Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Calcium Handling in Human Induced Pluripotent Stem Cell Derived Cardiomyocytes

  • Ilanit Itzhaki ,

    Contributed equally to this work with: Ilanit Itzhaki, Sophia Rapoport

    Affiliation Sohnis Family Research Laboratory for Cardiac Electrophysiology and Regenerative Medicine, Bruce Rappaport Faculty of Medicine, Technion - Israel Institute of Technology, Haifa, Israel

  • Sophia Rapoport ,

    Contributed equally to this work with: Ilanit Itzhaki, Sophia Rapoport

    Affiliation Department of Biophysics Physiology, Technion - Israel Institute of Technology, Haifa, Israel

  • Irit Huber,

    Affiliation Sohnis Family Research Laboratory for Cardiac Electrophysiology and Regenerative Medicine, Bruce Rappaport Faculty of Medicine, Technion - Israel Institute of Technology, Haifa, Israel

  • Itzhak Mizrahi,

    Affiliation Sohnis Family Research Laboratory for Cardiac Electrophysiology and Regenerative Medicine, Bruce Rappaport Faculty of Medicine, Technion - Israel Institute of Technology, Haifa, Israel

  • Limor Zwi-Dantsis,

    Affiliation Sohnis Family Research Laboratory for Cardiac Electrophysiology and Regenerative Medicine, Bruce Rappaport Faculty of Medicine, Technion - Israel Institute of Technology, Haifa, Israel

  • Gil Arbel,

    Affiliation Sohnis Family Research Laboratory for Cardiac Electrophysiology and Regenerative Medicine, Bruce Rappaport Faculty of Medicine, Technion - Israel Institute of Technology, Haifa, Israel

  • Jackie Schiller,

    Affiliation Department of Biophysics Physiology, Technion - Israel Institute of Technology, Haifa, Israel

  • Lior Gepstein

    mdlior@tx.technion.ac.il

    Affiliation Sohnis Family Research Laboratory for Cardiac Electrophysiology and Regenerative Medicine, Bruce Rappaport Faculty of Medicine, Technion - Israel Institute of Technology, Haifa, Israel

Abstract

Background

The ability to establish human induced pluripotent stem cells (hiPSCs) by reprogramming of adult fibroblasts and to coax their differentiation into cardiomyocytes opens unique opportunities for cardiovascular regenerative and personalized medicine. In the current study, we investigated the Ca2+-handling properties of hiPSCs derived-cardiomyocytes (hiPSC-CMs).

Methodology/Principal Findings

RT-PCR and immunocytochemistry experiments identified the expression of key Ca2+-handling proteins. Detailed laser confocal Ca2+ imaging demonstrated spontaneous whole-cell [Ca2+]i transients. These transients required Ca2+ influx via L-type Ca2+ channels, as demonstrated by their elimination in the absence of extracellular Ca2+ or by administration of the L-type Ca2+ channel blocker nifedipine. The presence of a functional ryanodine receptor (RyR)-mediated sarcoplasmic reticulum (SR) Ca2+ store, contributing to [Ca2+]i transients, was established by application of caffeine (triggering a rapid increase in cytosolic Ca2+) and ryanodine (decreasing [Ca2+]i). Similarly, the importance of Ca2+ reuptake into the SR via the SR Ca2+ ATPase (SERCA) pump was demonstrated by the inhibiting effect of its blocker (thapsigargin), which led to [Ca2+]i transients elimination. Finally, the presence of an IP3-releasable Ca2+ pool in hiPSC-CMs and its contribution to whole-cell [Ca2+]i transients was demonstrated by the inhibitory effects induced by the IP3-receptor blocker 2-Aminoethoxydiphenyl borate (2-APB) and the phosopholipase C inhibitor U73122.

Conclusions/Significance

Our study establishes the presence of a functional, SERCA-sequestering, RyR-mediated SR Ca2+ store in hiPSC-CMs. Furthermore, it demonstrates the dependency of whole-cell [Ca2+]i transients in hiPSC-CMs on both sarcolemmal Ca2+ entry via L-type Ca2+ channels and intracellular store Ca2+ release.

Introduction

The breakthrough technology introduced by Takahashi and Yamanka in 2006 enables the derivation of pluripotent stem cells by reprogramming of somatic cells with a set of transcription factors [1]. Application of this reprogramming strategy to human fibroblasts resulted in the establishment of human induced pluripotent stem cells (hiPSCs) [2], [3]. The hiPSC lines generated were demonstrated to closely-resemble the previously described human embryonic stem cell (hESC) lines [4], including in their ability to differentiate into advanced cell-derivatives of all three germ layers.

Only a limited number of studies described the ability to direct hiPSC differentiation towards the desired cardiac-lineage [5], [6], [7], [8]. As a consequence very little is known about these human iPSC-derived cardiomyocytes' (hiPSC-CMs) functional capabilities, and even less is known about their excitation-contraction (E-C) coupling and Ca2+-handling properties [5]. Thorough characterization of the functional nature of hiPSC-CMs must be conducted before these cells can be considered as candidates for the emerging fields of regenerative medicine (potentially providing autologous cardiomyocytes for myocardial regeneration strategies) and personalized medicine (for the derivation of patient/disease-specific in-vitro models of human cardiac tissue). The suitability of hiPSC-CMs for such tasks depends, in part, on their contractile characteristics which in turn greatly depend on the Ca2+-handling nature of these cells.

In adult ventricular cardiomyocytes, Ca2+-handling displays a well-defined sequence of events. Ca2+ influx into the cells via depolarization-activated L-type Ca2+ channels serves as an initial trigger that is then amplified several folds by sarcoplasmic reticulum (SR) Ca2+-store release via Ca2+-sensitive ryanodine receptors (RyRs); a process known as Ca2+ induced Ca2+ release (CICR) [9], [10]. Nevertheless, exceptions to the CICR model have been reported in different species and in developing cardiomyocytes with whole-cell [Ca2+]i transients being derived solely from Ca2+ influx through the membrane Ca2+ channels [11], [12], [13] or by spontaneous release from the intracellular Ca2+ stores [14].

In the current study, we tested the hypothesis that whole-cell [Ca2+]i transients in hiPSC-CMs are dependent on both transsarcolemmal Ca2+ entry via L-type Ca2+ channels and on intracellular store Ca2+ release. To test this hypothesis, we initially carried out gene expression and immunostaining studies to show that key Ca2+-handling proteins are expressed in hiPSC-CMs. To test for their functionality we then performed detailed laser-confocal Ca2+ imaging coupled with targeted pharmacological interventions. Initial studies confirmed the importance of transsarcolemmal Ca2+ entry through the L-type Ca2+ channels for modulation of the whole-cell [Ca2+]i transients in these cells. We then demonstrated that hiPSC-CMs display functional and loaded RyR-regulated intracellular Ca2+ stores that contribute as well to the whole-cell [Ca2+]i transient. In addition, we investigated the functionality of SR Ca2+ ATPase (SERCA) pumps, which serve as an important SR Ca2+ sequestration pathway. We found the SERCA pumps to be functional and responsible for the refilling of hiPSC-CMs' SR Ca2+ store content. Finally, we also present evidence showing the expression and functionality of inositol-1,4,5-trisphosphate receptors (IP3Rs) in hiPSC-CMs and demonstrate the important contribution of this alternative pathway to Ca2+-handling in these cells.

Methods

Differentiation of hiPSCs into cardiomyocytes

The hiPSC line utilized in the current study (hIH-1) was recently established in our laboratory [15] by retroviral delivery of three reprogramming factors: OCT4, SOX2, and KLF4 together with valproic acid (VPA), a histone deacetylase inhibitor potentiating the reprogramming ability of these factors [16]. This hiPSCs line was demonstrated to fulfill all the criteria defining the iPSC state including full reprogramming, pluripotency, and genetic stability [15]. In the current study we used two clones (hIH-1 clones 1&2) of this line that were derived independently during reprogramming of the human fibroblasts. In addition, we also studied a second well-characterized hiPSCs line (hFib2-iPS cells; kindly provided by G.Q Daley) [8], [17], which was established by retroviral transduction of human fibroblasts with OCT4, SOX2, c-MYC, KLF4, together with hTERT and SV40-large T [17].

Undifferentiated hiPSC colonies were cultured on a mitotically-inactivated MEF feeder layer [8]. The culture medium consisted of 80%-knockout high-glucose glutamine-free DMEM with sodium-pyruvate supplemented with 20%-serum-replacement, 1 mM L-glutamine, 0.1 mM mercaptoethanol, 4 ng/mL human recombinant basic fibroblast-growth-factor, and 1%-nonessential amino acid stock (Invitrogen).

To induce differentiation, hiPSCs were dispersed into small clumps using collagenase-IV (Life-Technologies, 1 mg/mL, 20 min) and cultivated in suspension where they aggregated to form embryoid bodies (EBs). The EBs were plated after 10 days on gelatin-coated culture dishes and observed for the appearance of spontaneous contracting areas. The beating areas within the EBs were mechanically microdissected at 30–50 days following the appearance of spontaneous beating to allow comparison with studies assessing hESC-derived cardiomyocytes at similar developmental stages [18], [19]. This was followed by enzymatic dispersion (1 mg/ml collagenase B, Roche, Mannheim, Germany) at 37°C for 30 min to derive single cardiomyocytes or small monolayered clusters. After dispersal, the cells were plated onto fibronectin-coated glass coverslips for experimentation.

Reverse transcription PCR (RT-PCR) analysis

The expression of various cardiac Ca2+-handling genes was tested using semi-quantitative RT-PCR. Total RNA was extracted from human embryonic kidney cells (HEK293, American Type Culture Collection (ATCC), Manassas, VA) and from beating hiPSC-CMs using RNeasy Mini Kit (Qiagen). cDNA was synthesized using Verso™-RT-PCR-Kit (Thermo-Scientific). 1.6 ng of cDNA was subjected to the following PCR program: 3 min at 93°C, 30 sec at 93°C, 30 sec at 60°C, and 30 sec at 72°C. The PCR-related primers are detailed in Table 1.

Immunocytochemistry

Single cells and small monolayered clusters were fixed using 4% paraformaldehyde and permeabilized with 1% Triton-X-100 (Sigma). Cells were blocked with 5% horse serum and 1% bovine serum albumin and incubated overnight at 4°C with primary antibodies targeting sarcomeric α-actinin (1∶200, Sigma-Aldrich), ryanodine receptor (1∶1,000, Chemicon), and pan-IP3R (reacting with the C-terminal cytoplasmic domain of IP3R types 1, 2, and 3) (1∶10, Millipore). The secondary antibodies were Cy3-conjugated donkey anti-mouse and Cy2-conjugated goat anti-rabbit IgG antibodies at 1∶200 for 1 h (Jackson). Nuclei were counterstained with DAPI (Sigma). Preparations were examined using a laser-scanning confocal microscope (Zeiss LSM-510-PASCAL).

Ca2+ Imaging

Cells were loaded with 5 µM fluo-4 fluorescent Ca2+ indicator (Molecular Probes) in the presence of Pluronic F-127 (Molecular Probes) at a dilution of 2∶1 to allow the recording of intracellular Ca2+-transients (whole-cell [Ca2+]i transients) as previously described [18]. Experiments were conducted in the presence of indicator-free tyrode solution containing in mM: NaCl-140, KCl-5.4, CaCl2-1.8, MgCl2-1, HEPES-10, glucose-10 at pH 7.4 via 1 mM NaOH. The tyrode solution was superfused at a rate of 1 ml/minute and at a temperature of 37°C. Intracellular Ca2+-transients were recorded using a confocal imaging system (Fluoview; Olympus) mounted on an upright BX51WI Olympus microscope equipped with a X60 water objective [20]. Data were analyzed utilizing MatLab-based custom-written software.

To investigate caffeine mobilization of store Ca2+, temporally-limited puffs of caffeine were applied. This specific technique [18] was chosen to overcome the technical issue of the rate of caffeine delivery to the cells. The caffeine puff (20 mM, <2s) was applied by pressure-ejection through a pipette situated approximately 100 µm away from the target cells' position. The focal puff pipette was positioned central to the plane of the line scan and in the direction of constant bulk flow of the tyrode solution.

Statistical analysis

Data are presented as mean ± standard error of mean (SEM). Student's paired t-test was used to compare between means. When the effects of multiple concentrations of the pharmacological agents were studied, then one way ANOVA was used followed by Dunnet's post-hoc comparison to baseline values. p<0.05 was considered to be statistically significant.

Results

Expression of Ca2+-handling molecules in hiPSC-CMs

To evaluate whether cardiac-related Ca2+-handling molecular components are present in the hiPSC-CMs, we first used semi-quantitative RT-PCR analysis. We assessed for the expression of the following Ca2+-handling proteins: RyR2, IP3R2, SERCA2a, Cav1.2 (encoding the α-1C subunit of the L-type Ca2+ channel), Calsequestrin, and Phospholamban. As can be seen in Figure 1, all of these genes were expressed in the hiPSC-CMs while absent in control HEK293 cells.

thumbnail
Figure 1. Semi-quantitative RT-PCR analysis for cardiac-specific Ca2+-handling proteins in hiPSC-CMs.

Gene expression pattern of proteins involved in Ca2+-handling in hiPSC-CMs (upper panel) and control HEK293 cells. The lower panel displays genes expression in control HEK293 cells. β-tubulin was used as a normalizing house-keeping gene. Abbreviations: RyR2, ryanodine receptor; IP3R2, inositol 1,4,5-triphosphate receptor; SERCA2a, SR Ca ATPase; CaV1.2, L-type Ca2+ channels; Cal, calsequestrin, and PLN, phospholamban.

https://doi.org/10.1371/journal.pone.0018037.g001

Spontaneous whole-cell [Ca2+]i transients in hiPSC-CMs

Spontaneous whole-cell [Ca2+]i transients in hiPSC-CMs were recorded from spontaneously beating dispersed single cells or small monolayered clusters, under native cytosolic conditions. These transients were monitored in Fluo-4 (a fluorescent Ca2+ indicator) loaded cells, examined under a laser scanning confocal microscope utilizing the line-scan mode. Line scans were adjusted to avoid the cell nuclei and were localized at mid-cell z-section depth. Under control conditions, in the presence of 1.8 mM external bath Ca2+, spontaneous whole-cell [Ca2+]i transients were recorded as cell-wide rhythmic events in all cells tested (n = 96) (Figure 2A).

thumbnail
Figure 2. Whole-cell [Ca2+]i transients in hiPSC-CMs and their requirement for Ca2+ influx via L-type Ca2+ channels.

(A) Confocal line-scan images showing changes in intracellular Ca2+ in a fluo-4 loaded hiPSC-CM (derived from the hIH1 line). The images showing spontaneous whole-cell [Ca2+]i transients are displayed either as a distance-time plot (upper panel) or as a line-scan tracing (lower panel). (B) Whole-cell [Ca2+]i transients in the presence (left) and absence (right) of bath Ca2+. (C) Whole-cell [Ca2+]i transients before (left) and after (right) application of nifedipine (1 µM). (D) Dose-response curve showing the percentage of decrease in [Ca2+]i transients amplitude relative to baseline conditions as function of nifedipine concentration (0.05, 0.1 and 1 µM, n = 6). *p<0.05 and **p<0.01 when mean absolute values were compared to baseline values. Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.g002

Ca2+ influx via L-type Ca2+ channels contributes to whole-cell [Ca2+]i transients

Transmembranal Ca2+ influx is an important initial trigger for excitation-contraction coupling in adult cardiomyocytes [9] and in hESC-derived cardiomyocytes (hESC-CMs) [19]. Therefore, the next step was to investigate whether the development of hiPSC-CMs whole-cell [Ca2+]i transients require external Ca2+. To this end, we recorded whole-cell [Ca2+]i transients in the presence (1.8 mM bath Ca2+) and absence of Ca2+ in the bath solution (n = 6). As can be appreciated in Figure 2B, in the absence of bath Ca2+ the whole-cell [Ca2+]i transients were completely abolished.

To test whether the L-type Ca2+ channel serves as an important transmembranal Ca2+ influx pathway in hiPSC-CMs, as documented in adult cardiomyocytes [9], we tested the effect of Nifedipine, a L-type Ca2+ channel blocker. Whole-cell [Ca2+]i transients were recorded before and after the application of 1 µM Nifedipine (Figure 2C). Similarly to what was observed in the absence of bath Ca2+ (Figure 2B), 1 µM Nifedipine led to the complete elimination of whole-cell [Ca2+]i transients (n = 9). Dose-response studies using lower concentrations of nifedipine (0.05–1 µM, n = 6) demonstrated that the cells were extremely sensitive to L-type channel blockade with a steep decrease in [Ca2+]i transients amplitude observed at a very low concentration (Figure 2D).

To verify that the results obtained were not due to clonal or line variations, we compared the results obtained in cardiomyocytes differentiated from two different clones of the main hiPSC line studied (hIH1 - clones 1&2 derived independently from the human fibroblasts) as well as from an additional well-characterized hiPSC line (hfib2-5) derived using the traditional 4-factors method [8], [17]. The dependency of whole-cell [Ca2+]i transients on the presence of functional L-type Ca2+ channels was found to be independent of the specific hiPSC clone or line used. Thus, Nifedipine (1 µM) application resulted in complete elimination of whole-cell [Ca2+]i transients in all cases (Figure S1).

Taken together these data confirm that transmembranal Ca2+ influx and specifically Ca2+ entry via L-type Ca2+ channels are important requirements for the generation of whole-cell [Ca2+]i transients in hiPSC-CMs.

Functional RyR-mediated intracellular Ca2+ stores exist and contribute to whole-cell [Ca2+]i transients

We next conducted immunocytostaining studies of hiPSC-CMs probing for both RyR2 and sarcomeric α-actinin in small monolayered clusters (Figure 3A). As previously shown in hESC-CMs [18], [21] sarcomeric α-actinin staining in hiPSC-CMs displayed a relatively disorganized striated sarcomeric arrangement (Figure 3A, left). RyR2 expression was exhibited throughout the cytosol (Figures 3A, middle), with some myofilaments co-localization (Figure 3A, right). The perinuclear region displayed intense staining as was similarly observed in mouse ESC-CMs [22] and hESC-CMs [18].

thumbnail
Figure 3. Localization and functionality of Ca2+ store ryanodine receptors.

(A) A hIH1 hiPSC-CM co-labeled with antibodies for sarcomeric α-actinin (left) and RyR2 (middle). The merged image is displayed in the right panel. (B) A line-scan presenting the effect of 20 mM caffeine puff application (noted by the arrows). (C) A dose-response curve displaying the magnitude of the caffeine-induced Ca2+ release (quantified by the percentage increase of the amplitude of the caffeine-induced transient relative to the preceding action-potential induced [Ca2+]i transient) as function of the caffeine bolus dosage (10, 20 and 40 mM). *p<0.05 when compared to the lowest (10 mM) dosage. (D) A line-scan tracings presenting caffeine-induced Ca2+ transients in the presence (left) and absence (right) of bath Ca2+. Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.g003

To determine whether hiPSC-CMs possess loaded SR Ca2+ stores that release Ca2+ via functional RyRs we tested for caffeine responsiveness. Caffeine mobilization of store Ca2+ and its effect on whole-cell [Ca2+]i transients were measured by applying a pressure-ejected caffeine puff (20 mM, 2 sec) on to fluo-4 loaded hiPSC-CMs. As can be observed in Figure 3B, caffeine application elicited an instantaneous, rapid, and large release of Ca2+ from the intracellular stores, resulting in a high amplitude caffeine-induced Ca2+ transient (n = 26). This was followed by reversible quiescence of whole-cell [Ca2+]i transients (Fig. 3B) postulated to be a consequence of intracellular Ca2+ stores depletion. This phenomenon was noted in cardiomyocytes derived from all hiPSCs clones and lines studied (Figure S2A). Finally, dose-response studies (10, 20, and 40 mM caffeine puffs, n = 6) showed an escalating effect with an increase in the relative magnitude of caffeine-induced Ca2+ release (Figure 3C).

Next, it was important to validate that the caffeine-induced [Ca2+]i transient was indeed a consequence of RyR-mediated SR Ca2+ release. To exclude the plausible contribution of Ca2+ influx through voltage-gated Ca2+ channels 20 mM caffeine puffs were applied in the presence (1.8 mM Ca2+; Figure 3D, left) and absence (Figure 3D, right) of bath Ca2+. Similarly to what was observed under control conditions (Figure 3D, left), caffeine puffs applied in the absence of bath Ca2+ (n = 6) induced an instantaneous rapid caffeine-induced [Ca2+]i transient (Figure 3D, right) displaying an amplitude similar to that observed under control conditions. To exclude the possibility that the caffeine-induced [Ca2+]i transient was a result of a mechanical stimulation to the cell surface, caused by the actual pressure injected puff, control puff (tyrode solution) trials were conducted. These control puffs did not trigger any apparent intracellular Ca2+ response (data not shown).

Finally, we also tested the effect of ryanodine, a RyR antagonist. For this purpose we monitored whole-cell [Ca2+]i transients before (Figure 4A, left) and after (Figure 4A, right) application of ryanodine (10 µM). Ryanodine administration led to a significant reduction in Ca2+ release, as observed by the decrease in whole-cell [Ca2+]i transients amplitude (Figure 4B; n = 14; p<0.05) and also to significant slowing of whole-cell [Ca2+]i transients frequency (Figure 4C; n = 6; p<0.05). The effect of ryanodine was noted in cardiomyocytes derived from all hiPSC clones and lines studied (Figure S2B) and was dose-dependent, as increasing doses of ryanodine (5, 10, 15, and 20 µM, n = 8) led to a gradual decrease in whole-cell [Ca2+]i transients amplitude in both lines studied (Figures 4D and S2C). Taken together, these data demonstrate that hiPSC-CMs display caffeine-responsive and ryanodine-sensitive SR Ca2+ stores capable of unloading Ca2+ via RyR-mediated Ca2+ release and contributing to whole-cell [Ca2+]i transients.

thumbnail
Figure 4. Ryanodine-sensitive Ca2+ stores.

(A) Line-scan tracings of whole-cell [Ca2+]i transients recorded from a representative hIH1 hiPSC-CM under control conditions (left) and in the presence of 10 µM Ryanodine (right). (B–C) Summary of the ryanodine-mediated (10 µM) diminution in whole-cell [Ca2+]i transients amplitude (B) and frequency (C). *p<0.05 when compared to baseline (control) values. (D) Dose-response curve for the effects of ryanodine (5, 10, 15 and 20 µM) on whole-cell [Ca2+]i transients amplitude displayed as the percentage of decrease from baseline values. **p<0.01 when mean absolute values were compared to baseline values. Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.g004

SERCA-mediated SR Ca2+ uptake is required for whole-cell [Ca2+]i transients

We next tested for the functionality and contribution of another important Ca2+-handling protein situated on the SR membrane, SERCA. To test for SERCA functionality in hiPSC-CMs we recorded whole-cell [Ca2+]i transients before (Figure 5A, left) and after (Figure 5A, right panel) application of 10 µM thapsigargin, a specific SERCA inhibitor. Application of 10 µM thapsigargin (15 min incubation) resulted in a decrease in whole-cell [Ca2+]i transients' amplitude in both lines studied (n = 7, Figure 5A, and Figure S3A). The effect of thapsigargin was dose-related with a greater decrease in whole-cell [Ca2+]i transients amplitude (observed at 15 min following drug application) with escalating doses (1, 5, 10 and 20 µM, Figure 5B).

thumbnail
Figure 5. The effects of SERCA inhibition.

(A) Line-scan tracings of whole-cell [Ca2+]i transients in a hIH1 hiPSC-CM under control conditions (left) and after 15 minutes (right) of continuous 10 µM thapsigargin exposure. (B) Dose-response effects of thapsigargin (1, 5, 10 and 20 µM) on [Ca2+]i transients amplitude (displayed as percentage decrease versus baseline values). *p<0.05 and **p<0.01 when mean absolute values were compared to baseline values. (C) Caffeine-induced Ca2+ transients (denoted by arrows) under control conditions (upper), after 30 minutes (middle), and 32 minutes (lower) of thapsigargin exposure. Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.g005

We have shown above (Figures 3, 4, 5A) that in the case of hiPSC-CMs SR Ca2+ release is an important contributor of whole-cell [Ca2+]i transients. Therefore, we wanted to confirm whether the inhibiting effect of thapsigargin on whole-cell [Ca2+]i transients was due to a decrease in SR Ca2+ content, as a consequence of SERCA Ca2+ uptake inhibition. To this end we conducted repeated SR Ca2+ load measurements by applying 20 mM caffeine puffs under control conditions (Figure 5C, upper panel) and in the presence of 10 µM thapsigargin (Figure 5C, middle and lower panels). In contrast, to the enhanced caffeine-induced [Ca2+]i transient observed under control condition (Figure 5C, upper panel), application of a caffeine puff, when complete succession of all whole-cell [Ca2+]i transients has taken place as a result of the thapsigargin uptake inhibition, produced only a minor effect. This effect was displayed as a miniscule caffeine-induced [Ca2+]i transient (Figure 5C, middle panel), that was completely omitted in the subsequent caffeine puff (n = 7; Figure 5C, lower panel). A similar phenomenon was also observed in cardiomyocytes derived from a second hiPSC line (hfib2-5, Figure S3B). The absent caffeine-induced signal at this stage is postulated to be a consequence of the inability of the SR to reload due to SERCA uptake inhibition by thapsigargin.

IP3-mediated calcium release contributes to whole-cell [Ca2+]i transients

IP3-dependent signaling has been shown to play an important role during the process of cardiac development [23], [24]. It was recently demonstrated that in both mouse [25], [26] and human [18], [27] ESC-CMs an IP3-releasable Ca2+ pool is expressed and functional. To evaluate the potential role of an IP3-releasable Ca2+ pool in hiPSC-CMs, we first utilized immunocytostaining to detect the presence of the IP3R. These stainings displayed positive IP3R immunosignal, strongly distributed around the nucleus (Figure 6A), in a similar fashion to that observed in neonatal rat cardiomyocytes [28], mouse ESC-CMs [26], and hESC-CMs [18].

thumbnail
Figure 6. IP3Rs localization and function.

(A) A hIH1 hiPSC-CM co-stained for pan-IP3R (middle) and sarcomeric α-actinin (left). Merged images are presented in the right panel. (B) Whole-cell [Ca2+]i transients before (left) and after (right) 2-APB (2 µM) application. (C) Whole-cell [Ca2+]i transients before (left) and after (right) U73122 (2 µM) application. (D) Summary of the effects of 2-APB and U73122 on whole-cell [Ca2+]i transients amplitude and frequency. *p<0.05 when compared to baseline values. (E–F) Dose-response curves displaying the effects of 2-APB (0.1, 0.5, 1, 2, 5 and 10 µM) (E) and U73122 (0.5, 1, 2 and 3 µM) (F) on whole-cell [Ca2+]i transients amplitude (displayed as the relative decrease (%) versus baseline values). *p<0.05 and **p<0.01 when mean absolute values were compared to baseline values. Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); IP3R, inositol-1,4,5-trisphosphate receptor; sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.g006

To assess the potential contribution of the IP3R to the modulation of whole-cell [Ca2+]i transients in hiPSC-CMs, we tested the effect of IP3R blockade utilizing two different antagonistic approaches. First, whole-cell [Ca2+]i transients were recorded before (Figure 6B, left) and after (Figure 6B, right) application of a low concentration of 2-aminoethoxyphenyl borate (2-APB, 2 µM, n = 14), a well-known cell-permeate IP3R antagonist. 2-APB application resulted in a significant decrease in whole-cell [Ca2+]i transients amplitude (Figures 6B,D) and significantly slowed down whole-cell [Ca2+]i transients frequency (Figure 6B,D). This effect was also dose-related (0.1, 0.5, 1, 2, 5, and 10 µM; n = 8) as displayed in Figure 6E. Strengthening the above-mentioned results, application of 2 µM of U73122 (a phosopholipase C inhibitor, n = 15) led to a significant diminution of whole-cell [Ca2+]i transients amplitude (Figure 6C–D) and a slowing of whole-cell [Ca2+]i transients frequency (Figure 6C–D). This effect was dose-dependent (0.5, 1, 2, and 3 µM; n = 8; Figure 6E). The effects of 2-APB and U73122 were found to be independent of the hiPSC line used (Figure 6 and Figure S4). These observations imply that in hiPSC-CMs an IP3-releasable Ca2+ pool is functional and contributes to the modulation of Ca2+-handling in these cells.

Discussion

The ability to derive hiPSCs by reprogramming of adult human fibroblasts [2], [3] coupled with the capability to coax the differentiation of the generated hiPSCs into the cardiac lineage [6], [7], [8] opens unique opportunities for basic and translational cardiovascular research. Yet, to fulfill the promise of this unique technology in areas such as cardiovascular regenerative and personalized medicine, it is important that the generated hiPSC-CMs express a cardiomyocyte-specific phenotype; possessing, among other requirements, functional excitation-contraction coupling components.

In the current study, we focused on studying the Ca2+-handling properties of hiPSC-CMs. Our major findings consist of the following: (1) Mature molecular components RyR2, IP3R2, SERCA2a, CaV1.2, calsequestrin, and phospholamban are expressed in these cells; (2) hiPSC-CMs display spontaneous whole-cell [Ca2+]i transients; (3) Ca2+ entry via L-type Ca2+ channels is required for triggering whole-cell [Ca2+]i transients; (4) Caffeine-responsive and ryanodine-sensitive Ca2+ stores are loaded and functional; (5) RyR-mediated SR Ca2+ release contributes to whole-cell [Ca2+]i transients; (6) SERCA pumps are functional and enable the refilling of SR Ca2+ content, required for the modulation of whole-cell [Ca2+]i transients; (7) An IP3-releasable Ca2+ pool is expressed, functional, and contributes to whole-cell [Ca2+]i transients; and (8) the results obtained are comparable in cardiomyocytes derived from different differentiation experiments of the same hiPSC line, from different hiPSC clones (derived independently from fibroblasts of the same individual), and from different hiPSCs lines established using different techniques (using either three or four reprogramming factors). Taken together we conclude that whole-cell [Ca2+]i transients in hiPSC-CMs depend on both Ca2+ influx via L-type Ca2+ channels and intracellular Ca2+ store release, as previously documented in mouse [25], [26] and human ESC-CMs [18], [19].

Whole-cell [Ca2+]i transients in hiPSC-CMs depend on Ca2+ entry via L-type Ca2+ channels and Ca2+ release from RyR-mediated SR Ca2+ stores

In adult cardiomyocytes the development of an action potential triggers the opening of L-type Ca2+ channels. The resulting Ca2+ influx serves as a trigger for activating RyR-mediated SR Ca2+ release by a mechanism known as CICR [9], [10]. In addition to the well-established CICR mechanism in adult mammalian cardiomyocytes other mechanistic models were also proposed to be responsible for E-C coupling in various species and at earlier cardiomyocyte developmental stages. These include reports in frog [13] and turtle [11] adult ventricular cells as well as in primary embryonic murine myocytes [12], in which whole-cell [Ca2+]i transients were shown to be derived solely from Ca2+ influx through membrane Ca2+ channels. In contrast, global whole-cell [Ca2+]i transients in early cardiac cells derived from mouse ESC [14] were reported to be the result of spontaneous Ca2+ release from intracellular Ca2+ stores without the triggering of membrane Ca2+ currents.

The mechanism underlying E-C coupling in hESC-CMs is somewhat controversial. While some reports suggested the absence of a functional SR Ca2+ store in hESC-CMs and postulated that essentially all of the [Ca2+]i transients in these cells were a consequence of transsarcolemmal Ca2+ influx via membranal Ca2+ channels [29], others have argued for a more mature-like CICR mechanism [18], [19], [30]. The latter studies reported the presence of a functional caffeine-responsive and ryanodine-sensitive SR Ca2+ store in at least a subset [30] if not all [18], [19], of the cells tested; in a similar manner to the hiPSC-CMs studied in the current study.

Our results support the contribution of both the transsarcolemmal Ca2+ influx and intracellular Ca2+ store release to whole-cell [Ca2+]i transients in hiPSC-CMs. The importance of the L-type Ca2+ current in generating whole-cell [Ca2+]i transients in these cells was manifested by the elimination of these transients in the absence of external Ca2+ or in the presence of nifedipine, a selective L-type Ca2+ channel antagonist. A similar requirement for external Ca2+ and the consequent transsarcolemmal Ca2+ influx was documented in adult cardiomyocytes [9], [31], [32], hESC-CMs [19] and mouse ESC-CMs [33], [34].

The contribution of Ca2+ release from intracellular SR Ca2+ stores to whole-cell [Ca2+]i transients was demonstrated by pharmacological studies interfering either with SR Ca2+ release (caffeine and ryanodine) or reuptake (thapsigargin). Caffeine increases RyR2s opening, leading to a single large-amplitude caffeine-induced Ca2+ transient, considered to be a descriptive index of the level of SR Ca2+ load [35]. In hiPSC-CMs a local pressure ejected puff of caffeine elicited a local bolus release of Ca2+, followed by a short and reversible succession of whole-cell [Ca2+]i transients. These results suggest that caffeine induced depletion of the SR Ca2+ store and point to a whole-cell [Ca2+]i transient dependency on SR Ca2+ content.

The main Ca2+ source for the caffeine induced [Ca2+]i increase is RyR-mediated SR Ca2+ release and is not dominated by Ca2+ influx via voltage-gated Ca2+ channels. This was demonstrated by the similar caffeine-induced rise in intracellular Ca2+ documented in the absence of extracellular bath Ca2+ further confirming the presence of a caffeine-responsive intracellular Ca2+ store. Similar results were also acquired in hESC-CMs [18], [19].

We also applied the RyR antagonist, ryanodine, to further study hiPSC-CMs RyR-mediated SR Ca2+ release. Ryanodine, has been reported to reduce by approximately twofold the conductance of RyRs in the SR [36]. In hiPSC-CMs, ryanodine application led to a dose-dependent diminution in Ca2+ release observed as a significant decrease in the amplitude of whole-cell [Ca2+]i transients. A similar ryanodine induced effect was also reported in pacemaker cells isolated from rabbit sinoatrial node [37], in ESC-CMs [18], [30] and in mouse ESC-CMs [33]. In addition, in hiPSC-CMs ryanodine application also resulted in slowing of the spontaneous whole-cell [Ca2+]i transients firing rate. This phenomenon was also previously documented in rabbit sinoatrial node pacemaker cells where a similar slowing in firing rate was detected in the presence of ryanodine [37].

Functional SERCA pumps enable the loading of SR Ca2+ store content required for whole-cell [Ca2+]i transients

For cellular relaxation to take place Ca2+ must be removed from the cytosol. In adult cardiomyocytes, one of the main Ca2+ removal pathways is the SR Ca2+ ATPase pump (SERCA) [9], [38]. These pumps decrease intracellular Ca2+, by sequestering Ca2+ back into the SR, and in this manner also regulate SR Ca2+ load [38]. In adult human cardiomyocytes, SERCA pumping activity is responsible for 70% of Ca2+ sequestration from the cytosol back into the SR [9]. To investigate the functionality and contribution of the SERCA pumps to whole-cell [Ca2+]i transients through their ability to reload the SR Ca2+ stores in hiPSC-CMs we applied the SERCA inhibitor thapsigargin. Thapsigargin acted slowly to progressively decrease the amplitude of whole-cell [Ca2+]i transients, eventually leading to their complete inhibition. A similar effect was observed in spontaneously beating fluo-4 loaded isolated mouse ESC-CMs34. An antagonistic effect of thapsigargin on [Ca2+]i transients was also reported in human ESC-CMs [19], [30].

The key role of SERCA in reloading the SR, and thereby indirectly modulating hiPSC-CMs whole-cell [Ca2+]i transients, was further demonstrated by the miniscule effect of caffeine in hiPSC-CMs pretreated with thapsigargin, as a result of a pronounced diminution in SR Ca2+ content. Interestingly, under conditions of SERCA uptake inhibition (thapsigargin) a low SR Ca2+ content was retained yet [Ca2+]i transients were completely abolished. This can be explained by reports showing that decrease in SR Ca2+ content can disproportionately inhibit SR Ca2+ release [39], which as shown here is an important contributor to hiPSC-CMs whole-cell [Ca2+]i transients. In an immediate subsequent caffeine puff the caffeine-induced [Ca2+]i transient was completely omitted. The absent caffeine-induced signal at this stage is postulated to be a consequence of caffeine induced depletion of the SR Ca2+ store and the inability of the SR to accumulate Ca2+ as a result of the thapsigargin treatment.

IP3R expression, function, and contribution to whole-cell [Ca2+]i transients in hiPSC-CMs

IP3-mediated Ca2+ release presents a fundamental pathway for intracellular Ca2+ release in electrically non-excitable adult cells [40]. While, in adult cardiomyocytes IP3Rs contribution to cardiac physiology has remained elusive and controversial [9], [41], [42] they have been shown to play an important role during the process of cardiac development [23], [24], [42]. In fact, in the embryo the IP3R is reported to be the first expressed Ca2+ release channel [24], [42].

The IP3Rs have been reported to contribute to spontaneous activity in mouse ESC-CMs [25], [26] and are expressed and functional in hESC-CMs [18], [27]. To test for functionality and a potential contribution of an IP3-releasable Ca2+ pool to the modulation of Ca2+-handling in hiPSC-CMs we first examined the expression and localization of the IP3R at the protein level. Immunostainings of these hiPSC-CMs stained positive for IP3R with a strong subcellular distribution of the immunosignal around the nucleus (Figure 6A) in a similar manner to that observed in hESC-CMs [18], mouse ESC-CMs [26], and neonatal rat cardiomyocytes [28].

Next, to evaluate for IP3-releasable Ca2+ pool functionality and participation in the regulation of Ca2+-handling in hiPSC-CMs we tested the effect of IP3R blockade utilizing two different antagonistic approaches. First, to block IP3Rs we used the potent cell-permeate inhibitor 2-aminoethoxydiphenyl borate (2-APB) [25]. Application of 2-APB resulted in a significant dose-dependent diminution of whole-cell [Ca2+]i transients amplitude, as was also reported in human ESC-CMs under these conditions [18]. In addition, a slowing of whole-cell [Ca2+]i transients frequency was observed under the influence of 2-APB. Next we applied U73122, a phosopholipase C blocker (PLC). Blocking the activation of PLC inhibits a receptor-stimulated increase in the production of the second messenger IP3 required as a trigger for IP3R-mediated Ca2+ release [25], [27]. Superfusion of hiPSC-CMs with U73122 also significantly decreased whole-cell [Ca2+]i transients amplitude and frequency. A U73122 PLC inhibitory effect was also reported in mouse ESC-CMs [25]. These observations imply that an IP3-releasable Ca2+ pool is expressed and functional in hiPSC-CMs and that the resulting IP3R-mediated Ca2+ release contributes to the modulation of Ca2+-handling of these cells.

Potential clinical and research applications

The hiPSC technology has raised significant excitement with regards to its unique potential for regenerative medicine and for the study of various genetic disorders as well as for drug discovery and screening. In the current work we focused on the characterization of the Ca2+-handling properties of cardiomyocytes differentiated from hiPSCs and demonstrated that they share components that are present in adult cardiomyocytes, such as functional RyR-mediated SERCA-sequestering SR Ca2+ stores. Importantly, the results of this study showing similar properties in cardiomyocytes derived from different differentiation batches, from different hiPSCs clones, and from different hiPSCs lines may have important implications for their potential use for the aforementioned tasks. The hiPSC-CMs may serve as attractive cell-candidates for myocardial cell replacement therapy because of their inherent cardiac-specific properties and the potential for autologous therapy. Nevertheless, since functional compatibility between donor hiPSC-CMs and host myocardium is likely to contribute to an improved functional outcome of the cell engraftment as well as a reduction in potential pro-arrhythmic risk, detailed characterization of their Ca2+-handling characteristics is mandatory.

Similarly, phenotypic characterization of Ca2+-handling in these cells may be important if one wishes to utilize the iPSC technology for establishment of personalized in-vitro models of cardiac tissue for the development and testing of pharmacological compounds targeting these functional properties. Finally, detailed understanding of the Ca2+-handling properties of hiPSC-CMs generated from healthy individuals may be used as a future reference when studying E-C coupling in hiPSC-CMs derived from patients with genetic disorders involving Ca2+-handling. Such patient/disease-specific models can be established, for example, from families suffering from catecholinergic polymorphic ventricular tachycardia (CPVT) [43], [44], a potentially lethal disorder resulting from mutations in either the ryanodine receptor or calsequestrin.

Conclusion

In the present work we investigated basic Ca2+-handling components of hiPSC-CMs. Our results show that hiPSC-CMs display functional and loaded RyR-regulated intracellular Ca2+ stores. These stores can release Ca2+ via RyRs and can reload their content through SR Ca2+ uptake utilizing functional SERCA pumps. We present evidence showing the expression and functionality of inositol-1,4,5-trisphosphate receptors (IP3Rs). Furthermore, our findings demonstrate that the observed whole-cell [Ca2+]i transients in hiPSC-CMs depend on both sarcolemmal Ca2+ entry via L-type Ca2+ channels and on intracellular store Ca2+ release. Taken together hiPSC-CMs recapitulate functional key Ca2+ handling proteins that have been shown to be expressed and functional in mouse ESC-CMs [25], [26], [33], hESC-CMs [18], [19], [30], and adult cardiac tissue [9]. The results of the current study may have important implications for the potential applications of the iPSC technology in basic and translational cardiac research.

Supporting Information

Figure S1.

Whole-cell [Ca2+]i transients' requirement for Ca2+ influx via L-type Ca2+ channels as observed in cardiomyocytes derived from different hiPSCs clones and lines. Whole-cell [Ca2+]i transients recorded from (A): hIH1 clone 1, (B): hIH1 clone 2 and (C): hfib2-5 before (left) and after (right) application of nifedipine (1 µM). Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.s001

(PDF)

Figure S2.

Caffeine and Ryanodine-sensitive Ca2+ stores as displayed in cardiomyocytes derived from different hiPSCs clones and lines. (A) A line-scan presenting the effect of 20 mM caffeine puff application (noted by the arrows) in hIH1 clone 1 (left), hIH1 clone 2 (middle) and hfib2-5 (right). (B) Line-scan tracings of whole-cell [Ca2+]i transients recorded from hIH1 clone 1 (top), hIH1 clone 2 (middle) and hfib2-5 (bottom) under baseline conditions (left) and in the presence of 10 µM Ryanodine (right). (C) Dose-response curve for ryanodine (5, 10, and 20 µM) displayed as the percent decrease in [Ca2+]i transients amplitude from baseline values (n = 5, **p<0.01 when mean absolute values were compared to baseline values). Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.s002

(PDF)

Figure S3.

The effects of SERCA inhibition in cardiomyocytes derived from the hfib2-5 line. (A) Line-scan tracings of whole-cell [Ca2+]i transients in a representative hfib2-5 hiPSC-CM under baseline conditions (left) and after 30 minutes (right) of constant 10 µM thapsigargin exposure. (B) Caffeine-induced Ca2+ transients (denoted by arrows) under baseline conditions (upper), after 30 minutes (middle), and 32 minutes (lower) of thapsigargin exposure. Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.s003

(PDF)

Figure S4.

The effect of IP3R inhibition in cardiomyocytes derived from the hfib2-5 line. Whole-cell [Ca2+]i transients in representative hfib2-5 hiPSC-CMs before (left) and after (right) 2-APB (2 µM) application (A) and before (left) and after (right) U73122 (2 µM) application (B). Abbreviations: F/Fo, fluorescence (F) normalized to baseline fluorescence (Fo); IP3R, inositol-1,4,5-trisphosphate receptor; sec, seconds.

https://doi.org/10.1371/journal.pone.0018037.s004

(PDF)

Author Contributions

Conceived and designed the experiments: II SR JS LG. Performed the experiments: II SR IH IM LZD GA. Analyzed the data: II SR. Contributed reagents/materials/analysis tools: II IH IM LZD GA. Wrote the paper: II LG.

References

  1. 1. Takahashi K, Yamanaka S (2006) Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126: 663–676.
  2. 2. Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, et al. (2007) Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131: 861–872.
  3. 3. Yu J, Vodyanik MA, Smuga-Otto K, Antosiewicz-Bourget J, Frane JL, et al. (2007) Induced pluripotent stem cell lines derived from human somatic cells. Science 318: 1917–1920.
  4. 4. Thomson JA, Itskovitz-Eldor J, Shapiro SS, Waknitz MA, Swiergiel JJ, et al. (1998) Embryonic stem cell lines derived from human blastocysts. Science 282: 1145–1147.
  5. 5. Germanguz I, Sedan O, Zeevi-Levin N, Shtreichman R, Barak E, et al. (2009) Molecular characterization and functional properties of cardiomyocytes derived from human inducible pluripotent stem cells. J Cell Mol Med.
  6. 6. Haase A, Olmer R, Schwanke K, Wunderlich S, Merkert S, et al. (2009) Generation of induced pluripotent stem cells from human cord blood. Cell Stem Cell 5: 434–441.
  7. 7. Zhang J, Wilson GF, Soerens AG, Koonce CH, Yu J, et al. (2009) Functional cardiomyocytes derived from human induced pluripotent stem cells. Circ Res 104: e30–41.
  8. 8. Zwi L, Caspi O, Arbel G, Huber I, Gepstein A, et al. (2009) Cardiomyocyte differentiation of human induced pluripotent stem cells. Circulation 120: 1513–1523.
  9. 9. Bers DM (2002) Cardiac excitation-contraction coupling. Nature 415: 198–205.
  10. 10. Cannell MB, Cheng H, Lederer WJ (1995) The control of calcium release in heart muscle. Science 268: 1045–1049.
  11. 11. Galli GL, Taylor EW, Shiels HA (2006) Calcium flux in turtle ventricular myocytes. Am J Physiol Regul Integr Comp Physiol 291: R1781–1789.
  12. 12. Liu W, Yasui K, Opthof T, Ishiki R, Lee JK, et al. (2002) Developmental changes of Ca(2+) handling in mouse ventricular cells from early embryo to adulthood. Life Sci 71: 1279–1292.
  13. 13. Morad M, Goldman YE, Trentham DR (1983) Rapid photochemical inactivation of Ca2+-antagonists shows that Ca2+ entry directly activates contraction in frog heart. Nature 304: 635–638.
  14. 14. Viatchenko-Karpinski S, Fleischmann BK, Liu Q, Sauer H, Gryshchenko O, et al. (1999) Intracellular Ca2+ oscillations drive spontaneous contractions in cardiomyocytes during early development. Proc Natl Acad Sci U S A 96: 8259–8264.
  15. 15. Itzhaki I, Maizels L, Huber I, Zwi-Dantsis L, Caspi O, et al. (2011) Modelling the long QT syndrome with induced pluripotent stem cells. Nature.
  16. 16. Huangfu D, Maehr R, Guo W, Eijkelenboom A, Snitow M, et al. (2008) Induction of pluripotent stem cells by defined factors is greatly improved by small-molecule compounds. Nat Biotechnol 26: 795–797.
  17. 17. Park IH, Zhao R, West JA, Yabuuchi A, Huo H, et al. (2008) Reprogramming of human somatic cells to pluripotency with defined factors. Nature 451: 141–146.
  18. 18. Satin J, Itzhaki I, Rapoport S, Schroder EA, Izu L, et al. (2008) Calcium handling in human embryonic stem cell-derived cardiomyocytes. Stem Cells 26: 1961–1972.
  19. 19. Zhu WZ, Santana LF, Laflamme MA (2009) Local control of excitation-contraction coupling in human embryonic stem cell-derived cardiomyocytes. PLoS One 4: e5407.
  20. 20. Polsky A, Mel BW, Schiller J (2004) Computational subunits in thin dendrites of pyramidal cells. Nat Neurosci 7: 621–627.
  21. 21. Snir M, Kehat I, Gepstein A, Coleman R, Itskovitz-Eldor J, et al. (2003) Assessment of the ultrastructural and proliferative properties of human embryonic stem cell-derived cardiomyocytes. Am J Physiol Heart Circ Physiol 285: H2355–2363.
  22. 22. Sauer H, Theben T, Hescheler J, Lindner M, Brandt MC, et al. (2001) Characteristics of calcium sparks in cardiomyocytes derived from embryonic stem cells. Am J Physiol Heart Circ Physiol 281: H411–421.
  23. 23. Poindexter BJ, Smith JR, Buja LM, Bick RJ (2001) Calcium signaling mechanisms in dedifferentiated cardiac myocytes: comparison with neonatal and adult cardiomyocytes. Cell Calcium 30: 373–382.
  24. 24. Rosemblit N, Moschella MC, Ondriasa E, Gutstein DE, Ondrias K, et al. (1999) Intracellular calcium release channel expression during embryogenesis. Dev Biol 206: 163–177.
  25. 25. Kapur N, Banach K (2007) Inositol-1,4,5-trisphosphate-mediated spontaneous activity in mouse embryonic stem cell-derived cardiomyocytes. J Physiol 581: 1113–1127.
  26. 26. Mery A, Aimond F, Menard C, Mikoshiba K, Michalak M, et al. (2005) Initiation of embryonic cardiac pacemaker activity by inositol 1,4,5-trisphosphate-dependent calcium signaling. Mol Biol Cell 16: 2414–2423.
  27. 27. Sedan O, Dolnikov K, Zeevi-Levin N, Leibovich N, Amit M, et al. (2008) 1,4,5-Inositol trisphosphate-operated intracellular Ca(2+) stores and angiotensin-II/endothelin-1 signaling pathway are functional in human embryonic stem cell-derived cardiomyocytes. Stem Cells 26: 3130–3138.
  28. 28. Jaconi M, Bony C, Richards SM, Terzic A, Arnaudeau S, et al. (2000) Inositol 1,4,5-trisphosphate directs Ca(2+) flow between mitochondria and the Endoplasmic/Sarcoplasmic reticulum: a role in regulating cardiac autonomic Ca(2+) spiking. Mol Biol Cell 11: 1845–1858.
  29. 29. Dolnikov K, Shilkrut M, Zeevi-Levin N, Gerecht-Nir S, Amit M, et al. (2006) Functional properties of human embryonic stem cell-derived cardiomyocytes: intracellular Ca2+ handling and the role of sarcoplasmic reticulum in the contraction. Stem Cells 24: 236–245.
  30. 30. Liu J, Fu JD, Siu CW, Li RA (2007) Functional sarcoplasmic reticulum for calcium handling of human embryonic stem cell-derived cardiomyocytes: insights for driven maturation. Stem Cells 25: 3038–3044.
  31. 31. Bers DM, Perez-Reyes E (1999) Ca channels in cardiac myocytes: structure and function in Ca influx and intracellular Ca release. Cardiovasc Res 42: 339–360.
  32. 32. Fabiato A (1983) Calcium-induced release of calcium from the cardiac sarcoplasmic reticulum. Am J Physiol 245: C1–14.
  33. 33. Fu JD, Li J, Tweedie D, Yu HM, Chen L, et al. (2006) Crucial role of the sarcoplasmic reticulum in the developmental regulation of Ca2+ transients and contraction in cardiomyocytes derived from embryonic stem cells. FASEB J 20: 181–183.
  34. 34. Fu JD, Yu HM, Wang R, Liang J, Yang HT (2006) Developmental regulation of intracellular calcium transients during cardiomyocyte differentiation of mouse embryonic stem cells. Acta Pharmacol Sin 27: 901–910.
  35. 35. Weber A, Herz R (1968) The relationship between caffeine contracture of intact muscle and the effect of caffeine on reticulum. J Gen Physiol 52: 750–759.
  36. 36. Rousseau E, Smith JS, Meissner G (1987) Ryanodine modifies conductance and gating behavior of single Ca2+ release channel. Am J Physiol 253: C364–368.
  37. 37. Li J, Qu J, Nathan RD (1997) Ionic basis of ryanodine's negative chronotropic effect on pacemaker cells isolated from the sinoatrial node. Am J Physiol 273: H2481–2489.
  38. 38. Bassani JW, Bassani RA, Bers DM (1994) Relaxation in rabbit and rat cardiac cells: species-dependent differences in cellular mechanisms. J Physiol 476: 279–293.
  39. 39. Shannon TR, Ginsburg KS, Bers DM (2000) Potentiation of fractional sarcoplasmic reticulum calcium release by total and free intra-sarcoplasmic reticulum calcium concentration. Biophys J 78: 334–343.
  40. 40. Berridge MJ (1997) Elementary and global aspects of calcium signalling. J Physiol 499 (Pt 2): 291–306.
  41. 41. Marks AR (2000) Cardiac intracellular calcium release channels: role in heart failure. Circ Res 87: 8–11.
  42. 42. Moschella MC, Marks AR (1993) Inositol 1,4,5-trisphosphate receptor expression in cardiac myocytes. J Cell Biol 120: 1137–1146.
  43. 43. Katz G, Arad M, Eldar M (2009) Catecholaminergic polymorphic ventricular tachycardia from bedside to bench and beyond. Curr Probl Cardiol 34: 9–43.
  44. 44. Priori SG, Napolitano C (2005) Intracellular calcium handling dysfunction and arrhythmogenesis: a new challenge for the electrophysiologist. Circ Res 97: 1077–1079.