be pinpointed for subsequent immunofluorescence analyses, aimed

21

at determining the exact height of cell anchorage on the pillars. This parameter, acquired through

22

anti-vinculin immunofluorescence staining and confocal sectioning, was essential for the correct

23

computation of the mathematical model employed to derive contraction force values (Figure 6B).

24

16

ACS Paragon Plus Environment

Page 17 of 39

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

Langmuir

1 2

Figure 6: Detection of calcium transient and quantification of contraction force on single hESCs-

3

CMs. A) Live-cell confocal analysis. A.1) Confocal image of a single hESCs-CM anchored between

4

two micropillars. Fluo-4 (green signal) allows to detect calcium transients while di-8-ANEPPS

5

adsorbed to micropillar walls (red) allows to detect micropillar deflection. A.2) Image of the

6

simultaneous acquisition of calcium transient (green signal) and micropillar deflection (red lines).

7

A.3) Plot of calcium transient (green) and consequent micropillar deflection (red). B) Quantification

8

of the contraction force from micropillar deflection. B.1) Confocal 3D image of a hESCs-CM

9

anchored between two elastomeric micropillars. After immunostaining of the cell with vinculin it is

10

possible to determine the location of the focal adhesion points of the cell, corresponding to the

11

points of application of the force in the model. B.2) FE modeling of the distribution of the tangential

12

force exerted by the cell when the contraction force is applied at a height of 16 µm.

13 14

Extension of time in culture of hPSC-CMs has been frequently linked to improvement in cardiac

15

functional features in single cardiomyocytes [48, 49]. To assess if our platform could detect such

16

trend, we analyzed the calcium transients and contraction force of hESCs-CMs after 1 week and 5

17

weeks from seeding on the microstructured substrates (Figure 7 A). Within this interval of time, the

18

calcium transient significantly shortened, both for calcium release rate and calcium reuptake rate

17

ACS Paragon Plus Environment

Langmuir

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

Page 18 of 39

1

(Figure 7B). Calcium release time (Ca2+ Rel) decreased from 0.35 ± 0.03 s at week 1 to 0.19 ± 0.02

2

s at week 5, whereas the half-life of calcium decay (Ca2+ Up) decreased from 0.36 ± 0.03 s at week

3

1 to 0.16 ± 0.02 s at week 5. We observed similar trends in calcium transients recorded in hESCs-

4

CMs cultured on flat P(nBA-co-4%MABP) substrates with Ca2+ Rel decreasing from 0.38 ± 0.05 s

5

at week 1 to 0.31 ± 0.03 s at week 5, and Ca2+ Up from 0.50 ± 0.09 s at week 1 to 0.38 ± 0.05 s at

6

week 5. Cells cultured on microstructured substrates showed better functional performance with

7

shorter calcium transients at both 1 week and 5 weeks of culture. The generated force associated

8

with the contraction in cardiomyocytes cultured on microstructured substrates has been quantified

9

for the same cells in culture for 1 and 5 weeks. We observed an increase in the developed force

10

from 502 ± 122 nN at week 1 to 719± 101 nN at week 5 of culture (Figure 7C). Interestingly, we

11

found a negative correlation (R=-0.784) between the length of the calcium transient and the

12

developed contraction force with cardiomyocytes displaying shorter calcium transients exerting

13

higher force on the micropillars (Figure 7D). Nevertheless, analyzing our data points for correlation

14

between Ca2+ transient length (R=0.081 and R=0.104, respectively week 1 and week 5) or

15

contraction force (R=0.032 and R=0.091, respectively week 1 and week 5) against beating

16

frequency, no statistical correlation was found. Assuming the immature phenotype of hPSC-CMs

17

[8], this observation is consistent with reported literature data, making consideration over inotropic

18

and lusitropic effects difficult (Figure 7E).

18

ACS Paragon Plus Environment

Page 19 of 39

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

Langmuir

1 2

Figure 7: Analysis of single hESCs-CM calcium transients and contraction force after 1 week and

3

after 5 weeks of culture. A) Examples of simultaneous acquisition of calcium transients and

4

contraction forces of single hESCs-CM after 1 week (A.1) and after 5 weeks (A.2). B) Comparison

5

of calcium release (B.1) and calcium reuptake (B.2) of single hESCs-CMs analyzed after 1 week

6

and after 5 weeks of culture on flat photopatterned substrates (light gray) and on microstructured

7

and photopatterned substrates (black). C) Contraction force of single hESCs-CMs quantified on

8

cells after 1 week (light gray) and after 5 weeks (black) of culture on microstructured and

9

photopatterned substrates. Cells analyzed after 5 weeks from seeding developed stronger forces

10

and shorter calcium transients than cells analyzed after 1 week from seeding. Student t-test p

11

values, p ≤ 0.01 (*), p ≤ 0.001 (**), n=15. D) Maximum contraction force vs calcium transients of

19

ACS Paragon Plus Environment

Langmuir

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

Page 20 of 39

1

cells analyzed after 1 week and after 5 weeks of culture. E) Distribution plot of calcium transients

2

duration (E.1) and maximum contraction force (E.2) against beating frequency.

3

4. Discussion

4

In this work, we presented an in vitro platform based on a micro-engineered elastomeric

5

biomaterial suitable for assaying functional features of human embryonic stem cell-derived

6

cardiomyocytes (hESC-CMs) through optical imaging. Our main goal was to provide a viable

7

technology for contraction-force measurement in a human single cardiomyocyte, with simultaneous

8

acquisition of calcium dynamics. To this end, we developed a microengineered substrate

9

fabricated with a novel material, tuning its surface chemistry in order to achieve geometrically

10

constrained human cardiac cultures. In particular, we showed that calcium transients and

11

contraction force of single hESCs-CMs can be simultaneously acquired in cell cultures that can

12

span several weeks (up to 8 weeks). This time frame is suitable for both testing the long-term

13

effects of pharmacological treatment and analyze the progression of cardiac features upon

14

extended culture time.

15

Through proper biomaterial design, micro-fabrication and micro-patterning techniques, we

16

successfully achieved fine control of cell shape in an end-to-end cell configuration between two

17

micropillars. Previous studies are based on PDMS micropillared surfaces [38-43]. However PDMS

18

surface chemistry is difficult to modify permanently; atmospheric air plasma and argon plasma

19

oxidation can be used to alter the surface chemistry, adding silanol (SiOH) groups to the surface.

20

Nevertheless oxidized surfaces are stable for only short periods of times, independently from the

21

surrounding medium, i.e. regardless of whether the surfaces are stored in vacuum, air, or water.

22

To overcome PDMS limits and obtain a surface having tunable surface mechanical and chemical

23

properties, we combined the cell-adhesive photocrosslinkable elastomeric polymer P(nBAco-%-

24

MABP) with the cell-repellent polymer PAA. Fabrication of substrates in P(nBA-co-%MABP)

25

showed different advantages: i) dilution of polymer in organic solvent allows easy replica-molding

26

fabrication processes; ii) crosslinking the pre-polymer through UV light exposure allow tunable

27

mechanical properties; iii) the polymer is transparent and has a low fluorescence background; iv)

28

crosslinked polymer can be easily functionalized by a layer of proteins that are adsorbed by the

29

surface; v) thanks to the presence of the benzophenone groups, the surface of the polymer can be

30

activated by UV light and easily functionalized.

31

These peculiar characteristics allowed fine tuning of substrate geometry both in 3D (configuration

32

of micro-pillars: cross-section, interspace or height) and in 2D (surface chemistry leading to cell-

33

adhesive and cell-repellent areas). Moreover, the photolithographic process allowed to accurately

34

overlap the surface chemistry pattern with the micro-pillar grid. These specific features allowed for

35

accurate and reproducible large scale hESC-CMs patterning.

20

ACS Paragon Plus Environment

Page 21 of 39

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

Langmuir

1

Force measurement in cardiac cultures has been matter of several studies in the past, and

2

microstructured substrates found vast application in this field. A number of works employed dense

3

arrays of microposts deflected by contracting cell cultured on top of them [25-28, 32-34, 38]. This

4

experimental approach shares similar downfalls to the more common traction force microscopy

5

with compliant hydrogels embedded with glass or fluorescent beads [35, 50, 51]: it is a 2D system

6

with traction forces derived by tangential shear of the spread cell membrane against a broad

7

surface. The golden standard for force measurement is uni-axial traction measurement and it has

8

been proven to work for engineered cardiac tissues [52] but is not for single cell systems. Only

9

recently, Taylor and colleagues developed the sacrificial layer techniques on which hESC-CMs

10

were able to randomly adhere between two or three adjacent pillars [40]. The authors show a proof

11

of concept that this technique could be used for axial force detection in primary neonatal rat

12

cardiomyocytes by phase contrast imaging.

13

In order to provide a more comprehensive analysis of physiological and pathophysiological

14

mechanisms in heart muscle cells, simultaneous detection of uni-axial force and additional cardiac

15

features (such as calcium handling) would be of great value.

16

For this reason, we developed a biomaterial-based platform that combines the measurement of the

17

axial contractile force of hESCs-CMs with the detection of the corresponding calcium transients

18

using confocal imaging.

19

In order to obtain accurate measurements of forces during contraction, micropillar displacement

20

was correlated to a force value through the application of a finite element model (FEM) developed

21

according to the specific geometry and mechanical properties of the substrate. The contraction

22

force was quantified considering the actual points of cell anchorage to the micropillars, detected by

23

3D reconstruction of cell-pillar system by immunofluorescence for cell focal adhesion and confocal

24

sectioning. This computational effort provided proper analysis of single cell contraction parameters

25

regardless of the individual cell-pillar coupling conditions.

26

We also showed that the morphology and chemistry of the substrate are fitting to long-term cell

27

culture requirements (up to 8 weeks). hESC-CMs on the substrate showed a highly oriented

28

sarcomeric organization. As proof of concept, we applied this experimental setup to follow calcium

29

transients and contraction force evolution of hESC-CMs after 1 week and after 5 weeks of culture

30

on the microstructured substrate. Extension of time in culture has been often reported as a

31

determinant in cardiac structural and functional maturation [9, 48, 49, 53] and we showed here how

32

our experimental setup can be instrumental in analyzing the functional features of human

33

cardiomyocytes maturing in vitro. Cardiomyocytes analyzed after 5 weeks in culture showed on

34

average shorter calcium transients (figure 7B) and an associated increased contraction force (fig

35

7C). Although hPSC-CMs are an early and immature cell type, often referred as fetal-like [8],

36

characterized by absence of structural feature such as t-tubules [54-56], underdeveloped SR [57]

21

ACS Paragon Plus Environment

Langmuir

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

Page 22 of 39

1

and mixed contribution of SR-stores and trans-sarcolemmal Ca2+ entry to calcium transients [58-

2

60], the hESC-CMs still represent an excellent biological substrate for our experimental platform

3

validation, holding the potential to become a reliable model for the human myocardium in vitro.

4

To our knowledge, this is a first measurement of purely uni-axial contraction force for hESC-CMs,

5

with simultaneous acquisition of calcium dynamics. The feasibility of our experimental approach for

6

contraction force measurements is validated by comparison to the absolute values of reported

7

contraction force measurements made with different experimental setups, although challenging

8

because mostly based on cell in adhesion [41, 42] and reported as tangential stress (i.e. nN/mm2

9

of cell area or nN/post in the case of micro-pillared substrate) [34, 35, 38, 39]. For instances,

10

measurements of 25-45 nN/post have been reported by Rodriguez et al. on human induced

11

pluripotent stem cells-derived cardiomyocytes, using elastomeric force post arrays [38], whereas

12

Ribeiro et al. reported contraction stresses ranging from 0.26 to 0.25 mN/mm2 for both hESC-CM

13

and hiPS-CM [35]. Using micropost arrays, semi-quantitative estimation of 150 nN uni-axial force

14

have been performed by Taylor et al. for neonatal rat cardiomyocytes [40].

15

Altogether, we provide a microengineered platform suitable for precise and high-throughput

16

contraction force measurement in single-cardiomyocyte systems. Pairing force measurements with

17

calcium dynamics increases considerably the experimental value of the acquired data.

18

Furthermore, since micropillar displacement can be simultaneously detected along with other

19

optically encoded measures, further cardiac functional parameters (such as membrane potential

20

detected by fluorescent probes [61]) could be simultaneously detected. With the correct imaging

21

protocol in place, this platform is suitable for high-throughput monitoring of drug-induced changes

22

on more than one cardiac feature at once and can be a valuable tool in drug-screening.

23 24

Conclusions

25

We presented a biomaterial-based platform to perform quantitative in vitro studies on hESC-CMs.

26

In particular we showed simultaneous detection of the uni-axial contraction force and calcium

27

transients of hESCs-CMs with a defined morphology. To guide cell shape, we combined the

28

generation of a 3-dimensional micropillar array with a photochemical surface modification process.

29

As result, an elastomeric micropillar grid coupled with a matrix of single cell adhesive regions were

30

successfully developed. hESC-CMs, selectively grown only on confined areas of the surface,

31

extended between two micropillars and assumed an elongated shape. Calcium transients together

32

with uni-axial contraction force were acquired through confocal analysis. The use of a finite

33

elements model (FEM), which describes the deflection of the micropillars when a tangential force is

34

applied and 3D confocal reconstruction allowed accurate estimation of contraction forces. We

35

reported that the shortening of calcium transient during several weeks of culture is inversely

36

correlated to the uni-axial contraction force detected in single hESC-CMs. The platform can be

22

ACS Paragon Plus Environment

Page 23 of 39

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

Langmuir

1

suitable for physiological and pathophysiological in vitro assays as well as for monitoring functional

2

maturation of hPSCs-CMs.

3 4

Acknowledgements

5

The authors are grateful for the financial support from the Deutsche Forschungsgemeinschaft

6

(DFG) through the IRTG Soft Matter Science, from the University of Padova through the Progetto

7

Strategico TRANSAC and the grant Giovani Studiosi 2010 (DIRPRGR10), from “Amici del cuore di

8

Montebelluna” and from “ASCM”.

9 10 11 12 13

References

14

[1] Yin, S.; Zhang, X.; Zhan, C.; Wu, J.; Xu, J.; Cheung, J. Measuring single cardiac myocyte

15

contractile force via moving a magnetic bead. Biophys. J. 2005, 88, 1489-1495.

16

[2] Chang, W.T.; Yu, D.; Lai, Y.C.; Lin, K.Y.; Liau, I. Characterization of the mechanodynamic

17

response of cardiomyocytes with atomic force microscopy. Anal. Chem. 2013, 85, 1395-1400.

18

[3] Walker, C.A.; Spinale, F.G. The structure and function of the cardiac myocyte: a review of

19

fundamental concepts. J. Thorac. Cardiovasc. Surg. 1999, Aug, 118, 375-382.

20

[4] Bers, D. Cardiac excitation-contraction coupling. Nature 2002, 415, 198-205.

21

[5] Koivumaki, J.T.; Takalo, J.; Korhonen, T.; Tavi, P.; Weckstrom, M. Modelling sarcoplasmic

22

reticulum calcium ATPase and its regulation in cardiac myocytes. Philos. Trans. A Math. Phys.

23

Eng. Sci. 2009, 367, 2181-2202.

24

[6] Mathers, C.D.; Bernard, C.; Iburg, K.M.; Inoue, M.; Ma Fat, D.; Shibuya, K.; Stein, C.; Tomijima,

25

N.; Xu, H. Global burden of disease: data sources, methods and results. World Health

26

Organization, Geneva, Switzerland, 2008.

27

[7] Dick, E.; Rajamohan, D.; Ronksley, J.; Denning, C. Evaluating the utility of cardiomyocytes

28

from human pluripotent stem cells for drug screening. Biochem. Soc. Trans. 2010, 38, 1037-1045.

29

[8] Roberston, C.; Tran, D.D.; George, S.C. Concise review: maturation phases of human

30

pluripotent stem cell-derived cardiomyocytes. Stem Cells 2013, 31, 829-837.

23

ACS Paragon Plus Environment

Langmuir

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

Page 24 of 39

1

[9] Otsuji, T.G.; Minami, I.; Kurose, Y.; Yamauchi, K.; Tada, M.; Nakatsuji, N. Progressive

2

maturation in contracting cardiomyocytes derived from human embryonic stem cells: qualitative

3

effects on electrophysiological responses to drugs. Stem Cell Res. 2010, 4, 201-213.

4

[10] Gilchrist, K.H.; Lewis, G.F.; Gay, E.A.; Sellgren, K.L.; Grego, S. High-throughput cardiac

5

safety evaluation and multi-parameter arrhythmia profiling of cardiomyocytes using microelectrode

6

arrays. Toxicol Appl Pharmacol. 2015, 288, 249-257.

7

[11] Itzhaki, I.; Rapoport, S.; Huber, I.; Mizrahi, I.; Zwi-Dantsis, L.; Arbel, G.; Schiller, J.; Gepstein,

8

L. Calcium handling in human induced pluripotent stem cell derived cardiomyocytes. PLoS One

9

2011, 6, e18037.

10

[12] Lu, H.R.; Whittaker, R.; Price, J.H.; Vega, R.; Pfeiffer, E.R.; Cerignoli, F.; Towart, R.;

11

Gallacher, D.J. High throughput measurement of Ca++ dynamics in human stem cell-derived

12

cardiomyocytes by kinetic image cytometery: a cardiac risk assessment characterization using a

13

large panel of cardioactive and inactive compounds. Toxicol Sci. 2015, 148, 503-516.

14

[13] Dolnikov K, Shilkrut M, Zeevi-Levin N, Gerecht-Nir S, Amit M, Danon A, Itskovitz-Eldor J,

15

Binah O. (2006) Functional properties of human embryonic stem cell-derived cardiomyocytes:

16

intracellular Ca2+ handling and the role of sarcoplasmic reticulum in the contraction. Stem Cells.

17 18 19

Feb;24(2):236-45.

20

through small gaps. European Biophysics Journal with Biophysics Letters 2006, 35, 713-719.

21

[15] Kress, H.; Stelzer, E.H.K.; Holzer, D.; Buss, F.; Griffiths, G.; Rohrbach, A. Filopodia act as

22

phagocytic tentacles and pull with discrete steps and a load-dependent velocity. Proc. Natl. Acad.

23

Sci. U.S.A. 2007, 104, 11633-11638.

24

[16] Prass, M.; Jacobson, K.; Mogilner, A.; Radmacher, M. Direct measurement of the lamellipodial

25

protrusive force in a migrating cell. J. Cell Biol. 2006, 174, 767-772.

26

[17] Tanaka, Y.; Morishima, K.; Shimizu, T.; Kikuchi, A.; Yamato, M.; Okano, T.; Kitamori, T.

27

Demonstration of a PDMS-based bio-microactuator using cultured cardiomyocytes to drive

28

polymer micropillars. Lab Chip 2006, 6, 230-235.

29

[18] Brady, A.J.; Tan, S.T.; Ricchiuti, N.V. Contractile force measured in unskinned isolated adult

30

rat heart fibres. Nature 1979, 282, 728-729.

31

[19] Merkel, R.; Kirchgebner, N.; Cesa, C.M.; Hoffmann, B. Cell force Microscopy on elastic layers

32

of finite thickness. Biophys. J. 2007, 93, 3314-3323.

[14] Brunner, C.A.; Ehrlicher, A.; Kohlstrunk, B.; Knebel, D.; Kaes, J.A.; Goegler, M. Cell migration

24

ACS Paragon Plus Environment

Page 25 of 39

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

Langmuir

1

[20] Das, T.; Maiti, T.K.; Chakraborty, S. Traction force microscopy on-chip: shear deformation of

2

fibroblast cells. Lab Chip 2008, 8, 1308-1318.

3

[21] Harris, A.K.; Wild, P.; Stopak, D. Silicone rubber substrata: A new wrinkle in the study of cell

4

locomotion. Science 1980, 208, 177-179.

5

[22] Jacot, J.G.; McCulloch, A.D.; Omens, J.H. Substrate stiffness affects the functional maturation

6

of neonatal rat ventricular myocytes. Biophys. J. 2008, 95, 3479-3487.

7

[23] Hersch, N.; Wolters, B.; Dreissen, G.; Springer, R.; Kirchgessner, N.; Merkel, R.; Hoffmann, B.

8

The constant beat: cardiomyocytes adapt their forces by equal contraction upon environmental

9

stiffening. Biol. Open 2013, 15, 351-361.

10

[24] Vannier, C,; Chevassus, H.; Vassort, G. Ca-dependence of isometric force kinetics in single

11

skinned ventricular cardiomyocytes from rats. Cardiovasc. Res. 1996, 32, 580-586.

12

[25] Kajzar, A.; Cesa, C.M.; Kirchgessner, N.; Hoffmann, B.; Merkel, R. Toward physiological

13

conditions for cell analyses: Forces of heart muscle cells suspended between elastic micropillars.

14

Biophys J. 2008, 94, 1854-1866.

15

[26] Kim, K.; Taylor, R.; Sim, J.Y.; Pak, S.-J.; Norman, J.; Fajardo, G.; Bernstein, D.; Priutt, B.L.

16

Calibrated micropost arrays for biomechanical characterisation of cardiomyocytes. Micro & Nano

17

Letters 2011, 6, 317-322.

18

[27] Balaban, N.; Schwarz, U.; Riveline, D.; Goichberg, P.; Tzur, G.; Sabanay, I.; Mahalu, D.;

19

Safran, S.; Bershadsky, A.; Addadi, L.; Geiger, B. Force and focal adhesion assembly: a close

20

relationship studied using elastic micropatterned substrates. Nat. Cell Biol. 2001, 3, 466-472.

21

[28] Cesa, C.M.; Kirchgessner, N.; Mayer, D.; Schwarz, U.S.; Hoffmann, B.; Merkel, R.

22

Micropatterneded silicone elastomer substrates for high resolution analysis of cellular force

23

patterns. Rev. Sci. Instrum. 2007, 78, 1-10.

24

[29] Barentin, C.; Sawada, Y.; Rieu, J. An iterative method to calculate forces exerted by single

25

cells and multicellular assemblies from the detection of deformations of flexible substrates. Eur.

26

Biophys. J. 2006 , 35, 328-339.

27

[30] Fu, J.; Wang, Y.; Yang, M.T.; Desai, R.A.; Yu, X.; Liu, Z.; Chen C.S. Mechanical regulation of

28

cell function with geometrically modulated elastomeric substrates. Nature Methods 2010, 7, 733-

29

736..

25

ACS Paragon Plus Environment

Langmuir

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

Page 26 of 39

1

[31] Klein, F.; Striebel, T.; Fischer, J.; Jiang, Z.; Franz, C.M.; Von Freymann G.; Wegener M.;

2

Bastmeyer M. Elastic Fully Three-dimensional Microstructure Scaffolds for Cell Force

3

Measurements. Adv. Mater. 2010, 23, 868-871.

4

[32] Le Digabel, J.; Ghibaudo, M.; Trichet, L.; Richert, A.; Ladoux, B. Microfabricated substrates as

5

a tool to study cell mechanotransduction. Med. Biol. Eng. Comput. 2010, 48, 965-976.

6

[33] Tan, J.; Tien, J.; Pirone, D.; Gray, D.; Bhadriraju, K.; Chen, C. Cells lying on a bed of

7

microneedles: An approach to isolate mechanical force. Proc. Natl. Acad. Sci. U.S.A. 2003, 100,

8

1484-1489.

9

[34] Rodriguez, A.G.; Han, S.J.; Regnier, M.; Sniadecki, N.J. Substrate stiffness increases twitch

10

power of neonatal cardiomyocytes in correlation with changes in myofibril structure and

11

intracellular calcium. Biophys. J. 2011, 101, 2455-2464.

12

[35] Ribeiro, M.C.; Tertoolen, L.G.; Guadix, J.A.; Bellin M.; Kosmidis, G.; D'Aniello, C.;

13

Monshouwer-Klootsa, J.; Goumansc, M.-L.; Wangd, Y.-L.; Feinbergd, A.W., Mummery, C.L.;

14

Passier, R. Functional maturation of human pluripotent stem cell derived cardiomyocytes in vitro:

15

correlation between contraction force and electrophysiology. Biomaterials. 2015, 51, 138-150

16

[36] Kita-Matsuo, H.; Barcova, M.; Prigozhina, N.; Salomonis, N.; Wei, K.; Jacot, J.G.; Nelson, B.;

17

Spiering, S.; Haverslag, R.; Kim, C.; Talantova, M.; Bajpai, R.; Calzolari, D.; Terskikh, A.;

18

McCulloch, A.D.; Price, J.H.; Conklin, B.R.; Chen, H.S.; Mercola, M. Lentiviral vectors and

19

protocols for creation of stable hESC lines for fluorescent tracking and drug resistance selection of

20

cardiomyocytes. PLoS One 2009, 4, e5046.

21

[37] Hazeltine, L.B.; Simmons, C.S.; Salick, M.R.; Lian, X.; Badur, M.G.; Han, W.; Delgado, S.M.;

22

Wakatsuki, T.; Crone, W.C.; Pruitt, B.L.; Palecek, S.P. Effects of substrate mechanics on

23

contractility of cardiomyocytes generated from human pluripotent stem cells. Int. J. Cell Biol. 2012,

24

2012, 508294.

25

[38] Rodriguez, M.L.; Graham, B.T.; Pabon, L.M.; Han, S.J.; Murry, C.E.; Sniadecki,

26

N.J..Measuring the contractile forces

27

cardiomyocytes with arrays of microposts. J. Biomech. Eng. 2014, 136, 051005.

28

[39] Yang, X.; Rodriguez, M.; Pabon, L.; Fischer, K.A.; Reinecke, H.; Regnier, M.; Sniadecki, N.J.;

29

Ruohola-Baker, H.; Murry, C.E.. Tri-iodo-l-thyronine promotes the maturation of human

30

cardiomyocytes-derived from induced pluripotent stem cells. J. Mol. Cell Cardiol. 2014, 72, 296-

31

304.

of human induced

pluripotent

stem

cell-derived

26

ACS Paragon Plus Environment

Page 27 of 39

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

Langmuir

1

[40] Taylor, R.E.; Kim, K.; Sun, N.; Park, S.J.;, Sim, J.Y.; Fajardo, G.; Bernstein, D.; Wu, J.C.;

2

Pruitt, B.L. Sacrificial layer technique for axial force post assay of immature cardiomyocytes.

3

Biomed. Microdevices. 2013, 15, 171-181.

4

[41] Stoehr, A.; Neuber, C.; Baldauf, C.; Vollert, I.; Friedrich, F.W.; Flenner, F.; Carrier, L.; Eder, A.;

5

Schaaf, S.; Hirt, M.N.; Aksehirlioglu, B.; Tong, C.W.; Moretti, A.; Eschenhagen, T.; Hansen, A.

6

Automated analysis of contractile force and Ca2+ transients in engineered heart tissue. Am. J.

7

Physiol. Heart Circ. Physiol. 2014, 306, 1353-1363.

8

[42] Kensah, G.; Roa, L.A.; Dahlmann, J.; Zweigerdt, R.; Schwanke, K.; Hegermann, J.; Skvorc,

9

D.; Gawol, A.; Azizian, A.; Wagner, S.; Maier, L.S.; Krause, A.; Dräger, G.; Ochs, M.; Haverich, A.;

10

Gruh, I.; Martin, U. Murine and human pluripotent stem cell-derived cardiac bodies form contractile

11

myocardial tissue in vitro. Eur. Heart J. 2013, 34, 1134-1146.

12

[43] Thavandiran N., Dubois, N.; Mikryukov, A.; Massé, S.; Beca, B.; Simmons, C.A.; Deshpande,

13

V.S.; McGarry, J.P.; Chen, C.S.; Nanthakumar, K.; Keller, G.M.; Radisic, M.; Zandstra, P.W.

14

Design and formulation of functional pluripotent stem cell-derived cardiac microtissues. Proc. Natl.

15

Acad. Sci. U.S. 2013, 110, 698-707.

16

[44] Belardi, J.; Schorr, N.; Prucker, O.; Rühe, J. Artificial Cilia: Generation of magnetic actuators in

17

microfluidic systems. Adv. Funct. Mater. 2011, 21, 3314-3320.

18

[45] Schuh, K.; Prucker, O.; Rühe, J. Tailor-made polymer multilayers. Adv. Funct. Mat. 2013, 23,

19

6019–6023.

20

[46] Prucker, O.; Naumann, C.A.; Rühe, J.; Knoll, W.; Frank, C.W. Photochemical attachment of

21

polymer films to solid surfaces via monolayers of benzophenone Derivatives. J. Am. Chem. Soc.

22

1999, 121, 8766-8770.

23

[47] Pandiyarajan, C.K.; Prucker, O.; Zieger, B.; Rühe, J. Influence of molecular structure of

24

surface-attached Poly (N-alkylacrylamide) coatings on the interaction of surfaces with proteins,

25

cells and blood platelets. Macromolecular Bioscience 2013, 13, 873-884.

26

[48] Lundy, S.D.; Zhu, W.Z.; Regnier, M.; Laflamme, M.A. Structural and functional maturation of

27

cardiomyocytes derived from human pluripotent stem cells. Stem Cells Dev. 2013, 22, 1991-2002.

28

[49] Kamakura, T.; Makiyama, T.; Sasaki, K.; Yoshida, Y.; Wuriyanghai, Y.; Chen, J.; Hattori, T.;

29

Ohno, S.; Kita, T.; Horie, M., Yamanaka, S.; Kimura, T. Ultrastructural maturation of human-

30

induced pluripotent stem cell-derived cardiomyocytes in a long-term culture. Circ. J. 2010, 77,

31

1307-1314.

27

ACS Paragon Plus Environment

Langmuir

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

Page 28 of 39

1

[50] Polio, S.R.; Rothenberg, K.E.; Stamenovic, D.; Smith, M.L. A micropatterning and image

2

processing approach to simplify measurement. Acta Biomateralia 2012, 8, 82-88.

3

[51] Kosmidis, G.; Bellin, M.; Ribeiro, M.C.; van Meer, B.; Ward-van Oostwaard, D.; Passier, R.;

4

Tertoolen, L.G.; Mummery, C.L.; Casini. S. Altered calcium handling and increased contraction

5

force in human embryonic stem cell derived cardiomyocytes following short term dexamethasone

6

exposure. Biochem. Biophys. Res. Commun. 2015 , 467, 998-1005.

7

[52] Yasuda, S.I.; Sugiura, S; Kobayakawa, N.; Fujita, H.; Yamashita, H.; Katoh, K.; Saeki, Y.;

8

Kaneko, H.; Suda, Y.; Nagai, R.; Sugi, H. A novel method to study contraction characteristics of a

9

single cardiac myocyte using carbon fibers. Am. J. Physiol. Heart. Circ. Physiol. 2001, 281, 1442-

10

1446.

11

[53] Sartiani, L.; Bettiol, E.; Stillitano, F.; Mugelli, A.; Cerbai, E.; Jaconi, M.E. Developmental

12

changes in cardiomyocytes differentiated from human embryonic stem cells: a molecular and

13

electrophysiological approach. Stem Cells 2007, 5, 1136-1144.

14

[54] Snir, M.; Kehat, I.; Gepstein, A.; Coleman, R.; Itskovitz-Eldor, J.; Livne, E.; Gepstein, L.

15

Assessment of the ultrastructural and proliferative properties of human embryonic stem cell-

16

derived cardiomyocytes. Am. J. Physiol. Heart Circ, Physiol. 2003, 285, 2355-2363.

17

[55] Lieu, D.K.; Liu, J.; Siu, C.W.; McNerney, G.P.; Tse, H.F.; Abu-Khalil, A.; Huser, T.; Li, R.A.

18

Absence of transverse tubules contributes to non-uniform Ca(2+) wavefronts in mouse and human

19

embryonic stem cell-derived cardiomyocytes. Stem Cells Dev, 2009, 18, 1493-1500.

20

[56] Denning, C.; Borgdorff, V.; Crutchley, J.; Firth, K.S.; George, V.; Kalra, S.; Kondrashov, A.;

21

Hoang, M.D.; Mosqueira, D.; Patel, A.; Prodanov, L.; Rajamohan, D.; Skarnes, W.C.; Smith, J.G.;

22

Young, L.E. Cardiomyocytes from human pluripotent stem cells: From laboratory curiosity to

23

industrial biomedical platform. Biochim. Biophys. Acta 2016, 1863, 1728-1748.

24

[57] Liu, J.; Fu, J.D.; Siu, C.W.; Li, R.A. Functional sarcoplasmic reticulum for calcium handling of

25

human embryonic stem cell-derived cardiomyocytes: insights for driven maturation. Stem Cells

26

2007, 25, 3038-3044.

27

[58] Zhu, W.Z.; Santana, L.F.; Laflamme, M.A. Local control of excitation-contraction coupling in

28

human embryonic stem cell-derived cardiomyocytes. PLoS One 2009, 4, e5407.

29

[59] Lee, Y.K.; Ng, K.M.; Lai, W.H.; Chan, Y.C.; Lau, Y.M.; Lian, Q.; Tse, H.F.; Siu, C.W. Calcium

30

homeostasis in human induced pluripotent stem cell-derived cardiomyocytes. Stem Cell Rev.

31

2011, 7, 976-986.

28

ACS Paragon Plus Environment

Page 29 of 39

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

Langmuir

1

[60] Pillekamp, F.; Haustein, M.; Khalil, M.; Emmelheinz, M.; Nazzal, R.; Adelmann, R.; Nguemo,

2

F.; Rubenchyk, O.; Pfannkuche, K.; Matzkies, M.; Reppel, M.; Bloch, W.; Brockmeier, K.;

3

Hescheler, J. Contractile properties of early human embryonic stem cell-derived cardiomyocytes:

4

beta-adrenergic stimulation induces positive chronotropy and lusitropy but not inotropy. Stem Cells

5

Dev. 2012, 10, 2111-2121.

6

[61] Miller, E.W.; Lin, J.Y.; Frady, E.P.; Steinbach, P.A.; Kristan, W.B.; and Tsien, R.Y.; Optically

7

monitoring voltage in neurons by photo-induced electron transfer through molecular wires. PNAS

8

2011, 209, 2114-2119.

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 29

ACS Paragon Plus Environment

Langmuir

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

1

Page 30 of 39

TOC Grafic

2 3

30

ACS Paragon Plus Environment

Page 31 of 39

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28

Langmuir

ACS Paragon Plus Environment

Langmuir

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

ACS Paragon Plus Environment

Page 32 of 39

Page 33 of 39

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

Langmuir

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26

ACS Paragon Plus Environment

Page 34 of 39

Page 35 of 39

Langmuir

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

Langmuir

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

ACS Paragon Plus Environment

Page 36 of 39

Page 37 of 39

Langmuir

A

A.2 800

Fluo-4 (F/F0)

Ca2+ Force

Contraction force (nN)

1

Week 5

1

800 Ca2+ Force

700 600 500 400 300 200

Fluo-4 (F/F0)

Week 1

0.2

0.4

0.6

0.8

Time (s)

700 600 500 400 300 200

100 0

Contraction force (nN)

A.1

100

0 1.0 1.2

0

0.2

0.4

0.6

0.8

Time (s)

0 1.0 1.2

B

C B.2 Calcium Release

0.75

Week 1 Week 5 0.5

0.25

Fluo-4 (F/F0)

Calcium Reuptake

1000

Week 1 Week 5 0.5

0.25

0

0

E

D

E.1 900

4c

800 700 600 500 400 300 200 0.4

0.6

0.8

1.0

Calcium transient (s)

1.2

Calcium transient (s)

1.4 Week 1 Week 5

1.2 1.0 0.8 0.6 0.4 0.2

Week 1 Week 5

0 Paragon Plus Environment ACS 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3

Beating frequency (Hz)

800

Week 1 Week 5

600 400 200 0

E.2 Max contraction force (nN)

Fluo-4 (F/F0)

0.75

Max contraction force (nN)

B.1

Max contraction force (nN)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

900 800 700 600 500 400 300 200

Week 1 100 Week 5 0 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3

Beating frequency (Hz)

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

ACS Paragon Plus Environment

Page 38 of 39

Page 39 of 39

Langmuir

A

Ca2+ Up

Ca2+ Rel

1.5

2.5 Time (s)

Time (s)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

1 0.5

2 1.5 1 0.5

0

n=15

n=13

n=16

1

2.5

5

0

n=15

n=13

n=16

1

2.5

5

Fluo-4 concentration (µM)

Fluo-4 concentration (µM)

Ca2+ Rel

Ca2+ Up

B 2.5

**

1

Time (s)

Time (s)

1.5

*

0.5

*

2 1.5 1 0.5

0

n=25

1

n=25

n=27

n=26

n=22

n=25

0 5 10 20 ACS 40Paragon Plus Environment 1

Fluo-4 incubation time (min)

n=25

n=27

n=26

n=22

5 10 20 40

Fluo-4 incubation time (min)