Despite distinctive functional and anatomic differences, a precise understanding of the cardiac interventricular differences in excitation–contraction (E–C) coupling mechanisms is still lacking. Here, we directly compared rat right and left cardiomyocytes (RVCM and LVCM). Whole-cell patch clamp, the IonOptix system, and fura-2 fluorimetry were used to measure electrical properties (action potential and ionic currents), single-cell contractility, and cytosolic Ca2+ ([Ca2+]i), respectively. Myofilament proteins were analyzed by immunoblotting. RVCM showed significantly shorter action potential duration (APD) and higher density of transient outward K+ current (Ito). However, the triggered [Ca2+]i change (Ca2+ transient) was not different, while the decay rate of the Ca2+ transient was slower in RVCM. Although the relaxation speed was also slower, the sarcomere shortening amplitude (ΔSL) was smaller in RVCM. SERCA activity was ∼60% lower in RVCM, which is partly responsible for the slower decay of the Ca2+ transient. Immunoblot analysis revealed lower expression of the cardiac troponin complex (cTn) in RVCM, implying a smaller Ca2+ buffering capacity (κS), which was proved by in situ analysis. The introduction of these new levels of cTn, Ito, and SERCA into a mathematical model of rat LVCM reproduced the similar Ca2+ transient, slower Ca2+ decay, shorter APD, and smaller ΔSL of RVCM. Taken together, these data show reduced expression of cTn proteins in the RVCM, which provides an explanation for the interventricular difference in the E–C coupling kinetics.

The right ventricle (RV) propels blood into the pulmonary circulation, where the hydraulic impedance is significantly lower than that of the systemic circulation. Consequently, despite the same amount of cardiac output from the two ventricles, the thickness of the RV free wall is lower than that of the left ventricle (LV). As a whole, the RV uses approximately one-fifth of the energy of the LV (Friedberg and Redington, 2014; Foschi et al., 2017). In terms of the work of cardiac cycle, the pressure–volume relationship (P-V loop) of the RV is described as a triangular or trapezoidal shape, with a short isometric period (Redington et al., 1988), different from the square shape in the LV.

The contractile properties of isolated cardiomyocytes from LV (LVCM) and RV (RVCM) have rarely been investigated, while demonstrating variable results in sarcomere length (SL) changes depending on the animal species and research conditions. A study of mice showed significantly smaller shortening of SL (ΔSL) in RVCM than in endocardial LVCM (Kondo et al., 2006). In rats, the analysis of SL changes revealed the tendency of slower kinetics in RVCM (Carneiro-Júnior et al., 2013). In contrast, canine RVCM showed no significant difference from LVCM in basal SL and ΔSL (Carneiro-Júnior et al., 2013; Molina et al., 2014).

Previous studies of the interventricular differences in the rodent cardiac action potential duration (APD) showed shorter APD in RVCM than LVCM (Watanabe et al., 1983; Knollmann et al., 2001). In addition, canine RVCM showed a deeper notch of phase I in their action potential (AP) than LVCM (Di Diego et al., 1996). Consistent with the shorter APD, the peak amplitudes of cytosolic Ca2+ concentration changes triggered by a single AP (change in intracellular Ca2+ concentration [Δ[Ca2+]i] or Ca2+ transient) were lower in the RVCM than in the endocardial LVCM of mice (Kondo et al., 2006). However, in rat cardiomyocytes, the Δ[Ca2+]i was not different between the RV and LV (Sathish et al., 2006; Carneiro-Júnior et al., 2013).

In addition to the upstroke amplitude, the decay kinetics of the Ca2+ transient determine the contractile properties of cardiomyocytes. In an early study of rat RVCM, slower decay of Ca2+ transient was observed along with the experimental evidence suggesting a slower rate of Ca2+ sequestration by the sarcoplasmic reticulum Ca2+-ATPase (SERCA). However, in more recent studies of rodent hearts, the decay of Ca2+ transient in RVCM was not different from that in LVCM (Carneiro-Júnior et al., 2013; Molina et al., 2014). The controversial results from previous studies suggested that another Ca2+ buffering component, such as Ca2+ binding troponin C (TnC), might be different between RVCM and LVCM.

In cardiac myocytes, Ca2+ regulates contraction by binding to the thin filament regulatory protein TnC. Along with TnC, the troponin complex (cTn) comprises TnI, which prevents the interaction between myosin and actin filaments, and TnT, which links the TnI–TnC complex to tropomyosin in the thin filament. Ca2+ binding with TnC causes a conformational shift in TnI and TnT, allowing myosin to bind to actin (Sweeney and Hammers, 2018). In addition to the regulation of contraction, Ca2+-binding TnC could also contribute to the cytosolic Ca2+ buffering capacity (κS) of cardiomyocytes. However, the comparison of myofilament expression between RV and LV cardiomyocytes is very rare, let alone the expression of TnC. A recent study of isolated myosin molecules suggested that the mechanical characteristics of working myosin (kinetic rates and the distribution of spatial orientation of myosin lever arm) were the same in both ventricles (Duggal et al., 2017). However, there have been few reports comparing the expression of myofilament components between the two ventricles except 2-D electrophoresis data from porcine and human heart tissues (Phillips et al., 2011; Su et al., 2015).

In this study, we aimed to elucidate the functional differences in excitation–contraction (E–C) coupling along with the expression of myofilament proteins between RVCM and LVCM from rats, focusing on (1) the AP shapes and relevant ionic currents, (2) AP-triggered Ca2+ transients with contraction–relaxation kinetics, (3) the expression of Ca2+-binding TnC and other cTn proteins, and (4) the Ca2+ buffering capacity and Ca2+ sequestration rate by the SR. The results demonstrate intriguing discordant properties between the levels of E–C coupling (APD, Δ[Ca2+]i, and ΔSL) in terms of biventricular differences. The immunoblot assay revealed lower expression of cTn; in addition, rigorous in situ analysis revealed a lower κS, consistent with the lower TnC in RVCM. The SERCA activity was found to be lower in RVCM, consistent with the slower decay rate of the Ca2+ transient. The novel findings regarding cTn and cytosolic κS in RVCM could explain the biventricular gaps between the APD, Δ[Ca2+]i amplitude, and ΔSL, which were successfully simulated by using a computational model of rodent cardiomyocytes.

Isolation of ventricular cardiomyocytes

Ventricular cardiomyocytes from Sprague-Dawley rats (RRID:RGD_1566457) were used. The study protocol followed the Guide for the Care and Use of Laboratory Animals published by the US National Institutes of Health (Publication No. 85-23, revised 1996) and also conformed to the Institutional Animal Care and Use Committee of Seoul National University (approval no. SNU-160128-1). 46 male rats were used. At the time of the experiments, the rats were 12 wk old and had mean (± SD) body weight of 348.4 ± 4.68 g.

Cardiomyocytes were isolated using a standard enzymatic dispersion technique, as described previously (Jin et al., 2012), with minor modifications. Briefly, rats were anesthetized with a mixture of ketamine (90 mg ⋅ kg−1, i.p.) and xylazine (10 mg ⋅ kg−1, i.p.). After monitoring the anesthesia depth with a toe pinch, the hearts were extracted and rapidly mounted onto the Langendorff apparatus at 37°C. Cells were dissociated by a protocol consisting of a constant-flow perfusion (8–10 ml ⋅ min−1) for (1) 8 min with Ca2+-free Tyrode solution; (2) 8 min with Ca2+-free Tyrode solution containing enzyme (collagenase 1 mg ⋅ ml−1 [Worthington Biochemical Co.], protease, 0.133 mg ⋅ ml−1, BSA 1.65 mg ⋅ ml−1, and Ca2+ 0.05 mM); and (3) a further 10-min digestion period with collagenase-containing Tyrode’s solution. In this study, the septum between the two ventricles was not used. Cardiomyocytes were separately harvested from both ventricles and resuspended in cardioplegic storage solution (for IonOptix, in mM; 120 NaCl, 5.4 KCl, 10 HEPES, 5 MgSO4, 5 Na-pyruvate, 5.5 glucose, 20 taurine, 29 mannitol, and 0.2 CaCl2, pH 7.4) or Kraft-Brühe (K-B) solution (for electrophysiology, in mM; 50 L-glutamate, 70 KOH, 55 KCl, 10 HEPES, 0.5 EGTA, 20 KH2PO4, 20 taurine, 20 glucose, and 3 MgCl2, pH 7.3). The proportion of rod-shaped cardiomyocytes ranged from 60 to 80%. Cardiomyocytes for physiological assays were used immediately, and cardiomyocytes for biochemical assays were centrifuged and stored at −80°C.

Electrophysiology

Freshly isolated cardiomyocytes were transferred into a bath mounted on the stage of an inverted microscope (Ti; Nikon). The bath (0.15 ml) was continuously perfused at a constant rate (5 ml ⋅ min−1), and voltage-clamp experiments were performed at room temperature (22–25°C). Borosilicated glass pipettes with a free-tip resistance of ∼2.5 MΩ were used for whole-cell and perforated whole-cell patch clamp. The series resistance, estimated by dividing the time constants of the capacitive current, was kept <10 MΩ in the whole-cell configuration and <16 MΩ in the perforated whole-cell configuration. To correct for cell size, the absolute current amplitudes were divided by the cell capacitance and expressed as pA ⋅ pF−1. The pipettes were connected to the CV 203BU head stage of the Axopatch 200B, a patch-clamp amplifier (Molecular Devices). pCLAMP software v10.6.2 and Digidata-1440A (Molecular Devices) were used to acquire data and apply command pulses. The recorded currents were sampled at 10 kHz and low-pass Bessel filtered at 5 kHz. The recorded data were processed using Clampfit v11.0.1 (Molecular Devices).

Single cardiomyocyte contraction and Δ[Ca2+]i measurements using the IonOptix system

The physiological properties of cardiomyocytes were measured using the IonOptix system (IonOptix). The contractility was measured by detecting the length of two edges with a contractility recording system at a 2–6-Hz frequency of field stimulation. Soft-edge software (IonOptix) was used to capture and analyze the changes in SL, and the results were given as fractional shortening. For [Ca2+]i measurement, cardiomyocytes were incubated with 2 µM fura-2-acetoxymethyl ester (fura-2-AM) for 15 min at room temperature. After sedimentation, the supernatant was removed, and cardiomyocytes were reintroduced into a perfusion solution containing 500 µM Ca2+ for 10 min. Then, the fura-2-AM–loaded cardiomyocytes were placed in a bath and excited by 360- and 380-nm filtered fluorescent light. The emitted signal was measured simultaneously with photomultiplier tubes. [Ca2+]i was calculated using the in situ calibration described below and the estimated dissociation constant (Kd) of fura-2-AM (Kd,fura-2). Measurements from ≥20 steady-state contractions were averaged for each cardiomyocyte for each stage of the experimental protocol. All experiments were performed in a temperature-controlled bath solution at 37 ± 1°C.

Isolation of membrane/cytosol fraction and myofilament fraction

The sampling for Western blotting (immunoblotting) was performed in two different ways, as described previously (Jang et al., 2015). The membrane/cytosol fraction was extracted from cardiomyocytes using lysis buffer containing 0.5 mM EGTA, 25 mM Tris-HCl, 150 mM NaCl, and 1% Triton X-100 with phosphatase and protease inhibitor cocktail (Roche) following centrifugation. To obtain the myofilament fraction, cardiomyocytes were washed with cold PBS and transferred to screw-cap tubes containing 2.5 mm zirconia/silica beads and cold buffer (60 mM KCl, 30 mM imidazole, 2 mM MgCl2, and 1% Triton X-100 with protease and phosphatase inhibitor cocktail). The cells in the lysis buffer were homogenized with a bead beater (Mini-Beadbeater-8; BioSpec Products) for 20 s, and the homogenates were centrifuged at 8,000 rpm for 10 min at 4°C. The supernatant was discarded, and the process was repeated at least six times. The remaining pellets (myofilament-enriched fraction) were resuspended in the high-salt buffer (Bio-Rad) and used in subsequent experiments.

Immunoblotting

Lysates were quantified using the Bradford method (Bio-Rad) and heated at 95°C for 5 min. Lysates were fractionated by SDS-PAGE and transferred to PVDF membranes (Bio-Rad) in 25 mM Tris, 192 mM glycine, 0.01% SDS, and 20% methanol. Membranes were blocked in 1× TBS containing 1% Tween-20 and 5% BSA (blocking solution) for 1 h at room temperature with gentle rocking. Membranes were then incubated overnight at 4°C with TnC (RRID:AB_306435; Abcam), TnI (RRID:AB_2206278; Cell Signaling), TnT (RRID:AB_261723; Sigma-Aldrich), tropomyosin (RRID:AB_261817; Sigma-Aldrich), desmin (RRID:AB_306653; Abcam), actinin (RRID:AB_476766; Sigma-Aldrich), and α-actin (RRID:AB_476695; Sigma-Aldrich) primary antibodies, SERCA2a (Abcam), phospholamban (RRID:AB_2617049; Badrilla), and phosphor-PLB (RRID:AB_310352; Millipore) followed by secondary antibodies after washing. Blots were developed using ECL Plus Western blotting detection reagents (Merck). The relative densities were calculated by normalizing each blot to actin. The number of experiments means each different animal. We excluded the results that showed large differences in internal control (10%) between LV and RV protein samples.

In situ analysis of the Ca2+ binding ratio and Ca2+ removal

For these experiments, a whole-cell patch clamp was conducted with simultaneous fura-2 fluorescence ratiometry (PTI). The Cs-pipette solution contained 50 µM fura-2 pentapotassium salt (Table 1, Cs-pipette). Light from a xenon arc was collected and filtered at 340 and 380 nm with 10 nm full width at half-maximum. The filtered light of the region of interest surrounding the cardiomyocyte was passed through the epifluorescence port of the inverted microscope and the CFI Super Fluor objective lens (Nikon). The emitted light was filtered at 510 nm and measured simultaneously using a photomultiplier tube (PTI).

To estimate the endogenous cytosolic buffer components, we defined the endogenous (intrinsic) Ca2+ binding ratio (κS) in terms of a hypothetical Ca2+ binding species “S” (Mathias et al., 1990). The κS and Ca2+-binding ratio of the total buffer including fura-2 (κB) were defined as follows:

(1)
(2)
where [S]tot is the net concentration of cytosolic Ca2+ buffer, [B]tot is the concentration of the total buffer including fura-2, Kd,S is the dissociation constant of the cytosolic buffer, and Kd,B is the dissociation constant of the total buffer. κB was determined by measuring the change in [Ca2+]i and the net amount of calcium influx during depolarization (i.e., the integral of the Ca2+ current; see Fig. 7; Berlin et al., 1994). For this purpose, after establishing the whole-cell configuration, the cell was depolarized with a ramp-like pulse from −70 to −40 mV (300 ms) to inactivate the voltage-gated Na+ current, followed by 100 ms of 10-mV depolarization to activate the voltage-operated Ca2+ channel (VOCCL). After confirming the steady-state amplitudes of the Ca2+ transient in normal Tyrode (NT), the SERCA inhibitor (2.5 µM thapsigargin) was applied for 5 min, followed by 20 mM caffeine to deplete the SR Ca2+ store. In addition, the extracellular solution was changed to a 0 Ca/0 Na solution (Table 1) to inhibit the sodium-calcium exchanger (NCX). Then, Ca2+ was added back (0 Na/2 Ca solution; Table 1) with simultaneous depolarization from −40 to 10 mV (200 ms, 1 Hz). To measure the Ca-binding ratio (κB), the change in [Ca2+]i induced by each depolarizing pulse under the above conditions (d[Ca2+]i) was measured using fura-2 ratiometry at room temperature. The change in bound calcium (d[BCa]) was calculated using Eq. 3:
(3)
The mechanisms responsible for Ca2+ removal were investigated using Ca2+ imaging as described by Díaz et al. (2004). We defined the time constant of decay of [Ca2+]i as τtran measured when depolarizing stimulation was applied. In caffeine treatment, the time constant was defined as τcaff that would reflect the Ca2+ removal kinetics except for SERCA. The time constant of decay with caffeine in 0 Ca/0 Na solution (τ0Na0Ca) was regarded to reflect the Ca2+ removal kinetics of plasma membrane Ca2+-ATPase (PMCA) and other slow components. To quantify the relative contribution of Ca2+ removal, the reciprocal of the time constant was taken and converted into rate constants (ktran, kcaff, and k0Na0Ca). In this model, the released Ca2+ from SR could be removed by (1) three active components (SERCA, NCX, and PMCA) and (2) endogenous Ca2+ buffers. As our present study demonstrated the lower endogenous Ca2+ buffering capacity in RVCM (see Fig. 8), we expected that the amount of Ca2+ binding to endogenous Ca2+ buffer would be larger in LVCM during the caffeine-induced Ca2+ transient, which also suggested the possibility of underestimating the SERCA activity of LVCM. To correct the error from difference in endogenous buffer, we assumed that total Ca2+ is a linear function of free Ca2+ at the relatively higher range where SERCA activity is measured (dotted line in Fig. 8, G and H). Although the linear relationship was not observed at low free Ca2+ concentrations (∼300 nM), the concentration of Ca2+ transient used for fitting to the exponential function was higher than resting levels, implying that the assumption was reasonable. Given this assumption, the rate constants of three Ca2+ removal machinery was multiplied by the buffer power of each ventricle, which was named “relative rate constant” (see Fig. 4 F).

In vitro Kd,fura-2 mimicking the cytosolic environment

The properties of fura-2, such as its Kd,fura-2 value when associated with Ca2+, are known to change in the cytosolic environment due to its binding with proteins (Konishi et al., 1988). Therefore, we conducted an in vitro calibration according to the method described by Berlin et al. (1994). The free Ca2+ concentrations of the calibration solutions were calculated with Slider, v2.0 (written by Chris Patton, Pacific Grove, CA), which used the equation from Fabiato and Fabiato (1979). To obtain the Kd,fura-2 value in the cytosolic environment, we used Ca2+-buffered (10 mM EGTA) KCl solution containing 283 mM sucrose or 20 mg/ml aldolase to reflect the viscosity and the fura-2–binding property of cytosolic proteins, respectively (Ca2+-calibration solution in Table 1). The viscosity–protein and –sucrose relationships were measured using an Ubbelohde-type viscometer under the assumption of a Newtonian fluid (Fig. S1 A). To minimize the error factor from the depth of the bath solution, a calibration solution constrained between two round coverslips (diameter 18 mm) was used (Fig. S1 B). From the amount of the calibration solution and the size of coverslips, we estimated the distance between the coverslips to be 120–140 µm, which was ∼10 times the average thickness of cardiomyocytes, 13 µm (Satoh et al., 1996). To obtain a fluorescence light intensity similar that of fura-2 loaded into cardiomyocytes, the concentration of fura-2 in the in vitro calibration was lowered to 5 µM. We then measured Kd,fura-2 in three types of calibration solutions (Fig. S1 C): (1) standard calibration solution (normal), (2) viscosity-corrected calibration (sucrose), and (3) protein-containing calibration solution (protein). Because the Kd,fura-2 values of the standard and viscosity-corrected calibration solutions were significantly lower than those of the protein-containing solutions (Fig. S1 D), we used the Kd,fura-2 value of the protein-containing calibration, in accordance with previous measurement of Berlin et al. (1994). In this study, the in situ Kd,fura-2 of fura-2 was 381.8 nM.

In situ calibration of fura-2 fluorescence ratiometry

It is well known that there are several problems in calibrating Ca2+ concentration in cardiomyocytes, including hypercontraction of cells at [Ca2+]i >1 µM, which distorts the path length and membrane integrity (Cheung et al., 1989). To overcome this problem, we measured the calibration parameters, Rmin, Rmax, and β, in separate experiments according to the lanthanum-method described by Borzak et al. (1990). Conversion factors were measured using a cuvette-type spectrophotometer with an identical right source and calculated as C1 = Rmax,Ca/Rmax,La and C2 = F380max,Ca/F380max,La (Fig. S2). These La3+-derived proportionality constants were determined to convert the fluorescence signal of La3+ to that of Ca2+.

Determination of fura-2–accessible cytosolic volume and the effective [fura-2] in situ

There is a linear correlation between membrane capacitance (Cm) and total cellular volume (voltc), and the ratio Cm/voltc was 6.76 pF ⋅ pl−1 in 12-wk-old adolescent rats (Satoh et al., 1996). Because mitochondria and other organelles occupy ∼35% of total cellular volume of cardiomyocytes, we estimated the cytosolic cell volume accessible to fura-2 (volac) as 0.65 voltc (Berlin et al., 1994). Fura-2 dialyzed into the cells could be partly bound to proteins (e.g., myofilaments) while maintaining the Ca2+-binding property. Therefore, it is necessary to estimate their effective cytosolic concentration of fura-2. At the end of every experiment, the fluorescence intensity of the cell at the isosbestic wavelength of fura-2 was measured and normalized to the accessible volume (Fcell/volac). In addition, the fluorescence of a glass capillary with a 20-µm inner diameter containing 50 µM of fura-2 was measured when placed in the same position as the cell within the region of interest of known length to obtain a standard fluorescence per unit volume (Fstandard) at the isosbestic wavelength. The effective net concentration of fura-2 ([fura2]tot) was calculated as follows:
where Φ is the ratio of fluorescence intensity in the presence and absence of aldolase at pCa 7.0.

Solution and chemicals

Solutions for electrophysiology, Ca2+ measurement, and the IonOptix system are listed in Table 1. We used three external solutions: NT, 0 Na/0 Ca, and 0 Na/2 Ca. Four types of pipette solutions were used depending on the experimental conditions (AP-, VOCCL-, K-, and Cs-pipette). All drugs and chemicals were obtained from Sigma-Aldrich.

Data presentation and statistics

For analysis of the activation and inactivation kinetics of the current of VOCCL (ICa,L) and transient outward K+ current (Ito), the conductance was normalized to the maximal conductance and fitted by a Boltzmann equation. The time constants of inactivation, τ, were obtained by fitting the current trace with a dual exponential function. Data are presented as mean ± SD, with n denoting the number of cells and N denoting the number of animals. Data were compared using a nested t test was where appropriate, taking into account n cells and N animals (Sikkel et al., 2017; Eisner, 2021). Current amplitudes of Ito and ICa,L were assessed using MANOVA in Origin software (MANOVA app, v1.22). Differences were considered significant when P < 0.05. Asterisks were used in the graphs to indicate significance of statistical inference (*, P < 0.05; **, P < 0.01; ***, P < 0.001; and ****, P < 0.0001). Curve fitting with the least-squares method was performed using Origin (Microcal Software) or hand-written Python software.

Computational simulation of AP

Mathematical models of rat ventricular AP described by the Hodgkin–Huxley-type differential equation were used (Pandit et al., 2001). The model consists of three intracellular compartments: myoplasm, junctional SR, and network. The total ionic current (Iion) is given by
where INa is the fast inward Na+ current, ICa,L is the VOCCL current, Ito is the transient outward K+ current, Iss is the steady-state outward K+ current, If is the hyperpolarization-activated funny current, IK1 is the inward rectifier K+ current, IB is the background current, INaK is the Na+-K+ pump, ICaP is the Ca2+ pump, and INaCa is the NCX.
In the original Pandit model, the Ito current is modeled as
where gt is the maximum conductance for Ito, r is the activation gating variable, a and b are proportionality coefficients, V is the membrane potential, EK is the reversal potential of K+, and s and sslow represent fast and slow inactivation gating variables, respectively.

Computational simulation of [Ca2+]i and contraction

The Ca2+-buffering effect of the troponin system, as represented by Itropn, was used to calculate free cytosolic [Ca2+] in the mathematical model of rat ventricular AP (Pandit et al., 2001). We replaced the simple Ca2+-buffering troponin model by Pandit et al. (2001) with the cardiac muscle model developed by Negroni and Lascano (2008) to calculate free cytosolic [Ca2+] as well as SL and force. Systolic and diastolic [Ca2+] by those two models were in the similar range (systolic, 400–600 nM; diastolic, ∼100 nM). The model equations are as follows:
SL and force were calculated as
These variables are listed in the List of abbreviations, and the numerical values of the parameters are listed in Table 2. The diagram of the troponin states described by Negroni and Lascano (2008) was used in this study with minor modification (Fig. S3).

Online supplemental material

Fig. S1 shows the methods of in vitro calibration for measuring Kd,fura-2. Fig. S2 shows the spectra of fura-2-Ca2+ and fura-2-La3+ and converting factors between two measurements. Fig. S3 shows the diagram of troponin states used in this study. Fig. S4 shows the AP analysis for each animal. Fig. S5 shows the inactivation time constants of Ito and Ica,L. Fig. S6 shows the Ca2+ transient analysis for each animal. Fig. S7 shows the immunoblotting of cTn and tropomyosin while varying the total loading protein amount. Fig. S8 shows the sarcomere shortening analysis for each animal. Fig. S9 shows the changes in sarcomere shortening pattern according to the stimulation frequency. The box-and-whisker plots in Figs. S4, S6, and S8 display the five-number summary of a set of data including the minimum, first quartile, median, third quartile, and maximum.

Electrophysiological properties

The APs of RVCM and LVCM were triggered by a 6-ms square pulse (∼1.0 nA of current injection, 2 Hz) under the nystatin-perforated mode of current-clamp (Fig. 1 A). In separate experiments under the conventional whole-cell mode, electrical membrane capacitances (Cm) were measured, which showed a larger Cm in LVCM than in RVCM (Fig. 1 B). All the parameters analyzed in the AP recordings were obtained in the nystatin-perforated condition (Fig. 1, C–M). Peak amplitudes, time to reach peak depolarization (peak time), resting membrane potential (RMP), maximum increasing speed, and time of the maximum increasing speed were not different between LVCM and RVCM (Fig. 1, C–G). In contrast, the repolarization was more rapid in RVCM; and higher maximum decreasing speed (Fig. 1 H); shorter time of maximum decreasing speed from the stimulus (MdT) and from the peak time (MdT from peak; Fig. 1, I and J), and shorter APD at 10, 50, and 90% of repolarization (APD10, APD50, APD90) were seen in RVCM than in LVCM (Fig. 1, K and M). Six and five Sprague-Dawley rats were used for measuring capacitance and AP, respectively. The statistical significance between RVCM and LVCM was analyzed by nested t test (Fig. S4).

Because the faster repolarization suggested larger voltage-dependent K+ conductance, we measured the K+ currents activated by depolarizing step pulses from −90 mV holding voltage under the conventional whole-cell voltage-clamp. A VOCCL blocker (nifedipine, 10 µM) was added to the bath solution. Fast activating transient outward currents (Ito) and sustained outward currents were observed (Fig. 2 A). The amplitudes of Ito—that is, the inactivating component of outward current—were normalized to Cm and plotted against the test voltages (I/V curves; Fig. 2 B). Ito was larger in RVCM than LVCM (131 ± 5.8% at 60 mV, P < 0.0001). Voltage-dependent inactivation of Ito was analyzed using a two-step voltage-clamp protocol. A common test pulse to 30 mV (200 ms) was applied at the end of the conditioning step (1 s) to different potentials (inset of Fig. 2 C). Inactivation was determined as the ratio of the test current amplitude to the maximum test current. The voltage dependence of inactivation was fitted by a Boltzmann function, providing a half-inactivation voltage (V1/2,inact) and slope factor (k). The V1/2,inact and k values were not significantly different between RVCM and LVCM (Fig. 2 C). The kinetics of the inactivating components were further analyzed by fitting to double exponential functions. The time constants for the fast and slow components did not differ between RVCM and LVCM (Fig. S5, A and B).

The current of VOCCL (ICa,L) generates prolonged depolarization of the cardiac AP, which determines the APD; therefore, we measured the ICa,L using Cs-pipette solution under the nystatin-perforated patch clamp condition to minimize the decay of ICa,L. Steplike depolarizations from −40 mV holding voltage were applied. Voltage-dependent activation of inward currents with relatively slow inactivation was observed (Fig. 2 D). The peak amplitudes normalized to Cm were plotted against the test voltages, and the I/V curves did not differ between RVCM and LVCM (Fig. 2 E). Steady-state voltage-dependent inactivation was analyzed using a two-step voltage clamp protocol. The pulses were applied to a test potential of 10 mV (300 ms) from various conditioning pulses (−80 mV to 20 mV, 300 ms; inset of Fig. 2 F). The voltage dependence and kinetics of ICa,L inactivation did not differ between RVCM and LVCM (Fig. 2 F and Fig. S5, C and D).

Using a mathematical electrophysiology model of rat ventricular cardiomyocytes (Pandit et al., 2001), we simulated the effects of the higher Ito in RVCM (Fig. 2, dotted lines). The Pandit model simulating the cardiac AP of the “left” ventricular AP was modified according to the experimental data of the RV. As the measured parameters representing the kinetics of Ito did not differ from those of the Pandit model, we modified only the conductance of Ito to reflect the measured values. It was also notable that the ICa,L of the Pandit model, 7.8 ± 0.2 pA ⋅ pF−1 at 10 mV (Katsube et al., 1998), was smaller than our experimental data, 9.0 ± 0.78 pA ⋅ pF−1 at 10 mV (Fig. 2, D and E). Therefore, to reflect this difference, we modified the maximum conductance for ICa,L (gCaL in the original model) in our RVCM and LVCM models. The calculated APD of rat RVCM was shorter than that of LVCM (Fig. 2 G), similar to the experimental data (Fig. 1 A and Table 3).

The amount of net Ca2+ influx would also be limited by the shorter APD inducing faster termination of ICa,L. However, the peak amplitude of ICa,L during the actual AP might not be smaller or could even be larger in RVCM with larger Ito, owing to the electrical driving force for the Ca2+ influx through the not-yet-deactivated VOCCL (Harris et al., 2005). Consistent with the latter possibility, it has been reported that a loss of the notch due to reduction of Ito contributes to the reduced Ca2+ influx in failing hearts of larger mammals (Cooper et al., 2010). To elucidate the actual Ca2+ influx via VOCCL, we compared the ICa,L induced by the AP-clamp protocol in RVCM and LVCM. In fact, the peak amplitude of inward ICa,L became higher, and the decay became faster, when applying the RVCM-type versus the LVCM-type AP clamp. Overall, the cumulative Ca2+ influx—that is, mathematical integration of the inward current per membrane capacitance (∫ICa,LdtCm−1)—was smaller when applying the RVCM-type AP (Fig. 2 H).

Analyses of Ca2+ transient

The Ca2+ transient of intact cardiomyocytes triggered by repetitive electrical field stimulation (2 Hz) was observed in cells loaded with fura-2-AM (Fig. 3 A, representative traces). Summarized results from LVCM and RVCM are shown as bar graphs (Fig. 3, B–G). The [Ca2+]i during the diastolic period did not differ between the RVCM and LVCM (Fig. 3 B). Considering the shorter APD and the smaller Ca2+ influx via VOCCL (Figs. 1 and 2), it was anticipated that Δ[Ca2+]i would be smaller in RVCM. However, the peak [Ca2+]i and Δ[Ca2+]i were not different from those of LVCM, indicating a mismatch between AP and Ca2+ transient (Fig. 3, C and D). Furthermore, despite the shorter APD, the decay of the Ca2+ transient was slower in the RVCM than in the LVCM (Fig. 3, E–G). 14 animals were used for measuring Ca2+ transient, and the four parameters analyzed in each animal are depicted in Fig. S6.

In rat ventricular cardiomyocytes, SERCA is mainly responsible for the decay of the Ca2+ transient (Bassani et al., 1994). Therefore, we analyzed SERCA activity in RVCM using whole-cell patch clamp combined with fura-2 ratiometry. Under the current-clamp configuration, AP-triggering current injections (∼1.0 nA, 1 Hz) were applied to confirm that the steady-state Ca2+ transient reflected the responses of the Ca2+-loaded state of the SR (Fig. 4 A, stimulation). Three Ca2+ removal mechanisms, SERCA, NCX, and PMCA, eliminated intracellular Ca2+ into SR and extracellular space in this condition, and the time constant τtran reflected the sum of the rates of Ca2+ removal by the three mechanisms (Fig. 4 B, stimulation). After stopping stimulation, we immediately applied 20 mM of caffeine to induce Ca2+ transient (Fig. 4 A, caffeine). In this condition, where SERCA activity was nullified, NCX and PMCA were involved in the decay of caffeine-induced calcium transient, which could be fitted to a single exponential function (τcaff; Fig. 4 B, caffeine). As the SR was depleted by the previous caffeine application, the cardiomyocytes were then restimulated by current injection until steady state was achieved (Fig. 4 A, second stimulation). The cardiomyocytes were then superfused with 0 Na/0 Ca solution to inhibit the NCX, and the second application of caffeine induced the Ca2+ transient with a slow decay speed (τ0Na0Ca, Fig. 4 A). As the SERCA and NCX were inhibited by caffeine and 0 Na/0 Ca bath solution, the decay kinetics of Ca2+ transient would be mainly affected by PMCA (Fig. 4 B, right panel). The decay of the Ca2+ transient of this experiment could be fitted to single exponential function (Fig. 4 B).

The time constant reflecting all the three Ca2+ removal mechanisms (τtran) was significantly larger in the RVCM (Fig. 4 C), while the time constant was almost identical when SERCA was neglected (τcaff, Fig. 4 D), implying that the slower Ca2+ decay of RVCM was due to the lower SERCA activity. From the three rate constants (ktran, kcaff, and k0Na0Ca) obtained by taking the reciprocal of time constants (τtran, τcaff, and τ0Na0Ca), the relative contributions made to Ca2+ removal by SERCA, NCX, and PMCA (relative rate constant) were determined using the protocol described by Díaz et al. (2004), with correction for the plausible error from the difference in endogenous buffer (see In situ analysis of the Ca2+ binding ratio…). The total rate of Ca2+ removal was higher in LVCM than RVCM, mostly because of the 2.23-times-higher relative rate of SERCA in LVCM (Fig. 4 F). The relative NCX and PMCA activity were also lower in RVCM (Fig. 4 F). However, since >89% of the total calcium removal mechanism was performed by SERCA [(ktrankcaff)/ktran], the difference of NCX and PMCA between two ventricles seems to have a partial effect on the total Ca2+ removal rate.

Molecular mechanism of the difference in SERCA activity

To investigate the underlying mechanisms of the observed difference in SERCA activity, we compared the protein amount of SERCA2a, which actively sequesters Ca2+ back into the SR lumen to remove intracellular Ca2+ (Sathish et al., 2006). When assessed by quantification of bands, the protein level of SERCA2a was not significantly different in the two ventricles (Fig. 5). Because the small difference that was not statistically significant could not account for the difference in SERCA activity, we focused on the endogenous inhibitor of SERCA, PLB, and its phosphorylated form (p-PLB), which relieves the inhibitory function of PLB (Tada and Katz, 1982; Simmerman et al., 1986). We compared the expression of PLB, which was normalized to the expression of internal loading control. The amount of PLB was not different in the two ventricles, while that of p-PLB was lower in RVCM. The ratio of p-PLB to PLB was 0.40 ± 0.039 versus 0.28 ± 0.065, also significantly lower in RVCM (Fig. 5 E).

Ca2+-binding proteins of myofilament components

Although the lower SERCA activity could explain the slower Ca2+ decay of RVCM, the similar Ca2+ peaks despite the shorter APD suggested additional differences in the Ca2+-buffering components of RVCM. We hypothesized that the Ca2+-binding myofilament proteins, such as TnC, might be different between RVCM and LVCM. TnC is the major Ca2+-binding protein in the cytosol of cardiomyocytes (Smith and Eisner, 2019). The immunoblot assay of the myofilament fraction revealed that the expression levels of TnC, TnI, and TnT were commonly lower in RVCM compared with LVCM (Fig. 5). In contrast, the levels of tropomyosin, desmin, actinin, and myosin binding protein C did not differ between RVCM and LVCM.

Overloading of samples is a widespread problem, termed membrane saturation, that may compromise the reliability of quantitative analysis, especially in enriched samples (Ghosh et al., 2014; Pillai-Kastoori et al., 2020). To overcome this limitation, we performed immunoblot experiments while loading various amounts of protein (1–5 µg) to find the range of protein amount that does not cause membrane saturation (Fig. S7). In the case of TnI, TnC, and tropomyosin, the signal of protein bands was not saturated up to 5 µg of protein, while that of TnT showed a saturated signal from 4 µg of protein (Fig. S7 B). The expression of cTn was lower in RVCM at the unsaturated ranges of protein amount. All the signals of blotting in this experiment were normalized to the total signal of protein using Ponceau S staining as an internal loading control (Gilda and Gomes, 2013; Sander et al., 2019).

Ca2+ binding ratio of the endogenous calcium buffer (κs) in RVCM

Cytosolic Ca2+ buffers in cardiomyocytes are composed of various components, including proteins such as TnC, calmodulin, and myosin. The major Ca2+ buffers in the physiological range of Ca2+ concentration (100–1,000 nM) are regulatory Ca2+-specific binding sites of TnC and active sequestration by SERCA. The other constituents (calmodulin and ATP) are quantitatively small or have dissociation constants that are widely different from the physiological [Ca2+]i (Ca2+/Mg2+-binding site of TnC and myosin light chain; Smith and Eisner, 2019). We analyzed the κS values of the RVCM and LVCM after excluding the factors from calcium-induced calcium release (CICR), SERCA, and NCX using the SERCA inhibitor (thapsigargin, 2.5 µM), RYR activator (caffeine, 20 mM), and Na+-free bath solution (Fig. 6 and Fig. 7 A; see Eq. 1 in Materials and methods).

When depolarizing pulses (from −40 to 10 mV, 100 ms, 1 Hz) were applied in 0 Na/2 Ca solution, a cumulative increase in [Ca2+]i ≤2 µM was observed during the 8–10 consecutive pulses (upper panel of Fig. 7 B). Because the Ca2+ influx via VOCCL was the sole source of Δ[Ca2+]i, the integral of calcium currents was obtained (Fig. 7 B, bottom panel). The integrated Ca2+ influx was divided by the accessible cell volume (volac, in Materials and methods) to obtain the calculated net increase in [Ca2+] (Δ[Ca2+]tot,cal). The median values of [Ca2+]i before and immediately after the individual pulses were measured. The relationship between κB and the median [Ca2+] was plotted (circles in Fig. 8, A and B), and the κB and [B]tot values were obtained by fitting Eq. 2 (see Materials and methods).

In the above experimental measurement, it should be noted that when using fura-2, an “added-buffer” condition (Neher, 1995), the contribution from fura-2 (Furatot) has to be excluded from [B]tot to deduce the endogenous calcium buffer, [S]tot (Fig. 8, C–E). Because we know the dissociation constant of fura-2 (Kd,fura-2) from the in vitro calibration mimicking the cytosolic environment (see Materials and methods; Fig. S1) and the effective total concentration of fura-2 from Eq. 3 ([fura-2]tot; Fig. 8 D), the correlation between the Ca2+-binding ratio of fura-2 (κfura-2) and the corresponding median [Ca2+] could be fitted to a differential form of the Michaelis–Menten equation (square symbols in Fig. 8, A and B). Finally, the differences between κB and κfura-2, i.e., κS could be obtained by subtraction, the result of which was fitted by Eq. 1 to calculate the two parameters of the endogenous buffer: the total concentration of the intrinsic buffer ([S]tot) and the dissociation constant of the intrinsic buffer (Kd,S). In accordance with the expression level of myofilament proteins, [S]tot of RVCM was actually smaller than that of LVCM (195.8 ± 19.61 µM versus 127.5 ± 21.09 µM; Fig. 8 E). The reported dissociation constant of TnC (Kd,TnC < 0.6 µM, from Smith and Eisner, 2019) is smaller than that of measured intrinsic buffer, which was 0.90 ± 0.156 µM in LVCM. Therefore, the calcium affinity of the endogenous buffer would be lower in RVCM owing to the lower proportion of troponin among the total intrinsic buffer. Consistently, the Kd,S values were larger in RVCM than in LVCM (Kd,S of RVCM = 1.21 ± 0.239 µM). Using the parameters obtained from the experiments (κB, κS, κfura-2, [B]tot, [fura-2]tot, [S]tot), we calculated the bound Ca2+ (calculated bound Ca2+) from the median [Ca2+] using the Michaelis–Menten equation. The calculated buffering power of RVCM was lower than that of LVCM (Fig. 8, G and H).

SL changes and the kinetics of contraction relaxation

The above differences in the factors determining the Ca2+ transient kinetics (e.g., lower SERCA activity and endogenous Ca2+ buffer) suggested differential features of contraction–relaxation between RVCM and LVCM. In the IonOptix experiment, we analyzed the SL changes induced by repetitive electrical field stimulation (2 Hz; Fig. 8 A). 20 animals were used for measuring SL, and 6 parameters analyzed for each animal are depicted in Fig. S8. The basal SL of RVCM during the diastolic phase appeared to be longer than that of LVCM, but not statistically significantly when nested t test was applied (Fig. 9 C, P = 0.190). The average maximum shortening in 10 consecutive pulses (ΔSL) was smaller in RVCM (Fig. 9 D). The percent normalized values (ΔSL/SL %) were also smaller in RVCM than in LVCM (Fig. 9 E). To clarify the difference in relaxation speed, the relative SL changes normalized to the peak SL were analyzed (Fig. 9 B). The relaxation speed was lower in RVCM (Fig. 9, F–H), which seems to be consistent with the slower decay of the Ca2+ transient in RVCM (Fig. 3, E–G).

The physiological heart rate of rat is faster (5–6 Hz) than the pacing rate in the present study (2 Hz). To confirm whether the different pacing rates might have induced the different contractile properties between RVCM and LVCM, the SL analysis was conducted at different pacing frequencies (2–6 Hz), where the smaller ΔSL and the slower relaxation speed in RVCM were consistently observed (Fig. S9).

Mathematical simulation of the Ca2+ transient and contraction

Finally, we recapitulated the experimental findings from the Ca2+ transient and the single-cell contraction (half-SL change, ΔSLhalf) by using a combinational computational model of rat ventricular cardiomyocytes constructed from two mathematical models (Pandit et al., 2001; Negroni and Lascano, 2008). Because the models were based on the experimental values obtained from LVCM, the three parameters were modified: the larger Ito, the lower SERCA activity, and the lower cTn expression (Fig. 10; for the AP simulation, see Fig. 2). To consider the kinetics of Ca2+ binding with TnC, we also calculated the differential parameter expressing the flux of Ca2+ to the cTn (dJTRPN/dt) according to the Itropn in Negroni and Lascano (2008) (Fig. 10 B).

When a higher Ito (gto ⋅ 1.31) was introduced, all three amplitudes (Ca2+ transient, dJTRPN/dt, and ΔSLhalf) were decreased (Fig. 10 A, upper panels), which was consistent with the effect of the RVCM AP-clamp on ICa,L (Fig. 2 H). The combined lowering of SERCA activity (ISERCA ⋅ 0.44) with the higher Ito slowed the decay of the Ca2+ transient and ΔSLhalf, resulting in their amplitudes being only slightly increased (Fig. 10, middle panels, blue traces). In contrast, the combined lowering of cTn (cTn ⋅ 0.65) with a higher Ito significantly increased the amplitude of the Ca2+ transient while decreasing dJTRPN/dt and ΔSLhalf (Fig. 10, middle panels, green lines). By combining the changes in the three parameters, Ito, SERCA, and cTn, the simulated Ca2+ transient and ΔSLhalf were close to the experimental data of the RVCM when compared with the simulated LVCM (Fig. 10, lower panels).

In addition to the previously known differences in electrophysiology (larger Ito) and lower SERCA activity in RVCM (Afzal and Dhalla, 1992; Molina et al., 2016), our present study demonstrated novel aspects of biventricular differences regarding E–C coupling, the lower expression of cTn that seems to underlie the lower Ca2+ buffering capacity and smaller contraction in RVCM. The lower SERCA activity and cTn in RVCM could explain the apparent mismatches between the AP versus Ca2+ transient as well as the Ca2+ transient versus ΔSL, which was reproduced by mathematical simulations.

Mismatch between AP and Ca2+ transient

The whole-cell patch clamp showed shorter APD and larger Ito in RVCM than in LVCM (Fig. 1 and Fig. 2, A–C). These findings are consistent with those of previous studies on rodent cardiomyocytes (Di Diego et al., 1996; Brunet et al., 2004). Because the ICa,L was not different (Fig. 2, D–F), the higher Ito seems to be responsible for the faster repolarization in RVCM, which was also reproduced by the mathematical electrophysiological model of rat cardiomyocytes (Fig. 2 G). Heterogeneity of electrical properties has also been observed in LVCM of the epi- and endocardium (Watanabe et al., 1983; Kondo et al., 2006). In fact, a study comparing RVCM with epi- and endocardial LVCM showed that the AP configuration of RVCM was similar to that of epi-LVCM (Kondo et al., 2006). Regrettably, we did not rigorously divide the LVCM from the epi- and endocardium. Nevertheless, the consistent statistical significance of the different parameters in AP and Ito suggested that the LVCM were a relatively homogeneous group close to the endocardial myocytes (Figs. 1 and 2).

Despite the shorter APD and smaller net Ca2+ influx via VOCCL, the amplitude of the Ca2+ transient was not different between RVCM and LVCM (Fig. 3). Because the amount of Ca2+ released from the SR (CICR) is mainly responsible for the final amplitude of Ca2+ transient in cardiomyocytes (Eisner et al., 2017), one has to also consider the efficiency of CICR when comparing the biventricular difference. Unfortunately, we could not directly evaluate the CICR efficiency or activity of RYR channels. Because the inactivation of VOCCL is largely affected by the [Ca2+] of the junctional space near RYR, an indirect estimation from the inactivation speed of ICa,L in the perforated whole-cell patch clamp could be provided. The identical time constants of inactivation kinetics did not support the hypothesis that CICR efficiency was higher in RVCM. Therefore, the SR Ca2+-release via CICR could be smaller in the beating RVCM than in LVCM because of the smaller Ca2+ influx during the shorter AP.

Taking these findings together, the more plausible explanation for the mismatch between the APD and the Ca2+ transient was the lower level of endogenous Ca2+ buffer and removal mechanism. In cardiomyocytes, it is known that only 1% of the total calcium increase could be experimentally detected as free Ca2+; 99% is bound to the cytosolic buffers or removed by SERCA, NCX, and PMCA (Fabiato, 1983; Hove-Madsen and Bers, 1993). Among the latter Ca2+ removal mechanisms, SERCA predominantly operates in rodent cardiomyocytes (Bassani et al, 1994; Díaz et al, 2004). A previous study using rat myocardium demonstrated slower SERCA in the SR vesicles isolated from RVCM than in those isolated from LVCM (Sathish et al., 2006). A similar finding was also confirmed in the present study using intact cardiomyocytes (Fig. 4), and the underlying mechanism of difference in SERCA activity was due to the difference in the phosphorylation states of PLB, which was lower in RVCM (Fig. 5). However, the lower SERCA activity appears to be more responsible for the slower rate of the Ca2+ transient decay than the instantaneous increase, which could also be simulated by mathematical modeling (Fig. 10).

Among the cytoplasmic proteins of cardiomyocytes, TnC could provide quantitatively effective Ca2+ buffering capacity in the physiological range of Δ[Ca2+]i (Pan and Solaro, 1987; Smith and Eisner, 2019). The Ca2+-binding site of TnC is classified into two categories: Ca2+-specific site and Ca2+/Mg2+-binding site (Robertson et al., 1981). Because the Ca2+/Mg2+-binding site with low Kd values (0.0195 µM, from Pan and Solaro, 1987) would be largely occupied by Ca2+ or Mg2+ at their resting levels, binding with the Ca2+-specific site in TnC would change dynamically during the cardiac cycle. In this study, the expression of TnC, along with TnI and TnT, was actually lower by ∼30% in RVCM than LVCM (Figs. 6 and S7). Based on the buffering properties of cTn, we propose that the differential Ca2+ buffering capacity could explain the mismatch between APD and the amplitude of the Ca2+ transient.

The apparently lower ratio of cTn over other myofilament proteins was an unexpected finding. According to the recent cryo-electron microcopy study of the cardiac myofilament structure, cTn attaches to a specific position of tropomyosin at every seventh actin monomer. Therefore, the ratio of actin, tropomyosin, and cTn would be a tight stoichiometry of 7:1:1 (Oda et al., 2020; Yamada et al., 2020). In this respect, the lower cTn in RVCM could imply a discordance between the amount of tropomyosin and cTn and putative instability of the myofilament complexes. Unfortunately, our data from immunoblotting experiments could not provide a reasonable explanation for the controversy. Nevertheless, a previous 2-D electrophoresis analysis of porcine heart also suggested the relatively lower level of TnC than tropomyosin in the RV than LV (Phillips et al., 2011). In addition, a differential in-gel electrophoresis combined with mass spectrometry study of human cardiac tissues suggested higher tropomyosin in comparison with TnI in LV than RV (Su et al., 2015). The studies including our present study suggested a difference in cTn/tropomyosin between RV and LV. Although it is not very likely, the regularity of association between cTn and tropomyosin in RVCM might not be as strict as in LVCM. Otherwise, the immunoblot results might include putative soluble components of the myofilament proteins that are different between RV and LV.

To precisely compare the intrinsic Ca2+-buffering capacity, we calibrated the in situ Ca2+ buffering properties using the added-buffer method (Berlin et al., 1994; Neher, 1995). For this purpose, it was necessary to define the buffer components as a collective Ca2+-binding species (S), described by a single dissociation constant (Kd,S) and concentration ([S]tot). Although S partially reflects components other than TnC such as calmodulin, ATP, and myosin, it mostly depends on the Ca2+-binding affinity and concentration of TnC under physiological conditions (Smith and Eisner, 2019). The calibration and analysis actually revealed that the Ca2+-binding ratio, Kd,S, and [S]tot were lower in RVCM than in LVCM (Figs. 7 and 8). Taken together, these data suggest that the similar amplitudes of Ca2+ transients despite their shorter APD were due to the lower expression of the Ca2+-binding TnC in RVCM. However, there might be another plausible difference in posttranslational modification of TnC affecting Ca2+-binding properties (Robertson et al., 1982; Blanchard and Solaro, 1984), which remains to be investigated in the future.

Interventricular difference in SERCA activity

In rodent cardiomyocytes, SERCA is a major functional Ca2+ buffer along with cTn (Berlin et al., 1994; Smith and Eisner, 2019). The calculation of Ca2+ buffering using the methods by Berlin et al. (1994) could not take SERCA activity into account. To overcome the limitation, we calculated this effect of lower SERCA activity using mathematical in silico models (Fig. 10). Although the lower cTn increased the height of the Ca2+ transient to a greater extent, the lower SERCA activity also had an effect on the Δ[Ca2+]i. It appeared that those two features of RVCM, the lower expression of cTn and the activity of SERCA, were acting in concert in Δ[Ca2+]i of RVCM. Notably, the reduction in sarcomere shortening due to lower cTn expression became less significant by the combined lowering of SERCA (Fig. 10 C, green line versus gray line). Considering the unavoidable lag time for the Ca2+-dependent operation of myofilament proteins, the combined contribution of SERCA as well as cTn would determine the final amplitude of contraction, that is, ΔSL.

Although Kondo et al. (2006) reported no difference in SERCA2a gene expression and its inhibitory regulator PLB, Sathish et al. (2006) showed lower levels of the protein SERCA2a and PLB proteins in RVCM than LVCM. Functional studies have suggested lower SERCA activity in RVCM (Afzal and Dhalla, 1992; Sathish et al., 2006), consistent with the present study. However, there was a quantitative difference; the previous functional studies showed four to five times slower maximum velocity of SR Ca2+ uptake (Vmax) in RVCM (Afzal and Dhalla, 1992; Sathish et al., 2006). Our analysis of SERCA from the analysis of rate constants showed a 2.23-fold larger rate constant in RVCM than LVCM (Fig. 4). Even though the measured variables were not the same (Vmax versus τ), the difference was still significant. A plausible explanation is the discrepancy in measuring processes: SR-enriched homogenate (Sathish et al., 2006) versus in situ Ca2+ imaging (present study). Regardless of the quantitative difference, a lower SERCA activity and lower phosphorylation of PLB might be the key features of E–C coupling in RVCM.

Contractility of RVCM

Here, we found another notable point in the E–C coupling process: shorter ΔSL despite the similar Δ[Ca2+]i in RVCM, which might be explained by the lower expression of cTn. Under the physiological environment, the RV is exposed to smaller afterload with more rapid pressure decline in cardiac cycle, and the P-V loop shows shorter isovolumic contraction time with slower decay of systolic pressure (Dell’Italia and Walsh, 1988). The macroscopic properties of RV such as the thinner wall and nonoblique geometry of myofibers are thought to be associated with the hemodynamic differences (Redington et al., 1988; Sengupta et al., 2006; Walker and Buttrick, 2013). Although it is difficult to extrapolate the properties of a single myocyte to the whole ventricle, the smaller ΔSL and slower relaxation of RVCM might partly explain the triangular shape of the P-V loop of RV.

Although not directly comparable with our present results, a previous study using a point mutation in TnI (K118C) with dominant-negative effects showed decreased contractility and shorter length of cardiomyocytes in transgenic mice (Wei and Jin, 2014). Another case was missense mutations in the Ca2+-specific binding site of TnC (E59D and D75Y), which was discovered from a patient diagnosed with idiopathic dilated cardiomyopathy. Rat cardiomyocytes with adenoviral expression of D75Y showed less cell shortening than wild-type cardiomyocytes, even without changes in the Ca2+ transient (Lim et al., 2008). In addition, the increased proteolytic activity and subsequent lowering of cTn were responsible for the impaired contractility and relaxation in ventricular hypertrophy (van der Laarse, 2002). These studies suggest that the expression, Ca2+-binding affinity, and Ca2+ sensitivity of cTn closely regulate cell shortening. Although we did not directly investigate the Ca2+-dependent contraction of permeabilized cardiomyocytes to analyze Ca2+ sensitivity, we cautiously interpret that the smaller ΔSL might be due to, at least partly, the lower amount of cTn in RVCM. To confirm this hypothesis, it is worth investigating the relationship between the role of lowered cTn and smaller ΔSL of RVCM.

In the computational simulations combining the previous mathematical models and the kinetics of Ca2+ binding to the cTn (Pandit et al., 2001; Negroni and Lascano, 2008), we calculated the Ca2+ flux to the myofilament fraction (JTRPN) and its time-dependent change (dJTRPN/dt; Fig. 10 B). This approach was based on the assumption that the Ca2+ flux to the myofilament is totally engaged in the Tn-Ca2+ state cycle. In the Negroni and Lascano (2008) model, the force and length of contraction depend on the number of cross-bridges, that is, the amount of effective troponin. In this context, the cTn-dependent ΔSL in the virtual simulation would be a self-evident result. Nevertheless, the intriguing point of our present study is that the expression of cTn was actually lower in RVCM (Figs. 6 and S7).

Although the heart rate of rats is faster than that of humans, 5–6 Hz (300–350 bpm), we used a 2-Hz frequency for pacing cardiomyocytes due to constraints in the experimental system using isolated single cells. For instance, myocytes showed tendency of deterioration at high rates of stimulating frequencies, especially when loaded with fura-2, limiting the simultaneous rerecording of Ca2+ transient and ΔSL. In the IonOptix system solely recording SL, we could confirm that the differences in contractility (ΔSL and relaxation speed) between RVCM and LVCM were maintained at higher pacing frequencies (Fig. S9).

It was notable that both RVCM and LVCM showed a tendency toward decreasing ΔSL at the higher pacing rates (Fig. S9, B and C), suggesting a kind of negative force–frequency relation (FFR). In large animals, the ventricular myocytes exhibit positive FFR due to increased SR Ca2+ content (Kurihara and Sakai, 1985; Joulin et al., 2009). In contrast, rat cardiomyocytes exhibit very weakly positive or even negative FFR (Banijamali et al., 1991; Maier et al., 1998; Janssen and Periasamy, 2007) owing to their higher SR Ca2+ content, even at the low stimulation frequencies (Shattock and Bers, 1989; Maier et al., 2000).

While both RVCM and LVCM showed frequency-dependent acceleration of relaxation (FDAR), slower relaxation of RVCM was consistently observed at the physiological pacing frequency (Fig. S9, E and F). In physiological conditions, FDAR might be essential to refill more rapidly between beats (Schouten, 1990; Hussain et al., 1997). It was reported that the FDAR was SR dependent and modulated by SERCA (Bassani et al., 1995). Although calcium transient was not directly compared at the higher pacing frequencies, accelerated Ca2+ removal could be inferred from the recordings of ΔSL. Also, we cautiously concluded that slower recovery of Ca2+ transient of RVCM would be conserved at the physiological heart rate of rodents.

Pathophysiological implications

Along with the anatomic biventricular differences such as the ventricular wall thickness and the muscle fiber band folding pattern, the lower cardiac troponin and the smaller ΔSL in the RVCM might be related to the pathophysiological responses of the RV in vivo. Changes in Ca2+ sensitivity make a significant contribution to the contractility of cardiomyocytes during disease progression (Belin et al., 2007; Marston and de Tombe, 2008). Because the phosphorylation of TnI plays a critical role in the modulation of Ca2+ sensitivity, the lower cTn might be partly responsible for the less effective adaptation to the increased afterload.

The importance of Ca2+ buffering has been suggested in genetically engineered rats to express parvalbumin, a Ca2+-binding protein, in cardiomyocytes (Wahr et al., 1999). Although there was little effect on the peak Ca2+ transient due to the slow Ca2+ binding kinetics of parvalbumin, the gene-transferred cardiomyocytes showed accelerated decay of [Ca2+]i, implying improved diastolic performance of cardiomyocytes. These effects of parvalbumin suggest the potential importance of slow buffers, including the Ca2+/Mg2+-binding site on TnC, in the development and consequences of heart failure, as recently mentioned by Smith and Eisner (2019). While we did not directly measure the slow-acting buffer power of cardiomyocytes, the lower expression of cTn in RVCM might be partly associated with the vulnerability of the RV.

Beat-to-beat alternation of cardiac AP (alternans) is regarded as an arrhythmia mechanism with an increased risk of sudden cardiac death (Kulkarni et al., 2019). The delayed Ca2+ decay or abnormal Ca2+ signal could underlie the alternans, as suggested by the reduced probability of alternans by injecting EGTA into cardiomyocytes (Baudenbacher et al., 2008). The increased Ca2+ buffer could suppress the pathological Ca2+ waves triggered by SR Ca2+ release (Nivala et al., 2012; Bovo et al., 2015). In addition to the buffering effects on Ca2+ sparks, there was an idea that incomplete recovery at the time of the next stimulus followed by the slow decay of the Ca2+ transient might be attributed to the induction of alternans (Baudenbacher et al., 2008). In fact, genetically up-regulated SERCA2a suppressed the genesis of alternans (Lyon et al., 2011; Cutler et al., 2012). Interestingly, ventricular arrhythmia was more frequent in RV myocardial infarction than in inferior or anterior wall LV myocardial infarction (Mehta et al., 2001; Ondrus et al., 2013). According to the present study, we cautiously suppose that the lower Ca2+ buffer by TnC and the slower Ca2+ sequestration by SERCA in RVCM implied higher susceptibility to ventricular arrhythmia in the above conditions.

In summary, from the comparative analysis of the E–C coupling between the RVCM and LVCM of rat, in addition to the electrical difference (higher Ito and shorter AP), we found lower Ca2+-buffering mechanisms in RVCM, that is, TnC and SERCA (Fig. 11). The lower expression of TnC could explain the dual mismatch phenomena in the biventricular differences of the E–C coupling. The pathophysiological implications of the present findings, such as the putative changes in TnC expression, require further investigation in right heart disease models.

David A. Eisner served as editor.

This work was supported by the National Research Foundation of Korea funded by the Ministry of Science and ICT, Republic of Korea to S.J. Kim (grants NRF-2018R1D1A1B07048998, NRF-2018R1A5A2025964, and NRF-2021R1A2C2007243), and by EDISON (Education-Research Integration through Simulation on the Net; NRF-2016M3C1A6936605). This work was also supported by a grant of the M.D., Ph.D./Medical Scientist Training Program through the Korea Health Industry Development Institute to Y.K. Jeon, funded by the Ministry of Health and Welfare, Republic of Korea.

The authors declare no competing financial interests.

Author contributions: conception and design: Y.K. Jeon, J. Jang, J.B. Youm, Y.H. Zhang, and S.J. Kim; collection of data: Y.K. Jeon, J.W. Kwon, J. Jang, S.W. Choi, J. Woo, S.H. Cho, and B.I. Yu; analysis and/or interpretation: Y.K. Jeon, J.W. Kwon, J. Jang, Y.S. Chun, J.B. Youm, Y.H. Zhang, and S.J. Kim; manuscript writing: Y.K. Jeon, J.B. Youm, Y.H. Zhang, and S.J. Kim. All the authors reviewed and approved the final version of the manuscript.

Afzal
,
N.
, and
N.S.
Dhalla
.
1992
.
Differential changes in left and right ventricular SR calcium transport in congestive heart failure
.
Am. J. Physiol.
262
:
H868
H874
.
Banijamali
,
H.S.
,
W.D.
Gao
,
B.R.
MacIntosh
, and
H.E.
ter Keurs
.
1991
.
Force-interval relations of twitches and cold contractures in rat cardiac trabeculae. Effect of ryanodine
.
Circ. Res.
69
:
937
948
.
Bassani
,
J.W.
,
R.A.
Bassani
, and
D.M.
Bers
.
1994
.
Relaxation in rabbit and rat cardiac cells: species-dependent differences in cellular mechanisms
.
J. Physiol.
476
:
279
293
.
Bassani
,
R.A.
,
A.
Mattiazzi
, and
D.M.
Bers
.
1995
.
CaMKII is responsible for activity-dependent acceleration of relaxation in rat ventricular myocytes
.
Am. J. Physiol.
268
:
H703
H712
.
Baudenbacher
,
F.
,
T.
Schober
,
J.R.
Pinto
,
V.Y.
Sidorov
,
F.
Hilliard
,
R.J.
Solaro
,
J.D.
Potter
, and
B.C.
Knollmann
.
2008
.
Myofilament Ca2+ sensitization causes susceptibility to cardiac arrhythmia in mice
.
J. Clin. Invest.
118
:
3893
3903
.
Belin
,
R.J.
,
M.P.
Sumandea
,
E.J.
Allen
,
K.
Schoenfelt
,
H.
Wang
,
R.J.
Solaro
, and
P.P.
de Tombe
.
2007
.
Augmented protein kinase C-alpha-induced myofilament protein phosphorylation contributes to myofilament dysfunction in experimental congestive heart failure
.
Circ. Res.
101
:
195
204
.
Berlin
,
J.R.
,
J.W.
Bassani
, and
D.M.
Bers
.
1994
.
Intrinsic cytosolic calcium buffering properties of single rat cardiac myocytes
.
Biophys. J.
67
:
1775
1787
.
Blanchard
,
E.M.
, and
R.J.
Solaro
.
1984
.
Inhibition of the activation and troponin calcium binding of dog cardiac myofibrils by acidic pH
.
Circ. Res.
55
:
382
391
.
Borzak
,
S.
,
R.A.
Kelly
,
B.K.
Krämer
,
Y.
Matoba
,
J.D.
Marsh
, and
M.
Reers
.
1990
.
In situ calibration of fura-2 and BCECF fluorescence in adult rat ventricular myocytes
.
Am. J. Physiol.
259
:
H973
H981
.
Bovo
,
E.
,
S.R.
Mazurek
,
M.
Fill
, and
A.V.
Zima
.
2015
.
Cytosolic Ca2+ buffering determines the intra-SR Ca2+ concentration at which cardiac Ca2+ sparks terminate
.
Cell Calcium.
58
:
246
253
.
Brunet
,
S.
,
F.
Aimond
,
H.
Li
,
W.
Guo
,
J.
Eldstrom
,
D.
Fedida
,
K.A.
Yamada
, and
J.M.
Nerbonne
.
2004
.
Heterogeneous expression of repolarizing, voltage-gated K+ currents in adult mouse ventricles
.
J. Physiol.
559
:
103
120
.
Carneiro-Júnior
,
M.A.
,
T.N.
Prímola-Gomes
,
J.F.
Quintão-Júnior
,
L.R.
Drummond
,
V.N.
Lavorato
,
F.R.
Drummond
,
L.B.
Felix
,
E.M.
Oliveira
,
J.S.
Cruz
,
A.J.
Natali
, and
J.G.
Mill
.
2013
.
Regional effects of low-intensity endurance training on structural and mechanical properties of rat ventricular myocytes
.
J Appl Physiol (1985).
115
:
107
115
.
Cheung
,
J.Y.
,
D.L.
Tillotson
,
R.V.
Yelamarty
, and
R.C.
Scaduto
Jr
.
1989
.
Cytosolic free calcium concentration in individual cardiac myocytes in primary culture
.
Am. J. Physiol.
256
:
C1120
C1130
.
Cooper
,
P.J.
,
C.
Soeller
, and
M.B.
Cannell
.
2010
.
Excitation-contraction coupling in human heart failure examined by action potential clamp in rat cardiac myocytes
.
J. Mol. Cell. Cardiol.
49
:
911
917
.
Cutler
,
M.J.
,
X.
Wan
,
B.N.
Plummer
,
H.
Liu
,
I.
Deschenes
,
K.R.
Laurita
,
R.J.
Hajjar
, and
D.S.
Rosenbaum
.
2012
.
Targeted sarcoplasmic reticulum Ca2+ ATPase 2a gene delivery to restore electrical stability in the failing heart
.
Circulation.
126
:
2095
2104
.
Dell’Italia
,
L.J.
, and
R.A.
Walsh
.
1988
.
Acute determinants of the hangout interval in the pulmonary circulation
.
Am. Heart J.
116
:
1289
1297
.
Di Diego
,
J.M.
,
Z.Q.
Sun
, and
C.
Antzelevitch
.
1996
.
I(to) and action potential notch are smaller in left vs. right canine ventricular epicardium
.
Am. J. Physiol.
271
:
H548
H561
.
Díaz
,
M.E.
,
H.K.
Graham
, and
A.W.
Trafford
.
2004
.
Enhanced sarcolemmal Ca2+ efflux reduces sarcoplasmic reticulum Ca2+ content and systolic Ca2+ in cardiac hypertrophy
.
Cardiovasc. Res.
62
:
538
547
.
Duggal
,
D.
,
S.
Requena
,
J.
Nagwekar
,
S.
Raut
,
R.
Rich
,
H.
Das
,
V.
Patel
,
I.
Gryczynski
,
R.
Fudala
,
Z.
Gryczynski
, et al
.
2017
.
No Difference in Myosin Kinetics and Spatial Distribution of the Lever Arm in the Left and Right Ventricles of Human Hearts
.
Front. Physiol.
8
:
732
.
Eisner
,
D.A.
2021
.
Pseudoreplication in physiology: More means less
.
J. Gen. Physiol.
153
:e202012826.
Eisner
,
D.A.
,
J.L.
Caldwell
,
K.
Kistamás
, and
A.W.
Trafford
.
2017
.
Calcium and Excitation-Contraction Coupling in the Heart
.
Circ. Res.
121
:
181
195
.
Fabiato
,
A.
1983
.
Calcium-induced release of calcium from the cardiac sarcoplasmic reticulum
.
Am. J. Physiol.
245
:
C1
C14
.
Fabiato
,
A.
, and
F.
Fabiato
.
1979
.
Calculator programs for computing the composition of the solutions containing multiple metals and ligands used for experiments in skinned muscle cells
.
J. Physiol. (Paris).
75
:
463
505
.
Foschi
,
M.
,
M.
Di Mauro
,
F.
Tancredi
,
C.
Capparuccia
,
R.
Petroni
,
L.
Leonzio
,
S.
Romano
,
S.
Gallina
,
M.
Penco
,
M.
Cibelli
, and
A.
Calafiore
.
2017
.
The Dark Side of the Moon: The Right Ventricle
.
J. Cardiovasc. Dev. Dis.
4
:
18
.
Friedberg
,
M.K.
, and
A.N.
Redington
.
2014
.
Right versus left ventricular failure: differences, similarities, and interactions
.
Circulation.
129
:
1033
1044
.
Ghosh
,
R.
,
J.E.
Gilda
, and
A.V.
Gomes
.
2014
.
The necessity of and strategies for improving confidence in the accuracy of western blots
.
Expert Rev. Proteomics.
11
:
549
560
.
Gilda
,
J.E.
, and
A.V.
Gomes
.
2013
.
Stain-Free total protein staining is a superior loading control to β-actin for Western blots
.
Anal. Biochem.
440
:
186
188
.
Harris
,
D.M.
,
G.D.
Mills
,
X.
Chen
,
H.
Kubo
,
R.M.
Berretta
,
V.S.
Votaw
,
L.F.
Santana
, and
S.R.
Houser
.
2005
.
Alterations in early action potential repolarization causes localized failure of sarcoplasmic reticulum Ca2+ release
.
Circ. Res.
96
:
543
550
.
Hove-Madsen
,
L.
, and
D.M.
Bers
.
1993
.
Passive Ca buffering and SR Ca uptake in permeabilized rabbit ventricular myocytes
.
Am. J. Physiol.
264
:
C677
C686
.
Hussain
,
M.
,
G.A.
Drago
,
J.
Colyer
, and
C.H.
Orchard
.
1997
.
Rate-dependent abbreviation of Ca2+ transient in rat heart is independent of phospholamban phosphorylation
.
Am. J. Physiol.
273
:
H695
H706
.
Jang
,
J.H.
,
M.J.
Kang
,
G.P.
Ko
,
S.J.
Kim
,
E.C.
Yi
, and
Y.H.
Zhang
.
2015
.
Identification of a novel splice variant of neuronal nitric oxide synthase, nNOSβ, in myofilament fraction of murine cardiomyocytes
.
Nitric Oxide.
50
:
20
27
.
Janssen
,
P.M.
, and
M.
Periasamy
.
2007
.
Determinants of frequency-dependent contraction and relaxation of mammalian myocardium
.
J. Mol. Cell. Cardiol.
43
:
523
531
.
Jin
,
C.Z.
,
J.H.
Jang
,
Y.
Wang
,
J.G.
Kim
,
Y.M.
Bae
,
J.
Shi
,
C.R.
Che
,
S.J.
Kim
, and
Y.H.
Zhang
.
2012
.
Neuronal nitric oxide synthase is up-regulated by angiotensin II and attenuates NADPH oxidase activity and facilitates relaxation in murine left ventricular myocytes
.
J. Mol. Cell. Cardiol.
52
:
1274
1281
.
Joulin
,
O.
,
S.
Marechaux
,
S.
Hassoun
,
D.
Montaigne
,
S.
Lancel
, and
R.
Neviere
.
2009
.
Cardiac force-frequency relationship and frequency-dependent acceleration of relaxation are impaired in LPS-treated rats
.
Crit. Care.
13
:
R14
.
Katsube
,
Y.
,
H.
Yokoshiki
,
L.
Nguyen
,
M.
Yamamoto
, and
N.
Sperelakis
.
1998
.
L-type Ca2+ currents in ventricular myocytes from neonatal and adult rats
.
Can. J. Physiol. Pharmacol.
76
:
873
881
.
Knollmann
,
B.C.
,
A.N.
Katchman
, and
M.R.
Franz
.
2001
.
Monophasic action potential recordings from intact mouse heart: validation, regional heterogeneity, and relation to refractoriness
.
J. Cardiovasc. Electrophysiol.
12
:
1286
1294
.
Kondo
,
R.P.
,
D.A.
Dederko
,
C.
Teutsch
,
J.
Chrast
,
D.
Catalucci
,
K.R.
Chien
, and
W.R.
Giles
.
2006
.
Comparison of contraction and calcium handling between right and left ventricular myocytes from adult mouse heart: a role for repolarization waveform
.
J. Physiol.
571
:
131
146
.
Konishi
,
M.
,
A.
Olson
,
S.
Hollingworth
, and
S.M.
Baylor
.
1988
.
Myoplasmic binding of fura-2 investigated by steady-state fluorescence and absorbance measurements
.
Biophys. J.
54
:
1089
1104
.
Kulkarni
,
K.
,
F.M.
Merchant
,
M.B.
Kassab
,
F.
Sana
,
K.
Moazzami
,
O.
Sayadi
,
J.P.
Singh
,
E.K.
Heist
, and
A.A.
Armoundas
.
2019
.
Cardiac Alternans: Mechanisms and Clinical Utility in Arrhythmia Prevention
.
J. Am. Heart Assoc.
8
:e013750.
Kurihara
,
S.
, and
T.
Sakai
.
1985
.
Effects of rapid cooling on mechanical and electrical responses in ventricular muscle of guinea-pig
.
J. Physiol.
361
:
361
378
.
Lim
,
C.C.
,
H.
Yang
,
M.
Yang
,
C.K.
Wang
,
J.
Shi
,
E.A.
Berg
,
D.R.
Pimentel
,
J.K.
Gwathmey
,
R.J.
Hajjar
,
M.
Helmes
, et al
.
2008
.
A novel mutant cardiac troponin C disrupts molecular motions critical for calcium binding affinity and cardiomyocyte contractility
.
Biophys. J.
94
:
3577
3589
.
Lyon
,
A.R.
,
M.L.
Bannister
,
T.
Collins
,
E.
Pearce
,
A.H.
Sepehripour
,
S.S.
Dubb
,
E.
Garcia
,
P.
O’Gara
,
L.
Liang
,
E.
Kohlbrenner
, et al
.
2011
.
SERCA2a gene transfer decreases sarcoplasmic reticulum calcium leak and reduces ventricular arrhythmias in a model of chronic heart failure
.
Circ. Arrhythm. Electrophysiol.
4
:
362
372
.
Maier
,
L.S.
,
R.
Brandes
,
B.
Pieske
, and
D.M.
Bers
.
1998
.
Effects of left ventricular hypertrophy on force and Ca2+ handling in isolated rat myocardium
.
Am. J. Physiol.
274
:
H1361
H1370
.
Maier
,
L.S.
,
D.M.
Bers
, and
B.
Pieske
.
2000
.
Differences in Ca(2+)-handling and sarcoplasmic reticulum Ca(2+)-content in isolated rat and rabbit myocardium
.
J. Mol. Cell. Cardiol.
32
:
2249
2258
.
Marston
,
S.B.
, and
P.P.
de Tombe
.
2008
.
Troponin phosphorylation and myofilament Ca2+-sensitivity in heart failure: increased or decreased?
J. Mol. Cell. Cardiol.
45
:
603
607
.
Mathias
,
R.T.
,
I.S.
Cohen
, and
C.
Oliva
.
1990
.
Limitations of the whole cell patch clamp technique in the control of intracellular concentrations
.
Biophys. J.
58
:
759
770
.
Mehta
,
S.R.
,
J.W.
Eikelboom
,
M.K.
Natarajan
,
R.
Diaz
,
C.
Yi
,
R.J.
Gibbons
, and
S.
Yusuf
.
2001
.
Impact of right ventricular involvement on mortality and morbidity in patients with inferior myocardial infarction
.
J. Am. Coll. Cardiol.
37
:
37
43
.
Molina
,
C.E.
,
D.M.
Johnson
,
H.
Mehel
,
R.L.
Spätjens
,
D.
Mika
,
V.
Algalarrondo
,
Z.H.
Slimane
,
P.
Lechêne
,
N.
Abi-Gerges
,
H.J.
van der Linde
, et al
.
2014
.
Interventricular differences in β-adrenergic responses in the canine heart: role of phosphodiesterases
.
J. Am. Heart Assoc.
3
:e000858.
Molina
,
C.E.
,
J.
Heijman
, and
D.
Dobrev
.
2016
.
Differences in Left Versus Right Ventricular Electrophysiological Properties in Cardiac Dysfunction and Arrhythmogenesis
.
Arrhythm. Electrophysiol. Rev.
5
:
14
19
.
Negroni
,
J.A.
, and
E.C.
Lascano
.
2008
.
Simulation of steady state and transient cardiac muscle response experiments with a Huxley-based contraction model
.
J. Mol. Cell. Cardiol.
45
:
300
312
.
Neher
,
E.
1995
.
The use of fura-2 for estimating Ca buffers and Ca fluxes
.
Neuropharmacology.
34
:
1423
1442
.
Nivala
,
M.
,
C.Y.
Ko
,
M.
Nivala
,
J.N.
Weiss
, and
Z.
Qu
.
2012
.
Criticality in intracellular calcium signaling in cardiac myocytes
.
Biophys. J.
102
:
2433
2442
.
Oda
,
T.
,
H.
Yanagisawa
, and
T.
Wakabayashi
.
2020
.
Cryo-EM structures of cardiac thin filaments reveal the 3D architecture of troponin
.
J. Struct. Biol.
209
:107450.
Ondrus
,
T.
,
J.
Kanovsky
,
T.
Novotny
,
I.
Andrsova
,
J.
Spinar
, and
P.
Kala
.
2013
.
Right ventricular myocardial infarction: From pathophysiology to prognosis
.
Exp. Clin. Cardiol.
18
:
27
30
.
Pan
,
B.S.
, and
R.J.
Solaro
.
1987
.
Calcium-binding properties of troponin C in detergent-skinned heart muscle fibers
.
J. Biol. Chem.
262
:
7839
7849
.
Pandit
,
S.V.
,
R.B.
Clark
,
W.R.
Giles
, and
S.S.
Demir
.
2001
.
A mathematical model of action potential heterogeneity in adult rat left ventricular myocytes
.
Biophys. J.
81
:
3029
3051
.
Phillips
,
D.
,
A.M.
Aponte
,
R.
Covian
,
E.
Neufeld
,
Z.X.
Yu
, and
R.S.
Balaban
.
2011
.
Homogenous protein programming in the mammalian left and right ventricle free walls
.
Physiol. Genomics.
43
:
1198
1206
.
Pillai-Kastoori
,
L.
,
A.R.
Schutz-Geschwender
, and
J.A.
Harford
.
2020
.
A systematic approach to quantitative Western blot analysis
.
Anal. Biochem.
593
:113608.
Redington
,
A.N.
,
H.H.
Gray
,
M.E.
Hodson
,
M.L.
Rigby
, and
P.J.
Oldershaw
.
1988
.
Characterisation of the normal right ventricular pressure-volume relation by biplane angiography and simultaneous micromanometer pressure measurements
.
Br. Heart J.
59
:
23
30
.
Robertson
,
S.P.
,
J.D.
Johnson
, and
J.D.
Potter
.
1981
.
The time-course of Ca2+ exchange with calmodulin, troponin, parvalbumin, and myosin in response to transient increases in Ca2+
.
Biophys. J.
34
:
559
569
.
Robertson
,
S.P.
,
J.D.
Johnson
,
M.J.
Holroyde
,
E.G.
Kranias
,
J.D.
Potter
, and
R.J.
Solaro
.
1982
.
The effect of troponin I phosphorylation on the Ca2+-binding properties of the Ca2+-regulatory site of bovine cardiac troponin
.
J. Biol. Chem.
257
:
260
263
.
Sander
,
H.
,
S.
Wallace
,
R.
Plouse
,
S.
Tiwari
, and
A.V.
Gomes
.
2019
.
Ponceau S waste: Ponceau S staining for total protein normalization
.
Anal. Biochem.
575
:
44
53
.
Sathish
,
V.
,
A.
Xu
,
M.
Karmazyn
,
S.M.
Sims
, and
N.
Narayanan
.
2006
.
Mechanistic basis of differences in Ca2+ -handling properties of sarcoplasmic reticulum in right and left ventricles of normal rat myocardium
.
Am. J. Physiol. Heart Circ. Physiol.
291
:
H88
H96
.
Satoh
,
H.
,
L.M.
Delbridge
,
L.A.
Blatter
, and
D.M.
Bers
.
1996
.
Surface:volume relationship in cardiac myocytes studied with confocal microscopy and membrane capacitance measurements: species-dependence and developmental effects
.
Biophys. J.
70
:
1494
1504
.
Schouten
,
V.J.
1990
.
Interval dependence of force and twitch duration in rat heart explained by Ca2+ pump inactivation in sarcoplasmic reticulum
.
J. Physiol.
431
:
427
444
.
Sengupta
,
P.P.
,
J.
Korinek
,
M.
Belohlavek
,
J.
Narula
,
M.A.
Vannan
,
A.
Jahangir
, and
B.K.
Khandheria
.
2006
.
Left ventricular structure and function: basic science for cardiac imaging
.
J. Am. Coll. Cardiol.
48
:
1988
2001
.
Shattock
,
M.J.
, and
D.M.
Bers
.
1989
.
Rat vs. rabbit ventricle: Ca flux and intracellular Na assessed by ion-selective microelectrodes
.
Am. J. Physiol.
256
:
C813
C822
.
Sikkel
,
M.B.
,
D.P.
Francis
,
J.
Howard
,
F.
Gordon
,
C.
Rowlands
,
N.S.
Peters
,
A.R.
Lyon
,
S.E.
Harding
, and
K.T.
MacLeod
.
2017
.
Hierarchical statistical techniques are necessary to draw reliable conclusions from analysis of isolated cardiomyocyte studies
.
Cardiovasc. Res.
113
:
1743
1752
.
Simmerman
,
H.K.
,
J.H.
Collins
,
J.L.
Theibert
,
A.D.
Wegener
, and
L.R.
Jones
.
1986
.
Sequence analysis of phospholamban. Identification of phosphorylation sites and two major structural domains
.
J. Biol. Chem.
261
:
13333
13341
.
Smith
,
G.L.
, and
D.A.
Eisner
.
2019
.
Calcium Buffering in the Heart in Health and Disease
.
Circulation.
139
:
2358
2371
.
Su
,
Y.R.
,
M.
Chiusa
,
E.
Brittain
,
A.R.
Hemnes
,
T.S.
Absi
,
C.C.
Lim
, and
T.G.
Di Salvo
.
2015
.
Right ventricular protein expression profile in end-stage heart failure
.
Pulm. Circ.
5
:
481
497
.
Sweeney
,
H.L.
, and
D.W.
Hammers
.
2018
.
Muscle Contraction
.
Cold Spring Harb. Perspect. Biol.
10
:a023200.
Tada
,
M.
, and
A.M.
Katz
.
1982
.
Phosphorylation of the sarcoplasmic reticulum and sarcolemma
.
Annu. Rev. Physiol.
44
:
401
423
.
van der Laarse
,
A.
2002
.
Hypothesis: troponin degradation is one of the factors responsible for deterioration of left ventricular function in heart failure
.
Cardiovasc. Res.
56
:
8
14
.
Wahr
,
P.A.
,
D.E.
Michele
, and
J.M.
Metzger
.
1999
.
Parvalbumin gene transfer corrects diastolic dysfunction in diseased cardiac myocytes
.
Proc. Natl. Acad. Sci. USA.
96
:
11982
11985
.
Walker
,
L.A.
, and
P.M.
Buttrick
.
2013
.
The right ventricle: biologic insights and response to disease: updated
.
Curr. Cardiol. Rev.
9
:
73
81
.
Watanabe
,
T.
,
L.M.
Delbridge
,
J.O.
Bustamante
, and
T.F.
McDonald
.
1983
.
Heterogeneity of the action potential in isolated rat ventricular myocytes and tissue
.
Circ. Res.
52
:
280
290
.
Wei
,
H.
, and
J.P.
Jin
.
2014
.
A dominantly negative mutation in cardiac troponin I at the interface with troponin T causes early remodeling in ventricular cardiomyocytes
.
Am. J. Physiol. Cell Physiol.
307
:
C338
C348
.
Yamada
,
Y.
,
K.
Namba
, and
T.
Fujii
.
2020
.
Cardiac muscle thin filament structures reveal calcium regulatory mechanism
.
Nat. Commun.
11
:
153
.

Author notes

*

Y.K. Jeon and J.W. Kwon contributed equally to this paper.

This article is distributed under the terms of an Attribution–Noncommercial–Share Alike–No Mirror Sites license for the first six months after the publication date (see http://www.rupress.org/terms/). After six months it is available under a Creative Commons License (Attribution–Noncommercial–Share Alike 4.0 International license, as described at https://creativecommons.org/licenses/by-nc-sa/4.0/).