Direct Production of Human Cardiac Tissues by Pluripotent Stem Cell

Sep 1, 2016 - Direct Production of Human Cardiac Tissues by Pluripotent Stem Cell Encapsulation in Gelatin Methacryloyl ... Direct stem cell encapsula...
0 downloads 0 Views 4MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Direct Production of Human Cardiac Tissues by Pluripotent Stem Cell Encapsulation in Gelatin Methacryloyl Petra Kerscher, Jennifer A Kaczmarek, Sara E Head, Morgan Brazel, Wen Seeto, Subhrajit Bhattacharya, Joonyul Kim, Vishnu Suppiramaniam, and Elizabeth A Lipke ACS Biomater. Sci. Eng., Just Accepted Manuscript • DOI: 10.1021/acsbiomaterials.6b00226 • Publication Date (Web): 01 Sep 2016 Downloaded from http://pubs.acs.org on September 6, 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.

ACS Biomaterials Science & Engineering 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 42

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 Biomaterials Science & Engineering

Direct Production of Human Cardiac Tissues by Pluripotent Stem Cell Encapsulation in Gelatin Methacryloyl Authors: Petra Kerscher1, Jennifer A. Kaczmarek1,#, Sara E. Head1,#, Morgan Brazel1,#, Wen J. Seeto1, Joonyul Kim2, Subhrajit Bhattacharya3,Vishnu Suppiramaniam3, Elizabeth A. Lipke1,*

Affiliations: 1

Department of Chemical Engineering, 212 Ross Hall, Auburn University, Auburn, AL, 36849,

USA. 2

Department of Chemistry and Biochemistry, 179 Chemistry Building Auburn University, AL,

36849, USA. 3

Drug Discovery and Development, Harrison School of Pharmacy, 2316 Walker Building

Auburn University, Auburn, AL, 36829, USA. #

Equal contributors.

*Corresponding author, Email: [email protected]

1 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

Abstract Direct stem cell encapsulation and cardiac differentiation within supporting biomaterial scaffolds are critical for reproducible and scalable production of the functional human tissues needed in regenerative medicine and drug-testing applications. Producing cardiac tissues directly from pluripotent stem cells rather than assembling tissues using pre-differentiated cells can eliminate multiple cell-handling steps that otherwise limit the potential for process automation and production scale-up. Here we asked whether our process for forming 3D developing human engineered cardiac tissues using poly(ethylene glycol)-fibrinogen hydrogels can be extended to widely used and printable gelatin methacryloyl (GelMA) hydrogels. We demonstrate that low density GelMA hydrogels can be formed rapidly using visible light (< 1 min) and successfully employed to encapsulate human induced pluripotent stem cells while maintaining high cell viability. Resulting constructs had an initial stiffness of approximately 220 Pa, supported tissue growth and dynamic remodeling, and facilitated high efficiency cardiac differentiation (>70%) to produce spontaneously contracting GelMA human engineered cardiac tissues (GEhECTs). GEhECTs initiated spontaneous contractions on day 8 of differentiation, with synchronicity, frequency, and velocity of contraction increasing over time, and displayed developmentallyappropriate temporal changes in cardiac gene expression. GEhECT-dissociated cardiomyocytes displayed well-defined and aligned sarcomeres spaced at 1.85±0.1 µm and responded appropriately to drug treatments, including the β-adrenergic agonist isoproterenol and antagonist propranolol, as well as to outside pacing up to 3.0 Hz. Overall results demonstrate that GelMA is a suitable biomaterial for the production of developing cardiac tissues and has the potential to be employed in scale-up production and bioprinting of GEhECTs. Keywords: GelMA, ontomimetic, engineered cardiac tissue, development, hydrogel, cardiomyocyte, microenvironment 2 ACS Paragon Plus Environment

Page 3 of 42

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 Biomaterials Science & Engineering

Introduction Given that the demand for human cardiac tissue greatly exceeds the supply1, straight-forward, automatable in vitro generation of functional cardiac tissues is critical for driving advances in applications ranging from drug testing and clinical cardiac regeneration to mechanistic investigation of heart development and disease2-3. With the progress in patient-specific human induced pluripotent stem cell (hiPSC) line derivation, improvements in cellular engineering and genomic editing4, and establishment of highly efficient differentiation protocols5-6, autologous cell- and tissue-therapies possess great future medical potential, including for cardiac tissue regeneration. When engineering these developing cardiac tissues with differentiating hiPSCs, it is beneficial to mimic the physiological conditions in the native heart and enable positive interactions between cells and their surrounding extracellular matrix (ECM). Providing a suitable microenvironment to encapsulated cells can promote and support tissue formation by sustaining initial cell viability, facilitating subsequent proliferation, directing differentiation, and promoting cellular function7-9. Hydrogel matrices can provide the necessary biomimetic microenvironment while simultaneously facilitating scalable tissue production. Standard protocols for three-dimensional (3D) cardiac tissue production have been limited to the use of cardiomyocytes (CMs) obtained through pre-differentiation of stem cells10-13 or direct isolation from the heart14-15. To form engineered cardiac tissues, the CMs are then dissociated and encapsulated within a biomaterial that provides a connective tissue-like 3D microenvironment16-17. This tissue production process involves multiple cell-handling steps, negatively impacting CM viability and disrupting essential cell-cell junctions and ongoing developmental processes. If, however, pluripotent stem cells are directly differentiated within a 3D hydrogel, it is possible to generate developing and functional cardiac tissues with a high CM

3 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

yield and only one cell-handling step18. This ability to directly produce 3D cardiac tissues from hiPSCs in an ontomimetic (ontogeny-mimicking) manner has been successfully demonstrated using the biosynthetic material poly(ethylene glycol)-fibrinogen (PEG-fibrinogen)18 as the support matrix for hiPSC encapsulation and differentiation, which we called 3D developing human engineered cardiac tissues (3D-dhECTs). Understanding the potential of additional hydrogel materials to enable direct, ontomimetic cardiac tissue production is important in leveraging this approach for a range of future applications7. Collagen is an abundant ECM protein in the adult human heart19-20 and, therefore, may possess potentially beneficial properties as a matrix material for supporting temporal differentiation of hiPSCs into contracting CMs. However, collagen alone cannot form mechanically tunable and photocrosslinkable hydrogels and is associated with significant batchto-batch variability21. Therefore, to mimic properties of the native heart ECM while improving control over material properties, we selected gelatin methacryloyl (GelMA) for this study and tested whether it could support the growth, differentiation, and function of stem cell-derived cardiomyocytes (SC-CMs). GelMA is a degradable hydrogel composed of gelatin (denatured collagens) that has been tunably substituted with methacryloyl groups; different degrees of methacryloyl group substitution can be achieved to obtain a range of different structural and functional characteristics of the hydrogel, including cell-responsiveness from gelatin and polymerization capabilities from methacryloyl groups22. In particular, GelMA can be photocrosslinked to form a relatively soft hydrogel, which we have found to be advantageous for hiPSC survival and maintenance23-24. Here we report for the first time the successful production of GelMA human engineered cardiac tissues (GEhECTs) by direct hiPSC encapsulation and differentiation within soft GelMA

4 ACS Paragon Plus Environment

Page 5 of 42

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 Biomaterials Science & Engineering

hydrogels. We first synthesized GelMA with a methacryloyl group substitution degree of 22% and then produced acellular GelMA hydrogels by visible-light photocrosslinking to study the enzymatic hydrogel degradation. Cellular GelMA hydrogels were formed by incorporating hiPSCs within the liquid GelMA precursor solution followed by photocrosslinking for less than one minute. HiPSC viability after encapsulation was maintained and cardiac differentiation efficiency was validated by flow cytometry. First isolated areas of spontaneous contraction were consistently observed on day 8 of differentiation, with uniformly contracting tissues forming by day 14. Cells remodeled their 3D microenvironment and became more elongated, with larger and more elongated cell nuclei, over time. The proliferating and differentiating cells formed a connected tissue, allowing for a temporal increase in frequency and velocity of contraction and developmentally-appropriate temporal changes in gene expression. GEhECT-differentiated CMs were evaluated by α-sarcomeric actinin staining with subsequent sarcomere spacing and alignment quantification. CM functionality was verified by appropriate response to drug treatments and outside pacing. These results showed that hiPSCs can be directly differentiated in GelMA hydrogels using a straight-forward, single cell-handling approach to produce GEhECTs.

METHODS HiPSC culture IMR-90 Clone 1 and 19-9-11 hiPSCs were purchased from WiCell and cultured on hESC qualified Matrigel (BD Biosciences) in mTeSR-1 medium (Stem Cell Technologies). HiPSCs were passaged using Versene (Life Technologies) and mTESR-1 medium supplemented with 5 µM ROCK inhibitor (RI, Y-27632, Stem Cell Technologies).

5 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

PDMS mold preparation All chemicals, unless otherwise noted, were purchased from Sigma-Aldrich. A precursor solution for the polydimethylsiloxane (PDMS) molds was created as previously described18. Shortly, SLYGARD 184 silicone elastomer curing agent and SLYGARD 184 elastomer base (1:10 ratio, Dow Corning Corporation) were mixed, cured at 70 °C between two glass slides spaced at approximately 600 µm. PDMS was peeled off and a cork borer was used to create circular holes (4 mm in diameter). PDMS molds (650 µm thick) were soaked in 70% ethanol and exposed to UV-light. The mold was transferred to the bottom of a well of a 6-well plate prior to the encapsulation process.

Gelatin methacryloyl (GelMA) synthesis and precursor preparation Gelatin methacryloyl (GelMA) was synthesized following previous protocols22,

25

with

modifications. Briefly, gelatin (Type B, bovine) was mixed at 5% (w/v) into phosphate buffered saline (PBS, Gibco) at 60 °C with constant stirring until fully dissolved. Methacrylic anhydride (MA) was slowly added until the target concentration was reached (15% w/v) and reacted at 60 °C for 2 h. The reaction was stopped with PBS; gelatin methacryloyl was dialyzed for seven days and lyophilized for five days. Lyophilized GelMA was dissolved in deuterium oxide (Fisher Scientific) for NMR analysis. 1HNMR spectra were collected using a Bruker NMR spectrometer. Before integration, phase and baseline corrections were applied to ensure accurate methacryloyl group addition calculations. To prepare the precursor solution, lyophilized GelMA was re-dissolved into PBS at 80 °C at a concentration of 15mg/ml and combined with 1.5v/v% triethanolamine (TEOA), 3.96 µl/ml N-vinyl pyrrolidone (NVP) and 10 mM eosin Y (Fisher Scientific) photoinitiator26. To assess the

6 ACS Paragon Plus Environment

Page 7 of 42

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 Biomaterials Science & Engineering

ability for GelMA hydrogels to degrade in the presence of an enzyme, acellular GelMA hydrogels were incubated in collagenase solution27-29. For these degradation studies, acellular GelMA hydrogels (20 µl) were crosslinked for 60 s and incubated in PBS or 5 U/ml collagenase Type 2 (Worthington) at 37 °C. Hydrogel weights were recorded at 0, 3, and 6 h (n = 3 hydrogels).

Fabrication of GelMA human engineered cardiac tissues (GEhECTs) To form GelMA tissues, hiPSCs were dissociated using Versene, centrifuged, and re-suspended at 26 ± 4.1 x 106 hiPSCs/ml GelMA. GelMA cell suspension precursor solution (8 µl) was pipetted into one circular PDMS mold and photocrosslinked using visible light (203 mW/cm2) for 40 s (Fig. 1a). After crosslinking, the PDMS mold was removed and newly formed tissues were suspended in mTeSR-1 medium with 5 µM RI and cultured for 24 h (day -3, three tissues per well). For the next two days, (days -2 and -1) mTeSR-1 medium was changed daily. Cardiac differentiation was initiated on the third day after encapsulation (day 0). Three days after hiPSC encapsulation (day 0), cardiac differentiation was initiated using a well-established protocol5, 18. MTeSR-1 medium was replaced with 4 ml per well of RPMI/B27 without insulin (RPMI/B27-Ins, Invitrogen) supplemented with 12 µM CHIR99021 (Stem Cell Technologies); 24 h later (day 1), the medium was replaced with 4 ml RPMI/B27-Ins. After 48 h (day 3), half of the old media was combined with an equal amount of fresh RPMI/B27-Ins supplemented with 5 µM IWP2 (Stem Cell Technologies). On day 5, the media was exchanged for 4 ml of RPMI/B27-Ins, followed by a change with 4 ml RPMI/B27 (Invitrogen) on day 7. RPMI/B27 medium was replaced every three days thereafter.

7 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

Parallel plate mechanical compression testing Acellular and cellular (IMR90-1 hiPSCs) GelMA hydrogels were photocrosslinked for 40 s, cultured, and tested for their compressive mechanical properties 24 h after tissue formation (day -2) and on days 7 and 19 of differentiation using a micron-scale mechanical testing system (Microsquisher, CellScale)30. The force was calculated via cantilever beam deflection in response to user-defined displacement. All samples were tested in PBS at 35 ºC. The cantilever beam was composed of Tungsten (modulus = 411 GPa) with a diameter of 558.5 µm. Stressstrain characteristics of acellular and cellular tissues were obtained and the elastic modulus was directly obtained within the linear 5-20% strain regime (n = 3 samples).

Tissue dissociation and re-plating For assays requiring dissociated cells, tissues were incubated in a dissociation solution containing collagenase type 2 (1 mg/mL, Worthington) at 37 °C for 2 h. The dissociation solution also contained 120 mM NaCl, 5.4 mM KCl, 5 mM MgSO4, 5 mM Na-pyruvate, 20 mM glucose, 20 mM taurine, and 10 mM HEPES (pH 6.9) supplemented with 30 µM CaCl2. After incubation, the tissues were centrifuged and re-suspended in 0.25% trypsin (EDTA, Mediatech) and incubated at 37 °C for 5 min. Finally, the cells were singularized, centrifuged, resuspended in RPMI20 medium (RPMI 1640 medium with 20% FBS, Atlanta Biologicals) with 5 µM RI and plated on fibronectin (25µg/ml) coated PDMS-coated glass coverslips.

Viability and immunofluorescent visualization A LIVE/DEAD ® viability kit (Molecular Probes) was used to assess hiPSC viability one day after encapsulation (day -2). Staining was performing following manufacturer’s instructions;

8 ACS Paragon Plus Environment

Page 9 of 42

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 Biomaterials Science & Engineering

fluorescent Z-stacks were taken randomly within each tissue (n = 3 tissues). The viable and dead cells in acquired Z-stack slices were manually counted to determine percent viable cells. LIVE staining was also performed following cardiac differentiation (day 8, n = 4 tissues). Protein expression within GelMA tissues was characterized through immunofluorescence staining. Whole GEhECTs on day 0 and day 22 were stained with Alexa Fluor 568 phalloidin (actin filaments, Invitrogen) and both GEhECT-dissociated CMs (day 22) attached to fibronectin coated PDMS-coated glass coverslips and GEhECTs (day 23) were immunostained for the cardiac marker α-sarcomeric actinin (αSA, Mouse IgG1) or cardiac troponin T (cTnT, Mouse IgG1, Thermo Scientific) and for connexin 43 (Cx43, Rabbit IgG). The cells were fixed using 4% paraformaldehyde (Electron Microscopy Sciences) or 50/50 acetone/ethanol for Cx43 at -20 °C, permeabilized with PBS-T (PBS with 1% bovine serum albumin (BSA) and 0.2% Triton X-100), and blocked with 3% FBS in PBS. The samples were then successively incubated in primary (1:200) and secondary (Alexa Fluor 568 and 488, 1:200) antibodies for 24 h at 4 °C. All samples were counterstained with 4’ 6-diamidino-2-phenylindole (DAPI, Molecular Probes). Fluorescently labeled samples were visualized with a Nikon A1si confocal microscope. GEhECT cells’ nuclear circularities and cross-sectional areas were analyzed on day 0 and day 22 using ImageJ by manually outlining DAPI-stained nuclei (n = 30 nuclei per time-point). Sarcomere spacing and alignment from dissociated GEhECT samples were analyzed using standard ImageJ plugins. Sarcomere spacing was determined by drawing a perpendicular line through visible sarcomeres and obtaining its intensity profile to observe the intersecting sarcomere location (n = 20 CMs)18. Sarcomere alignment was determined by first drawing a perpendicular line through multiple sarcomeres within each cell. Then, a second line was created

9 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

overlaying one sarcomere at a time. The angle between the perpendicular line and the sarcomere line was recorded for at least five sarcomeres per cell (n = 36 CMs).

Early GEhECT growth In order to track early tissue growth, phase contrast images of whole tissues were acquired daily from day -2 to day 7 of differentiation to assess change in lateral tissue area (Ti Eclipse, Nikon equipped with an Andor Luca S camera). Tissue edges were manually tracked and the lateral surface area of whole tissues was quantified using standard plugins in ImageJ (n = 3 tissues). The change in tissue thickness was measured on days -2, 7 and 19 using the Microsquisher system (CellScale) which facilitates acquisition of side-view images.

Contraction analysis of GEhECTs Contractions of whole GEhECTs were characterized using motion tracking software31. Videos of spontaneously contracting GEhECTs were recorded on days 14, 17, and 40 (n = 3 tissues per time point). Every video frame was converted into a separate image and the motion between images was calculated. The DataEvaluation interface was used to generate a plot of average maximum contraction and relaxation velocities and to determine the frequency of contraction.

Flow cytometry On day 27, cardiac tissues were dissociated as described above. Instead of re-plating, the cells were resuspended in 4 % paraformaldehyde and incubated at 25 °C for 20 min. The cells were then centrifuged, 90% cold methanol was added and cells were incubated at 4 °C for 15 min. Fixed cells were blocked with 5% BSA in PBS for 5 min, washed, and labeled by incubation

10 ACS Paragon Plus Environment

Page 11 of 42

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 Biomaterials Science & Engineering

with 100 µl of primary antibody (cTnT, Thermo Scientific; Ki67, Abcam; CD90, Abcam) diluted in 0.5% BSA and 0.1% Triton X-100 PBS. After washing, 100 µl of secondary antibody (Alexa Fluor 488/647, Invitrogen) diluted in 0.5% BSA and 0.1% Triton X-100 in PBS was added. The cells were again washed and analyzed on a BD Accuri C6 (BD Biosciences).

Reverse transcription quantitative PCR (RT-qPCR) RNA was isolated with a Nucleospin RNA kit (Macherey-Nagel) from day 10, 22, and 73 GEhECTs (n = 3 different tissues per time point) and its concentration was determined by NanoDrop 1000 UV-spectrophotometer (ThermoFisher Scientific). SuperScript III Platinum One-Step RT-qPCR kit (ThermoFisher Scientific) was used to perform RT-qPCR with 25 ng of sample in total 12.5 µl reaction volume/ with one cycle at 50 °C (15 min), one cycle at 95 °C (3 min), 45 cycles at 95 °C (15 s) and 55 °C (30 s). Gene expression was determined by using Taqman probes (P) in conjunction with forward (F) and reverse (R) primers. MLC2v F: GGGCGGAGTGTGGAATTCTT, FAM/AGTGCTGGG

R:

R:

/3IABkFQ/;

R:

/56-

αMHC

F:

P:

/56-

/3IABkFQ/;

βMHC

CAGGCACGAAGACATCCTTCT,

/ZEN/GCAAGTCAGAGAAGG

TGAGCAGTCTGCCTTTCGTT,

P:

TTGCTTGGCACCAATGTCAC,

/ZEN/TGGATGAGGCAGAG

CACAGCCATGGGAGATTCGG, FAM/CCTACCTGC

CCCGGCTCTCTTCTTTGCTT,

/ZEN/TCCTTTCCACCAT

ACCAACCTGTCCAAGTTCCG, FAM/AGCATGAGC

R:

/3IABkFQ/;

CCAGAAGCGCACATGAGAGA,

F:

P:

/56-

Cx43

F:

P:

/56-

FAM/ACACTCAGC/ZEN/AACCTGGTTGTGAAA/3IABkFQ/.

11 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

Electrophysiological characterization and drug-response Day 30, GEhECTs were dissociated with collagenase (as described above) and plated on a S2 type MEA200/30-Ti-gr (Multichannel Systems) in RPMI20 + RI. After 24 h, MEA was perfused with Tyrode’s solution (1.8 mM CaCl2, 5 mM glucose, 5 mM HEPES, 1 mM MgCl2, 5.4 mM KCl, 135 mM NaCl, and 0.33 mM NaH2PO4, pH 7.4) at 37 °C. The MEA setup contained a Multichannel Systems 1060-Inv-BC amplifier and field potential recordings were acquired at a sampling frequency of 10 kHz. GEhECT-dissociated CMs were exogenously paced from 0.5 – 3.0 Hz using a rectangular pulse pattern at ±4,000 mV and pulse duration of 10 ms. The β-adrenergic agonist isoproterenol (Molecular Devices) and antagonist propranolol (Molecular Devices) were introduced to spontaneously contracting GEhECTs and their drugresponse was analyzed using video analysis. Isoproterenol (1 µM) was first added to GEhECTs with a subsequent addition of 1 µM propranolol. After that, drugs were washed out and baseline contraction frequencies were achieved. GEhECT-dissociated CMs from both IMR90-1 and 19-911 hiPSCs were analyzed. To visualize spontaneously propagating calcium waves across whole GEhETCs, tissues were first incubated with the calcium indicator Rhod-2 (Invitrogen) and then recordings were acquired using optical mapping (n = 3 tissues). To prepare the calcium indicator solution, 5 µl of 100 µM Rhod-2 and 5 µl of Pluronic/ml were added to 1 ml Tyrode’s solution containing blebbistatin (50 µM). GEhECTs were incubated in this solution at room temperature for 2 hrs and then rinsed with Tyrode’s solution (37° C) over 15 minutes. Videos of propagating calcium waves were recorded and processed as previously described32.

12 ACS Paragon Plus Environment

Page 13 of 42

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 Biomaterials Science & Engineering

Statistics Minitab 17 was used to analyze the reported results and the mean ± SD was reported of all replicates. After checking for normal distribution, one-way analysis of variance (ANOVA) with Turkey’s test was used for all studies with equal variance and equal sample size. If equal variance did not apply, the Games-Howell test was used. P < 0.05 was considered statistically significant.

Results GelMA forms photocrosslinked hydrogels that enzymatically degrade In the present study we show that GelMA can be used to encapsulate hiPSCs and support 3D differentiation to form uniformly contracting cardiac tissues. GelMA hydrogels have previously been successfully used to encapsulate and culture multiple cell types to form engineered tissues33-36. Prior work has shown that low crosslinking density GelMA (~20% methacrylation) forms soft hydrogels, while high crosslinking density GelMA (~80% methacrylation) forms stiff hydrogels22. Based on our previous experience with PEG-fibrinogen, GelMA with a low methacryloyl substitution degree was chosen for this study to produce soft hydrogels and successfully used to create GelMA human engineered cardiac tissues (GEhECTs). Synthesized GelMA was lyophilized to form a sponge-like solid. Using NMR, a substitution percentage of 22% was calculated (Fig. 1b) by dividing methacryloyl group peaks (located at 5.3 ppm and 5.5 ppm) and gelatin’s free amine group peak. Initial studies confirmed that GelMA hydrogels could be successfully formed by visible light photocrosslinking using the photoinitiator eosin Y. Hydrogel stability and temporal enzymatic degradation was validated; chemically modified gelatin can be enzymatically

13 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

degraded by cells, enabling cell migration and proliferation within the provided matrix to form physiologically relevant tissues. To demonstrate that our GelMA was enzymatically degradable, acellular hydrogels were incubated in either PBS or collagenase solution at 37 °C and temporal changes in mass were quantified. Whereas hydrogels in PBS alone maintained their mass after the expected initial increase as a result of swelling, hydrogels in collagenase decreased in mass over the first 6 h, suggesting successful, temporal GelMA degradation (Fig. 1c). Within 24 h, hydrogels in collagenase solution degraded completely; although the time course differs, this illustrates the feasibility of cell-mediated enzymatic degradation of the GelMA, a critical feature for enabling encapsulated cells to remodel their microenvironment to form a continuous tissue in subsequent experiments.

14 ACS Paragon Plus Environment

Page 15 of 42

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 Biomaterials Science & Engineering

FIG. 1. GelMA hydrogels were successfully used to produce GEhECTs. (a) HiPSCs were collected and combined with a liquid GelMA precursor solution; this mixture was then added into a circular PDMS mold and photocrosslinked for 40 s using visible light. (b) NMR spectrum showed synthesized GelMA with a methacryloyl substitution degree of 22% (signals of methacrylate double bonds at 5.3 and 5.5 ppm). (c) Photocrosslinked GelMA hydrogels initially swelled in the presence of PBS and degraded when incubated in collagenase, with over 50% mass loss by 6 h (n = 3 hydrogels).

15 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

GelMA hydrogels support hiPSC survival, growth, and cardiac differentiation To form cell-laden GelMA hydrogels, hiPSCs were encapsulated in GelMA hydrogels using visible light photocrosslinking for less than one minute in a manner analogous to formation of acellular GelMA hydrogels. Mechanical testing of acellular and cellular hydrogels verified that our low crosslinking density GelMA hydrogels formed soft constructs (Fig. 2a); GelMA stiffness has been shown to play an important role in successful cell encapsulation and function22, 37-38

. The stiffness of cellular hydrogels appeared to initially decrease with time in culture and

commencement of cardiac differentiation, with elastic moduli decreasing from 219 ± 140 Pa (n = 3 tissues) on day -2 (24 h after encapsulation) to 90 ± 8.4 Pa (n = 3 tissues) on day 7 of differentiation. As cardiac differentiation progressed, GEhECT elastic modulus appeared to increase, with day 19 GEhECTs having an elastic modulus of 248 ± 109 Pa (n = 3 tissues). Cellular hydrogels were softer than acellular GelMA hydrogels, which had an elastic moduli of 660 ± 307 Pa (Fig. 2a, n = 4 hydrogels). Following encapsulation, hiPSCs were uniformly distributed (Fig. 2b) and remained viable within GelMA hydrogels (>75% survival, Fig. 2c). Three days post-encapsulation (day 0), hiPSCs had begun to grow beyond the original GelMA hydrogel boundaries, resulting in an increase in lateral surface area (Fig. 2d). Even with this enlargement in tissue size, hiPSCs maintained a round cell morphology, visualized by phalloidin staining (F-actin, Fig. 2e), through initiation of differentiation (day 0).

16 ACS Paragon Plus Environment

Page 17 of 42

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 Biomaterials Science & Engineering

FIG. 2. GelMA encapsulated hiPSCs form soft, cell-laden hydrogels with hiPSCs surviving the encapsulation process to produce GEhECTs. (a) Acellular GelMA hydrogels exhibited an elastic modus of 660 ± 307 Pa; in comparison, cellular GelMA tissues 24 h after encapsulation (day -2) were softer, with an elastic modulus of 219 ± 140 Pa. Elastic modulus of the GEhECTs decreased to 90 ± 8.4 Pa by day 7. Later in the cardiac differentiation process (day 19), GEhECT elastic modulus was again higher at 248 ± 109 Pa. n = 3 tissues per group. Mean ± s.d. ANOVA P < 0.05, * vs. GelMA hydrogel. (b) Phase contrast image showed circular 3D cell-laden hydrogel 24 h after encapsulation (day -2) with (c) the majority of hiPSCs surviving the encapsulation process (green = viable, red = dead). (d) Three days after encapsulation (day 0), hiPSCs formed growing clusters, (e) displaying a round cell morphology (F-actin filaments 17 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

visualized by phalloidin, green) with cells expanding beyond the original boundary of the hydrogel.

Cell proliferation and growth both prior to and during cardiac differentiation was most pronounced at the GEhECT edges (Fig. 2, Fig. 3a), resulting in a darkening of phase contract images as a result of higher cell density and an increase in tissue surface area. From day -2 to day 5 of differentiation, the tissue surface area increased by 13.1 ± 1.7%, followed by a decrease of 2.8% between day 5 and 7 (Fig. 3b, n = 3 tissues). This observed decrease in tissue area was followed by initiation of GEhECT spontaneous contraction; first spontaneous contraction within GEhECTs was consistently detected by day 8 of differentiation (Supplemental Movie 1). Tissue thickness remained constant between day -2 and day 7, but increased significantly by day 19 of GEhECT culture (Fig. 3c); viability remained high (Supplemental Figure 1). Flow cytometry on day 27 of differentiation verified successful cardiac differentiation of 3D encapsulated hiPSCs. GEhECTs contained 70.4 ± 1.6% CMs (cTnT+), of which were 10.6 ± 2.6% proliferating CMs (cTnT+, Ki67+), and 8.4 ± 0.3% fibroblasts (CD90+, Fig. 3d, n = 3 tissues). Immunostaining of whole GEhECTs also showed expression of cTnT along with the gap junction protein Cx43 (confocal z-stack, Supplemental Movie 2). Ontomimetic changes in cardiac and functional gene expression verified temporal maturation of GEhECTs. Expression of the early cardiac marker MLC2v was significantly upregulated on days 20 and 70 compared to day 10, displaying a downward trend between days 20 and 70 (Fig. 3e). The first myosin heavy chain subunit, αMHC, was significantly downregulated from days 10 and 20 to day 70 (Fig. 3f). The more prominent second myosin heavy chain subunit, βMHC, showed a significant upregulation between day 10 and days 20 and 70, suggesting temporal maturation over extended

18 ACS Paragon Plus Environment

Page 19 of 42

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 Biomaterials Science & Engineering

culture time (Fig. 3g). The functional gene, Cx43, showed an increasing trend between day 10 and day 70, suggesting a positive trend towards maturing cell-cell coupling (Fig. 3h).

FIG. 3. Encapsulated hiPSCs grew and differentiated to form a continuous tissue over time. (a) After the onset of cardiac differentiation (day 0), cells continued to grow within the original GelMA construct, forming a more connected tissue over time. (b) Circular surface area of tissues increased by 13.1 ± 1.7% by day 5 and decreased somewhat thereafter to form a dense tissue by

19 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

day 7, immediately prior to the onset of contraction, observed consistently on day 8. n = 3 tissues per group. Mean ± s.d. ANOVA P < 0.05, # vs. Day -2, * vs. Day 0. (c) Tissue thickness was similar on day -2 and day 7, but increased by day 19 of differentiation. n = 3 tissues per group. Mean ± s.d. ANOVA P < 0.05, * vs. Day 7. (d) On day 17, GEhECTs contained approximately 70.4 ± 1.6% CMs (cTnT+), of which 10.6 ± 2.6% were proliferating (cTnT+, Ki67+), and 8.4 ± 0.3% fibroblasts (CD90+) (n = 3 tissues). (e) Gene expression of the cardiac markers MLC2v, αMHC, and βMHC on day 10, 20, and 70 GEhECTs showed temporal changes expected of maturing SC-CMs during culture. Cx43 gene expression in GEhECTs trended upward suggesting developing cell-cell coupling (n = 3 GEhECTs). Mean ± s.d. ANOVA * P < 0.05.

Uniformly contracting GEhECTs showed increased frequency and velocity of contraction over time GEhECTs consistently displayed spontaneously contracting areas initially on day 8 of differentiation, with more areas of contraction initiating over time, resulting into uniformly contracting tissues by day 14 (Supplementary Movie 3). Once individual areas joined and contracted uniformly, the frequency and average maximum contraction and relaxation velocity increased over time (Fig. 4, Supplementary Movie 4), with tissues visually deforming between contracted and relaxed states. Sarcomeres as well as the gap junction protein Cx43 were visualized through immunostaining of whole GEhECTs (day 23, Fig. 4a). The frequency of GEhECT spontaneous contraction increased significantly between day 14 (0.3 Hz) and day 40 (0.9 Hz, Fig. 4b). The maximum contraction velocity increased from 120 ± 48 µm/s (day 14) to 597 ± 127 µm/s (day 40, Fig. 4c). Similarly, the maximum relaxation velocity increased from

20 ACS Paragon Plus Environment

Page 21 of 42

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 Biomaterials Science & Engineering

86 ± 58 µm/s (day 14) to 564 ± 107 µm/s (day 40, Fig. 4d), which correlates with the visual observance of stronger tissue contraction over time (Supplementary Movies 3, 4).

FIG. 4. Frequency of spontaneous contraction and contractile velocity increased with GEhECT culture time. (a) Whole tissue staining of day 23 GEhECTs displayed striation (αSA, red) and gap junctions (Cx43, green). (b) GEhECTs formed uniformly contracting tissues by day 14, with frequencies of spontaneous contraction increasing significantly between days 14 and 40. (c, d) In addition to the increase in contraction frequency, average maximum contraction and relaxation velocities also increased from day 14 to day 40 of differentiation. This increase in contraction and relaxation velocity corresponded with visual observations of stronger contracting tissues (Supplementary Movies 3, 4). (e) Representative analyzed traces showing progression over time. n = 3-5 tissues. Mean ± s.d. ANOVA P < 0.05, * vs. Days 14, 17.

21 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

Cellular remodeling and development of CMs in GEhECTs Observed changes in actin filament morphology as well as nuclei morphology and size between day 0 and day 22 of differentiation confirmed intracellular as well as microenvironmental remodeling, thereby contributing to the formation of a connective and uniformly contracting tissue. Confocal images of fluorescently stained cell-laden tissues on day 0 of differentiation (Fig. 2e) showed encapsulated cells with round (0.8 ± 0.1 circularity, Fig. 5b) and small (91.2 ± 30.5 µm2, Fig. 5c) cell nuclei. After 22 days, GEhECT cells displayed elongated actin filaments (Fig. 5a) with more elongated (0.7 ± 0.1 circularity, Fig. 5b) and larger (149 ± 52.4 µm2, Fig. 5c) nuclei of 3D cultured cells (n = 30), an expected trend during stem cell differentiation39. In addition to flow cytometry and video microscopy of visually contracting GEhECTs, immunofluorescence verified successful cardiac differentiation within GelMA hydrogels, showing defined and aligned sarcomeres. Contracting GEhECTs were successfully dissociated, re-adhered, and continued to synchronously contract (day 22, Supplementary Movie 5). While maintained in 2D culture, GEhECT-dissociated CMs displayed well-defined (Fig. 5d) and aligned (Fig. 5e, n = 36) sarcomeres, spaced at 1.85 ± 0.1 µm (n = 20 cells).

22 ACS Paragon Plus Environment

Page 23 of 42

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 Biomaterials Science & Engineering

FIG. 5. GEhECT CMs remodel their microenvironment and show morphological differences expected from differentiating cells. (a) Day 22 tissues contained areas of high cell density primarily on the edges of GEhECTs where cells showed elongated F-actin filaments (phalloidin, green). (b, c) Day 0 (hiPSCs) encapsulated cells displayed more circular and smaller nuclei compared to day 22 (CMs) cells. n = 30 nuclei per condition. Mean ± s.d. ANOVA P < 0.05, * vs. HiPSC. (d) Following tissue dissociation, GEhECT-dissociated CMs continued to spontaneously contract (Supplementary Movie 5) and locally self-aligned with each other, displaying defined sarcomeres (αSA, red) with (e) high intracellular alignment (n = 36).

GEhECT CMs responded to extrinsic stimuli Whole GEhECTs supported spontaneous calcium wave propagation (Supplemental Movie 6). To evaluate the potential of these GEhECT-derived SC-CMs to be used for future clinical applications or drug testing, the functionality of CMs was evaluated by pharmacological and electrical stimulation. Spontaneously contracting whole GEhECTs were exposed to the β-

23 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

adrenergic agonist isoproterenol (1 µM), which caused an expected increase in frequency of contraction by ~0.2 Hz. Subsequently, propranolol (1 µM), a β-adrenergic antagonist, was added to GEhECTs which caused the frequency of contraction to slow down by ~0.36 Hz (n = 3 tissues, Fig. 6a). Although GEhECTs initial frequency of contraction varied between tissues, changes caused by the administered drugs were consistent, even with different hiPSC lines. GEhECTs were also successfully dissociated and plated onto a MEA chip (Fig. 6b). GEhECTdissociated CMs were electrically paced from 0.5 Hz to 3.0 Hz; field potential recordings showed 1:1 correspondence to outside pacing up to 3.0 Hz.

FIG. 6. Cardiac tissues responded to drug treatment and outside pacing. (a) GEhECTs produced with IMR90-1 and 19-9-11 hiPSC lines responded to the β-adrenergic agonist isoproterenol (Iso) 24 ACS Paragon Plus Environment

Page 25 of 42

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 Biomaterials Science & Engineering

and β-adrenergic antagonist propranolol (Prop); 1 µM Iso increased the frequency of spontaneous contraction, the subsequent addition of 1 µM Prop slowed down the contraction. (b) Day 30 GEhECT-dissociated CMs (from IMR90-1 hiPSCs) were cultured on MEA (Supplementary Movie 7) and (c) extrinsically paced. GEhECT-dissociated CMs responded to outside pacing frequencies up to 3.0 Hz.

Discussion In this study we show that GelMA is a suitable biomaterial for hiPSC encapsulation, culture, and cardiac differentiation to form uniformly contracting GelMA human engineered cardiac tissues (GEhECTs). GelMA is an ECM-based biomaterial that enables cell-responsive hydrogel degradation and subsequent remodeling of the cellular microenvironment. Soft GelMA hydrogels surrounding encapsulated hiPSCs were produced using GelMA with a methacryloyl substitution degree of 22%, which results in the formation of gels with a low crosslinking density. The elastic modulus of GelMA hydrogels can be altered by changing the methacryloyl substitution degree during synthesis and the exposure time and light intensity during photocrosslinking, in addition to being influenced by temperature40. During compression testing, acellular GelMA hydrogels showed a higher elastic modulus than GelMA hydrogels containing encapsulated hiPSCs, suggesting that acellular hydrogels are stiffer than their cellular counterparts, but still remain softer than 1 kPa. GelMA having a high degree of methacryloyl substitution, which would have resulted in gels with increased elastic modulus, was not chosen for this project based on its slower degradation rate and reported inhibition of cell spreading41. Comparing the mechanical properties to other, similar methacryloyl substitution degree GelMA hydrogels, the photocrosslinked GelMA hydrogel for this study was softer22. This could be a

25 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

result of differences in the photoinitiator employed and the crosslinking time. Although lineage specification is influenced by matrix stiffness37, microenvironmental cues exerted on cells by biomaterials exploited in 2D or 3D culture22 have to be carefully distinguished and evaluated separately. Identifying suitable biomaterials to produce developing human cardiac tissues using only a single cell-handling step is challenging; success depends on a variety of parameters, including but not limited to mechanical properties, rate and mode of material degradation, crosslinking density, initial cell-seeding density, and potential to support cell maintenance and differentiation processes. Biomimetic material design to facilitate cardiac tissue production and cardiac regeneration is the subject of significant research efforts7 and incorporation of additional biomimetic cues may further enhance GEhECT function and maturation18, 42. In the past, natural materials such as Matrigel, collagen, or fibrin have been used to create cardiac tissues43-46, recapitulating aspects of the native ECM composition. In order to better control and tune the properties of hydrogels, purely synthetic materials have also been used for 3D cell culture47. However, these systems frequently require the addition of bioactive sites necessary for the survival, adhesion, and growth of cells, as well as for the cell-responsive and temporal degradation of the material, and have generally not been amendable to formation of robust engineered cardiac tissues. Based on the drawbacks of purely synthetic or natural biomaterials alone, hybrid and functionally modified natural biomaterials that possess suitable mechanical, biological, and physical characteristics necessary for cell survival, attachment, and growth have unique advantages and are becoming more widely used in tissue engineering, bioprinting, and microfluidic applications48-49. Hybrid and functionally modified natural biomaterials, including

26 ACS Paragon Plus Environment

Page 27 of 42

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 Biomaterials Science & Engineering

PEG-fibrinogen and GelMA, are therefore advantageous for the direct production of developing human cardiac tissues, as reported here. We recently showed that PEG-fibrinogen is a favorable biomaterial for hiPSC encapsulation and ontomimetic cardiac differentiation18. PEG-fibrinogen has previously been tested for use in a range of biomedical applications, including cardiac regeneration50-51 and in clinical trials for cartilage repair52. However, although PEG-fibrinogen has high potential, it has not yet been assessed in bioprinting applications, which would facilitate the transition from bench top research to biomanufacturing and advancements in cell-therapy. GelMA has also been used alone and in combination with other materials40, 53 to successfully produce tissues for a range of applications. In addition, methacryloyl-containing biomaterials, including GelMA, have been used for bioprinting 3D tissues40,

54

, a technology with high

potential for advancing the field of biomanufacturing55. Bioprinting not only allows for fast production of 3D cell-laden hydrogels, but also for increased reproducibility and production speed. Experimentation with printing of hiPSCs has been challenging due to the fragile nature of this cell type. Nevertheless, pluripotent stem cells have been successfully printed with high viability and maintenance of pluripotency56-57. Printing hiPSCs in combination with a hybrid or naturally modified biomaterial followed by subsequent cardiac differentiation has not yet been investigated, but, based on our results, could have high potential for the fabrication of engineered heart tissues. To manufacture the high number of cells necessary for cell-therapy and other regenerative applications, microcarriers or hydrogel encapsulation to support cell attachment and survival can be advantageous in bioreactor setup; GelMA hydrogels containing hiPSCs could be fabricated to meet this need. By employing only a single cell-handling step and direct differentiation approach, our reported GEhECT platform has

27 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 42

advantages for use in both bioprinting and scalable cell production to obtain functional human cardiac tissues for drug testing and regenerative medicine applications. CM function and mechanical contraction are physical characteristics of healthy cardiac tissues that can be monitored easily throughout in vitro development and maturation. Without the necessity of tissue dissociation, changes in frequency, contraction and relaxation velocities of spontaneously contracting GEhECTs were monitored and showed significant increases with culture time. GEhECTs were maintained in suspension without physical constraints or applied load. Contraction and relaxation velocities at early time points (day 14) were of the same magnitude as other human pluripotent stem cell (hPSC)-CMs cultured 2D monolayers on solid substrates31. Starting on day 17, as CMs within GEhECTs developed more robust cell-cell connections, tissues physically contracted faster and with greater tissue-scale deformation; this continual increase in contraction and relaxation velocities progressed through at least day 40 of differentiation. Differences in frequency of contraction can be linked to the level of maturation of cultured cells and also differs between tissues in suspension culture versus those mechanically anchored to a substrate or other outside support, which apply an external load on contracting CMs58. Although GEhECTs increased in their frequency of spontaneous contraction with culture time, the rate of contraction was slower than other human cardiac tissues cultured in vitro with mechanical load applied by a stiff substrate18 or mechanical posts. Overall, tissues showed regular contraction patterns and uniform frequency of contraction, properties which have previously been correlated to uniform cell distribution in contracting tissue areas and cell alignment59. The in vitro functional responsiveness of these developing cardiac tissues is critical for use in future drug testing and regenerative medicine applications.

28 ACS Paragon Plus Environment

Page 29 of 42

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 Biomaterials Science & Engineering

Conclusions In recent years, significant progress has been made in the production of functional human pluripotent stem cell-derived cardiomyocytes. The importance of cell-cell and cell-material interactions created by 3D culture systems has been studied, but determining appropriate materials and favorable material properties to support hiPSC encapsulation and subsequent cardiac differentiation has been difficult. Here we have shown for the first time that the printable gelatin methacryloyl (GelMA) hydrogel biomaterial can be successfully used to encapsulate and differentiate hiPSCs to produce GelMA human engineered cardiac tissues (GEhECTs). Low density GelMA hydrogels with an elastic modulus of less than 1 kPa supported high hiPSC viability after encapsulation and temporal remodeling of the hydrogel, which facilitated tissue growth, high efficiency cardiac differentiation, and increasing frequency and velocity of spontaneous contraction over time. We also showed that GEhECT-dissociated CMs contained well-defined and aligned sarcomeres and responded appropriately to the β-adrenergic agonist isoproterenol and antagonist propranolol as well as to outside pacing frequencies up to 3.0 Hz. Our results have the possibility to be translated towards bioprinting and biomanufacturing applications, and provide a potentially advantageous platform for meeting the need to manufacture reproducible functional human cardiac tissues.

29 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 30 of 42

Acknowledgments This work was funded by the National Science Foundation (NSF-CBET-1150854) (P.K., S.E.H., E.A.L.), an Auburn University Undergraduate Research Fellowship (J.A.K.), an AUCMB/EPSCoR Summer Fellowship (P.K.), the Alabama EPSCoR Graduate Research Scholarship Program (P.K.), a Department of Education GAANN Fellowship (P200A150074) (M.B.) and the Auburn University Internal Grants Program (V.S.). We would like to thank Benjamin Samuel Selwyn Anbiah for providing assistance with optical mapping, Bryana Harris for assistance with contraction analysis, and the Ethier lab at Georgia Tech for providing access to the Microsquisher.

Supporting information Supplementary Movie 1. Localized initiation of contraction consistently occurred on day 8. Supplementary Movie 2. Confocal z-stack of day 23 GEhECT immunostained for cTnT and Cx43. Supplementary Movie 3. Contracting GEhECT on day 14 of differentiation. Supplementary Movie 4. Uniformly and visually deforming GEhECT on day 40 of differentiation. Supplementary Movie 5. Day 22 GEhECT-dissociated CMs re-adhered and spontaneously contracting. Supplementary Movie 6: Spontaneous calcium transient propagation across GEhECT on day 30 of differentiation. Supplementary Movie 7. GEhECT CMs plated and spontaneously contracting on MEA chip.

30 ACS Paragon Plus Environment

Page 31 of 42

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 Biomaterials Science & Engineering

References 1.

2.

3.

4.

5.

6.

7.

8.

9.

10.

11.

12.

Noseda, M.; Abreu-Paiva, M.; Schneider, M. D., The Quest for the Adult Cardiac Stem Cell. Circ J 2015, 79 (7), 1422-30. DOI: 10.1253/circj.CJ-15-0557. Kempf, H.; Kropp, C.; Olmer, R.; Martin, U.; Zweigerdt, R., Cardiac differentiation of human pluripotent stem cells in scalable suspension culture. Nat Protoc 2015, 10 (9), 1345-61. DOI: 10.1038/nprot.2015.089. Jeziorowska, D.; Korniat, A.; Salem, J. E.; Fish, K.; Hulot, J. S., Generating patientspecific induced pluripotent stem cells-derived cardiomyocytes for the treatment of cardiac diseases. Expert Opin Biol Ther 2015, Vol. 15, 1-11. DOI: 10.1517/14712598.2015.1064109. Hotta, A.; Yamanaka, S., From Genomics to Gene Therapy: Induced Pluripotent Stem Cells Meet Genome Editing. Annu Rev Genet 2015, 49, 47-70. DOI: 10.1146/annurevgenet-112414-054926. Lian, X.; Zhang, J.; Azarin, S. M.; Zhu, K.; Hazeltine, L. B.; Bao, X.; Hsiao, C.; Kamp, T. J.; Palecek, S. P., Directed cardiomyocyte differentiation from human pluripotent stem cells by modulating Wnt/beta-catenin signaling under fully defined conditions. Nat Protoc 2013, 8 (1), 162-75. DOI: 10.1038/nprot.2012.150. Burridge, P. W.; Matsa, E.; Shukla, P.; Lin, Z. C.; Churko, J. M.; Ebert, A. D.; Lan, F.; Diecke, S.; Huber, B.; Mordwinkin, N. M.; Plews, J. R.; Abilez, O. J.; Cui, B.; Gold, J. D.; Wu, J. C., Chemically defined generation of human cardiomyocytes. Nat Methods 2014, 11 (8), 855-60. DOI: 10.1038/nmeth.2999. Hodge, A. J.; Kerscher, P.; Dunn, D. A.; Lipke, E. A., BIOMIMETIC MATERIALS FOR CARDIAC REGENERATION. In Handbook of Biomimetics and Bioinspiration: 3 Tissue Models, 2014; pp 1037-1067. Trosper, N.; Kerscher, P.; Macadangdang, J.; Carson, D.; Lipke, E.; Kim, D.-H., Microand Nanofabrication Approaches to Cardiac Tissue Engineering. In Tissue and Organ Regeneration: Advances in Micro-and Nanotechnology, Pan Stanford Publishing: 2014; pp 725-754. Burdick, J. A.; Vunjak-Novakovic, G., Engineered microenvironments for controlled stem cell differentiation. Tissue Eng Part A 2009, 15 (2), 205-19. DOI: 10.1089/ten.tea.2008.0131. Lian, X.; Hsiao, C.; Wilson, G.; Zhu, K.; Hazeltine, L. B.; Azarin, S. M.; Raval, K. K.; Zhang, J.; Kamp, T. J.; Palecek, S. P., Robust cardiomyocyte differentiation from human pluripotent stem cells via temporal modulation of canonical Wnt signaling. Proc Natl Acad Sci U S A 2012, 109 (27), E1848-57. DOI: 10.1073/pnas.1200250109. Nguyen, D. C.; Hookway, T. A.; Wu, Q.; Jha, R.; Preininger, M. K.; Chen, X.; Easley, C. A.; Spearman, P.; Deshpande, S. R.; Maher, K.; Wagner, M. B.; McDevitt, T. C.; Xu, C., Microscale generation of cardiospheres promotes robust enrichment of cardiomyocytes derived from human pluripotent stem cells. Stem Cell Reports 2014, 3 (2), 260-8. DOI: 10.1016/j.stemcr.2014.06.002. Turnbull, I. C.; Karakikes, I.; Serrao, G. W.; Backeris, P.; Lee, J. J.; Xie, C.; Senyei, G.; Gordon, R. E.; Li, R. A.; Akar, F. G.; Hajjar, R. J.; Hulot, J. S.; Costa, K. D., Advancing functional engineered cardiac tissues toward a preclinical model of human myocardium. FASEB J 2014, 28 (2), 644-54. DOI: 10.1096/fj.13-228007.

31 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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

13.

14.

15.

16.

17. 18.

19.

20. 21.

22.

23. 24.

25.

26.

27.

Page 32 of 42

Schaaf, S.; Shibamiya, A.; Mewe, M.; Eder, A.; Stohr, A.; Hirt, M. N.; Rau, T.; Zimmermann, W. H.; Conradi, L.; Eschenhagen, T.; Hansen, A., Human engineered heart tissue as a versatile tool in basic research and preclinical toxicology. PLoS One 2011, 6 (10), e26397. DOI: 10.1371/journal.pone.0026397. Bian, W.; Badie, N.; Himel, H. D. t.; Bursac, N., Robust T-tubulation and maturation of cardiomyocytes using tissue-engineered epicardial mimetics. Biomaterials 2014, 35 (12), 3819-28. DOI: 10.1016/j.biomaterials.2014.01.045. Shapira-Schweitzer, K.; Seliktar, D., Matrix stiffness affects spontaneous contraction of cardiomyocytes cultured within a PEGylated fibrinogen biomaterial. Acta Biomater 2007, 3 (1), 33-41. DOI: 10.1016/j.actbio.2006.09.003. Masumoto, H.; Ikuno, T.; Takeda, M.; Fukushima, H.; Marui, A.; Katayama, S.; Shimizu, T.; Ikeda, T.; Okano, T.; Sakata, R.; Yamashita, J. K., Human iPS cell-engineered cardiac tissue sheets with cardiomyocytes and vascular cells for cardiac regeneration. Sci Rep 2014, 4, 6716. DOI: 10.1038/srep06716. Zhu, J.; Marchant, R. E., Design properties of hydrogel tissue-engineering scaffolds. Expert Rev Med Devices 2011, 8 (5), 607-26. DOI: 10.1586/erd.11.27. Kerscher, P.; Turnbull, I. C.; Hodge, A. J.; Kim, J.; Seliktar, D.; Easley, C. J.; Costa, K. D.; Lipke, E. A., Direct hydrogel encapsulation of pluripotent stem cells enables ontomimetic differentiation and growth of engineered human heart tissues. Biomaterials 2015, 83, 383-395. DOI: 10.1016/j.biomaterials.2015.12.011. Stoppel, W. L.; Kaplan, D. L.; Black, L. D., 3rd, Electrical and mechanical stimulation of cardiac cells and tissue constructs. Adv Drug Deliv Rev 2015. DOI: 10.1016/j.addr.2015.07.009. Caulfield, J. B.; Borg, T. K., The collagen network of the heart. Lab Invest 1979, 40 (3), 364-72. Kumar, V. A.; Taylor, N. L.; Jalan, A. A.; Hwang, L. K.; Wang, B. K.; Hartgerink, J. D., A nanostructured synthetic collagen mimic for hemostasis. Biomacromolecules 2014, 15 (4), 1484-90. DOI: 10.1021/bm500091e. Nichol, J. W.; Koshy, S. T.; Bae, H.; Hwang, C. M.; Yamanlar, S.; Khademhosseini, A., Cell-laden microengineered gelatin methacrylate hydrogels. Biomaterials 2010, 31 (21), 5536-44. DOI: 10.1016/j.biomaterials.2010.03.064. Lutolf, M. P.; Gilbert, P. M.; Blau, H. M., Designing materials to direct stem-cell fate. Nature 2009, 462 (7272), 433-41. DOI: 10.1038/nature08602. Banerjee, A.; Arha, M.; Choudhary, S.; Ashton, R. S.; Bhatia, S. R.; Schaffer, D. V.; Kane, R. S., The influence of hydrogel modulus on the proliferation and differentiation of encapsulated neural stem cells. Biomaterials 2009, 30 (27), 4695-9. DOI: 10.1016/j.biomaterials.2009.05.050. Van Den Bulcke, A. I.; Bogdanov, B.; De Rooze, N.; Schacht, E. H.; Cornelissen, M.; Berghmans, H., Structural and rheological properties of methacrylamide modified gelatin hydrogels. Biomacromolecules 2000, 1 (1), 31-8. Franco, C. L.; Price, J.; West, J. L., Development and optimization of a dualphotoinitiator, emulsion-based technique for rapid generation of cell-laden hydrogel microspheres. Acta Biomater 2011, 7 (9), 3267-76. DOI: 10.1016/j.actbio.2011.06.011. Hutson, C. B.; Nichol, J. W.; Aubin, H.; Bae, H.; Yamanlar, S.; Al-Haque, S.; Koshy, S. T.; Khademhosseini, A., Synthesis and characterization of tunable poly(ethylene glycol):

32 ACS Paragon Plus Environment

Page 33 of 42

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 Biomaterials Science & Engineering

28.

29. 30.

31.

32.

33.

34.

35.

36.

37. 38.

39.

gelatin methacrylate composite hydrogels. Tissue Eng Part A 2011, 17 (13-14), 1713-23. DOI: 10.1089/ten.TEA.2010.0666. Camci-Unal, G.; Cuttica, D.; Annabi, N.; Demarchi, D.; Khademhosseini, A., Synthesis and characterization of hybrid hyaluronic acid-gelatin hydrogels. Biomacromolecules 2013, 14 (4), 1085-92. DOI: 10.1021/bm3019856. Wade, R. J.; Bassin, E. J.; Rodell, C. B.; Burdick, J. A., Protease-degradable electrospun fibrous hydrogels. Nat Commun 2015, 6, 6639. DOI: 10.1038/ncomms7639. Kinney, M. A.; Hookway, T. A.; Wang, Y.; McDevitt, T. C., Engineering threedimensional stem cell morphogenesis for the development of tissue models and scalable regenerative therapeutics. Ann Biomed Eng 2014, 42 (2), 352-67. DOI: 10.1007/s10439013-0953-9. Huebsch, N.; Loskill, P.; Mandegar, M. A.; Marks, N. C.; Sheehan, A. S.; Ma, Z.; Mathur, A.; Nguyen, T. N.; Yoo, J. C.; Judge, L. M.; Spencer, C. I.; Chukka, A. C.; Russell, C. R.; So, P. L.; Conklin, B. R.; Healy, K. E., Automated Video-Based Analysis of Contractility and Calcium Flux in Human-Induced Pluripotent Stem Cell-Derived Cardiomyocytes Cultured over Different Spatial Scales. Tissue Eng Part C Methods 2015, 21 (5), 467-79. DOI: 10.1089/ten.TEC.2014.0283. Spearman, B. S.; Hodge, A. J.; Porter, J. L.; Hardy, J. G.; Davis, Z. D.; Xu, T.; Zhang, X.; Schmidt, C. E.; Hamilton, M. C.; Lipke, E. A., Conductive interpenetrating networks of polypyrrole and polycaprolactone encourage electrophysiological development of cardiac cells. Acta Biomater 2015, 28, 109-20. DOI: 10.1016/j.actbio.2015.09.025. Benton, J. A.; DeForest, C. A.; Vivekanandan, V.; Anseth, K. S., Photocrosslinking of gelatin macromers to synthesize porous hydrogels that promote valvular interstitial cell function. Tissue Eng Part A 2009, 15 (11), 3221-30. DOI: 10.1089/ten.TEA.2008.0545. Xiao, W.; He, J.; Nichol, J. W.; Wang, L.; Hutson, C. B.; Wang, B.; Du, Y.; Fan, H.; Khademhosseini, A., Synthesis and characterization of photocrosslinkable gelatin and silk fibroin interpenetrating polymer network hydrogels. Acta Biomater 2011, 7 (6), 2384-93. DOI: 10.1016/j.actbio.2011.01.016. Paul, A.; Hasan, A.; Kindi, H. A.; Gaharwar, A. K.; Rao, V. T.; Nikkhah, M.; Shin, S. R.; Krafft, D.; Dokmeci, M. R.; Shum-Tim, D.; Khademhosseini, A., Injectable graphene oxide/hydrogel-based angiogenic gene delivery system for vasculogenesis and cardiac repair. ACS Nano 2014, 8 (8), 8050-62. DOI: 10.1021/nn5020787. Aubin, H.; Nichol, J. W.; Hutson, C. B.; Bae, H.; Sieminski, A. L.; Cropek, D. M.; Akhyari, P.; Khademhosseini, A., Directed 3D cell alignment and elongation in microengineered hydrogels. Biomaterials 2010, 31 (27), 6941-6951. DOI: 10.1016/j.biomaterials.2010.05.056. Engler, A. J.; Sen, S.; Sweeney, H. L.; Discher, D. E., Matrix elasticity directs stem cell lineage specification. Cell 2006, 126 (4), 677-89. DOI: 10.1016/j.cell.2006.06.044. Engler, A. J.; Sweeney, H. L.; Discher, D. E.; Schwarzbauer, J. E., Extracellular matrix elasticity directs stem cell differentiation. J Musculoskelet Neuronal Interact 2007, 7 (4), 335. Dambrot, C.; Passier, R.; Atsma, D.; Mummery, C. L., Cardiomyocyte differentiation of pluripotent stem cells and their use as cardiac disease models. Biochem J 2011, 434 (1), 25-35. DOI: 10.1042/BJ20101707.

33 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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

40.

41.

42.

43.

44.

45.

46.

47.

48.

49.

50.

51.

Page 34 of 42

Kolesky, D. B.; Truby, R. L.; Gladman, A. S.; Busbee, T. A.; Homan, K. A.; Lewis, J. A., 3D bioprinting of vascularized, heterogeneous cell-laden tissue constructs. Adv Mater 2014, 26 (19), 3124-30. DOI: 10.1002/adma.201305506. Frisman, I.; Seliktar, D.; Bianco-Peled, H., Nanostructuring PEG-fibrinogen hydrogels to control cellular morphogenesis. Biomaterials 2011, 32 (31), 7839-46. DOI: 10.1016/j.biomaterials.2011.06.078. Shin, S. R.; Jung, S. M.; Zalabany, M.; Kim, K.; Zorlutuna, P.; Kim, S. B.; Nikkhah, M.; Khabiry, M.; Azize, M.; Kong, J.; Wan, K. T.; Palacios, T.; Dokmeci, M. R.; Bae, H.; Tang, X. S.; Khademhosseini, A., Carbon-nanotube-embedded hydrogel sheets for engineering cardiac constructs and bioactuators. ACS Nano 2013, 7 (3), 2369-80. DOI: 10.1021/nn305559j. Nunes, S. S.; Miklas, J. W.; Liu, J.; Aschar-Sobbi, R.; Xiao, Y.; Zhang, B.; Jiang, J.; Masse, S.; Gagliardi, M.; Hsieh, A.; Thavandiran, N.; Laflamme, M. A.; Nanthakumar, K.; Gross, G. J.; Backx, P. H.; Keller, G.; Radisic, M., Biowire: a platform for maturation of human pluripotent stem cell-derived cardiomyocytes. Nat Methods 2013, 10 (8), 7817. DOI: 10.1038/nmeth.2524. Radisic, M.; Yang, L.; Boublik, J.; Cohen, R. J.; Langer, R.; Freed, L. E.; VunjakNovakovic, G., Medium perfusion enables engineering of compact and contractile cardiac tissue. Am J Physiol Heart Circ Physiol 2004, 286 (2), H507-16. DOI: 10.1152/ajpheart.00171.2003. Yuan Ye, K.; Sullivan, K. E.; Black, L. D., Encapsulation of cardiomyocytes in a fibrin hydrogel for cardiac tissue engineering. J Vis Exp 2011, (55), e3251. DOI: 10.3791/3251. Stoehr, A.; Neuber, C.; Baldauf, C.; Vollert, I.; Friedrich, F. W.; Flenner, F.; Carrier, L.; Eder, A.; Schaaf, S.; Hirt, M. N.; Aksehirlioglu, B.; Tong, C. W.; Moretti, A.; Eschenhagen, T.; Hansen, A., Automated analysis of contractile force and Ca2+ transients in engineered heart tissue. Am J Physiol Heart Circ Physiol 2014, 306 (9), H1353-63. DOI: 10.1152/ajpheart.00705.2013. Loessner, D.; Stok, K. S.; Lutolf, M. P.; Hutmacher, D. W.; Clements, J. A.; Rizzi, S. C., Bioengineered 3D platform to explore cell-ECM interactions and drug resistance of epithelial ovarian cancer cells. Biomaterials 2010, 31 (32), 8494-506. DOI: 10.1016/j.biomaterials.2010.07.064. Mironi-Harpaz, I.; Wang, D. Y.; Venkatraman, S.; Seliktar, D., Photopolymerization of cell-encapsulating hydrogels: crosslinking efficiency versus cytotoxicity. Acta Biomater 2012, 8 (5), 1838-48. DOI: 10.1016/j.actbio.2011.12.034. Wang, Z.; Abdulla, R.; Parker, B.; Samanipour, R.; Ghosh, S.; Kim, K., A simple and high-resolution stereolithography-based 3D bioprinting system using visible light crosslinkable bioinks. Biofabrication 2015, 7 (4), 045009. DOI: 10.1088/17585090/7/4/045009. Bearzi, C.; Gargioli, C.; Baci, D.; Fortunato, O.; Shapira-Schweitzer, K.; Kossover, O.; Latronico, M. V.; Seliktar, D.; Condorelli, G.; Rizzi, R., PlGF-MMP9-engineered iPS cells supported on a PEG-fibrinogen hydrogel scaffold possess an enhanced capacity to repair damaged myocardium. Cell Death Dis 2014, 5, e1053. DOI: 10.1038/cddis.2014.12. Plotkin, M.; Vaibavi, S. R.; Rufaihah, A. J.; Nithya, V.; Wang, J.; Shachaf, Y.; Kofidis, T.; Seliktar, D., The effect of matrix stiffness of injectable hydrogels on the preservation 34 ACS Paragon Plus Environment

Page 35 of 42

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 Biomaterials Science & Engineering

52.

53.

54.

55. 56.

57.

58.

59.

of cardiac function after a heart attack. Biomaterials 2014, 35 (5), 1429-38. DOI: 10.1016/j.biomaterials.2013.10.058. Trattnig, S.; Ohel, K.; Mlynarik, V.; Juras, V.; Zbyn, S.; Korner, A., Morphological and compositional monitoring of a new cell-free cartilage repair hydrogel technology GelrinC by MR using semi-quantitative MOCART scoring and quantitative T2 index and new zonal T2 index calculation. Osteoarthritis Cartilage 2015, 23 (12), 2224-32. DOI: 10.1016/j.joca.2015.07.007. Du, M.; Chen, B.; Meng, Q.; Liu, S.; Zheng, X.; Zhang, C.; Wang, H.; Li, H.; Wang, N.; Dai, J., 3D bioprinting of BMSC-laden methacrylamide gelatin scaffolds with CBDBMP2-collagen microfibers. Biofabrication 2015, 7 (4), 044104. DOI: 10.1088/17585090/7/4/044104. Ma, Y.; Ji, Y.; Huang, G.; Ling, K.; Zhang, X.; Xu, F., Bioprinting 3D cell-laden hydrogel microarray for screening human periodontal ligament stem cell response to extracellular matrix. Biofabrication 2015, 7 (4), 044105. DOI: 10.1088/17585090/7/4/044105. Ozbolat, I. T., Bioprinting scale-up tissue and organ constructs for transplantation. Trends Biotechnol 2015, 33 (7), 395-400. DOI: 10.1016/j.tibtech.2015.04.005. Ouyang, L.; Yao, R.; Mao, S.; Chen, X.; Na, J.; Sun, W., Three-dimensional bioprinting of embryonic stem cells directs highly uniform embryoid body formation. Biofabrication 2015, 7 (4), 044101. DOI: 10.1088/1758-5090/7/4/044101. Faulkner-Jones, A.; Fyfe, C.; Cornelissen, D. J.; Gardner, J.; King, J.; Courtney, A.; Shu, W., Bioprinting of human pluripotent stem cells and their directed differentiation into hepatocyte-like cells for the generation of mini-livers in 3D. Biofabrication 2015, 7 (4), 044102. DOI: 10.1088/1758-5090/7/4/044102. Tulloch, N. L.; Muskheli, V.; Razumova, M. V.; Korte, F. S.; Regnier, M.; Hauch, K. D.; Pabon, L.; Reinecke, H.; Murry, C. E., Growth of engineered human myocardium with mechanical loading and vascular coculture. Circ Res 2011, 109 (1), 47-59. DOI: 10.1161/CIRCRESAHA.110.237206. Tsang, K. M.; Annabi, N.; Ercole, F.; Zhou, K.; Karst, D.; Li, F.; Haynes, J. M.; Evans, R. A.; Thissen, H.; Khademhosseini, A.; Forsythe, J. S., Facile One-step Micropatterning Using Photodegradable Methacrylated Gelatin Hydrogels for Improved Cardiomyocyte Organization and Alignment. Adv Funct Mater 2015, 25 (6), 977-986. DOI: 10.1002/adfm.201403124.

35 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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 36 of 42

For Table of Contents Use Only Direct Production of Human Cardiac Tissues by Pluripotent Stem Cell Encapsulation in Gelatin Methacryloyl Petra Kerscher1, Jennifer A. Kaczmarek1,#, Sara E. Head1,#, Morgan Brazel1,#, Wen J. Seeto1, Joonyul Kim2, Subhrajit Bhattacharya3,Vishnu Suppiramaniam3, Elizabeth A. Lipke1,*

36 ACS Paragon Plus Environment

Page 37 of 42

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 Biomaterials Science & Engineering

Figure 1 GelMA hydrogels were successfully used to produce GEhECTs. (a) HiPSCs were collected and combined with a liquid GelMA precursor solution; this mixture was then added into a circular PDMS mold and photocrosslinked for 40 s using visible light. (b) NMR spectrum showed synthesized GelMA with a methacryloyl substitution degree of 22% (signals of methacrylate double bonds at 5.3 and 5.5 ppm). (c) Photocrosslinked GelMA hydrogels initially swelled in the presence of PBS and degraded when incubated in collagenase, with over 50% mass loss by 6 h (n = 3 hydrogels). Figure 1 148x119mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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

Figure 2 GelMA encapsulated hiPSCs form soft, cell-laden hydrogels with hiPSCs surviving the encapsulation process to produce GEhECTs. (a) Acellular GelMA hydrogels exhibited an elastic modus of 660 ± 307 Pa; in comparison, cellular GelMA tissues 24 h after encapsulation (day -2) were softer, with an elastic modulus of 219 ± 140 Pa. Elastic modulus of the GEhECTs decreased to 90 ± 8.4 Pa by day 7. Later in the cardiac differentiation process (day 19), GEhECT elastic modulus was again higher at 248 ± 109 Pa. n = 3 tissues per group. Mean ± s.d. ANOVA P < 0.05, * vs. GelMA hydrogel. (b) Phase contrast image showed circular 3D cell-laden hydrogel 24 h after encapsulation (day 2) with (c) the majority of hiPSCs surviving the encapsulation process (green = viable, red = dead). (d) Three days after encapsulation (day 0), hiPSCs formed growing clusters, (e) displaying a round cell morphology (F-actin filaments visualized by phalloidin, green) with cells expanding beyond the original boundary of the hydrogel. Figure 2 130x191mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 38 of 42

Page 39 of 42

ACS Biomaterials Science & Engineering

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

ACS Biomaterials Science & Engineering

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

Figure 3 Encapsulated hiPSCs grew and differentiated to form a continuous tissue over time. (a) After the onset of cardiac differentiation (day 0), cells continued to grow within the original GelMA construct, forming a more connected tissue over time. (b) Circular surface area of tissues increased by 13.1 ± 1.7% by day 5 and decreased somewhat thereafter to form a dense tissue by day 7, immediately prior to the onset of contraction, observed consistently on day 8. n = 3 tissues per group. Mean ± s.d. ANOVA P < 0.05, # vs. Day -2, * vs. Day 0. (c) Tissue thickness was similar on day -2 and day 7, but increased by day 19 of differentiation. n = 3 tissues per group. Mean ± s.d. ANOVA P < 0.05, * vs. Day 7. (d) On day 17, GEhECTs contained approximately 70.4 ± 1.6% CMs (cTnT+), of which 10.6 ± 2.6% were proliferating (cTnT+, Ki67+), and 8.4 ± 0.3% fibroblasts (CD90+) (n = 3 tissues). (e) Gene expression of the cardiac markers MLC2v, αMHC, and βMHC on day 10, 20, and 70 GEhECTs showed temporal changes expected of maturing SC-CMs during culture. Cx43 gene expression in GEhECTs trended upward suggesting developing cell-cell coupling (n = 3 GEhECTs). Mean ± s.d. ANOVA * P < 0.05. Figure 3 169x165mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 40 of 42

Page 41 of 42

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 Biomaterials Science & Engineering

Figure 4 Frequency of spontaneous contraction and contractile velocity increased with GEhECT culture time. (a) Whole tissue staining of day 23 GEhECTs displayed striation (αSA, red) and gap junctions (Cx43, green). (b) GEhECTs formed uniformly contracting tissues by day 14, with frequencies of spontaneous contraction increasing significantly between days 14 and 40. (c, d) In addition to the increase in contraction frequency, average maximum contraction and relaxation velocities also increased from day 14 to day 40 of differentiation. This increase in contraction and relaxation velocity corresponded with visual observations of stronger contracting tissues (Supplementary Movies 3, 4). (e) Representative analyzed traces showing progression over time. n = 3-5 tissues. Mean ± s.d. ANOVA P < 0.05, * vs. Days 14, 17. Figure 4 105x61mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

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

Figure 5 GEhECT CMs remodel their microenvironment and show morphological differences expected from differentiating cells. (a) Day 22 tissues contained areas of high cell density primarily on the edges of GEhECTs where cells showed elongated F-actin filaments (phalloidin, green). (b, c) Day 0 (hiPSCs) encapsulated cells displayed more circular and smaller nuclei compared to day 22 (CMs) cells. n = 30 nuclei per condition. Mean ± s.d. ANOVA P < 0.05, * vs. HiPSC. (d) Following tissue dissociation, GEhECTdissociated CMs continued to spontaneously contract (Supplementary Movie 5) and locally self-aligned with each other, displaying defined sarcomeres (αSA, red) with (e) high intracellular alignment (n = 36). Figure 5 84x80mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 42 of 42

Page 43 of 42

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 Biomaterials Science & Engineering

Figure 6 Cardiac tissues responded to drug treatment and outside pacing. (a) GEhECTs produced with IMR90-1 and 19-9-11 hiPSC lines responded to the β-adrenergic agonist isoproterenol (Iso) and β-adrenergic antagonist propranolol (Prop); 1 µM Iso increased the frequency of spontaneous contraction, the subsequent addition of 1 µM Prop slowed down the contraction. (b) Day 30 GEhECT-dissociated CMs (from IMR90-1 hiPSCs) were cultured on MEA (Supplementary Movie 6) and (c) extrinsically paced. GEhECT-dissociated CMs responded to outside pacing frequencies up to 3.0 Hz. Figure 6 127x165mm (300 x 300 DPI)

ACS Paragon Plus Environment