Electronic Supplementary Material

Journal of Muscle Research and Cell Motility

L71F mutation in rat cardiac Troponin T augments crossbridge recruitment and detachment dynamics against α-myosin heavy chain, but not against β-myosin heavy chain

Sherif M. Reda, Sampath K. Gollapudi, and Murali Chandra*

Department of Integrative Physiology and Neuroscience (IPN), Washington State University, Pullman, WA, USA

*Address correspondence to: Murali Chandra, Ph.D, PO Box 647620, 251 Veterinary and Biomedical Research Building, Department of IPN, Washington State University, Pullman, WA-99164. Tel: +1 (509) 335-7561; Fax: +1 (509) 335-4650; Email:

SUPPORTING MATERIALS & METHODS

Propylthiouracil(PTU) treatment: PTU was administered in both drinking water (0.2 g L-1) and solid feed (0.15% PTU, Harlan Laboratories, Madison, WI) for approximately 5 weeks. PTU silences the α-myosin heavy chain (MHC) promoter, leading to activation of the β-MHC promoter in the cardiac ventricle (Chizzonite and Zak 1984). After 5 weeks of treatment with PTU, there was a complete switch of MHC isoform from α- to β-MHC in the ventricles (Michael and Chandra 2016; Michael et al. 2014). Previous studies have shown that this switch to β-MHC occurs without any alterations in thick or thin filament proteins, their stoichiometry, or phosphorylation status(Fitzsimons et al. 1998; Herron et al. 2001; Krenz et al. 2007; Locher et al. 2009; Metzger et al. 1999).

Detergent-skinned cardiac muscle fibers: Detergent-skinned cardiac muscle fibers were prepared as described previously (Chandra et al. 2006; Tschirgi et al. 2006). Hearts were quickly removed from deeply anesthetized (isofluorane) rats and immersed in an ice-cold high relaxing (HR) solution containing (in mM): 20 2,3-butanedione monoxime (BDM), 50 N,N-bis (2-hydroxyethyl)-2-amino-ethane-sulfonic acid (BES), 20 EGTA, 6.29 MgCl2, 6.09 Na2ATP, 30.83 potassium propionate, 10 sodium azide, 1.0 DTT, and 4 benzamidine-HCl. The pH of the HR solution was adjusted to 7.0 with KOH. HR also included protease inhibitors (in μM: 5 bestatin, 2 E-64, 10 leupeptin, 1 pepstatin, and 200 PMSF). Left ventricular papillary muscle bundles were dissected into smaller muscle fibers of ~0.15 mm in width and 2-2.5 mm in length in ice-cold HR solution. Muscle fibers were detergent-skinned overnight at 4°C in HR solution containing 1% Triton X-100.

Purification of recombinant rat cardiac Tn subunits: Previously described methods for purifying recombinant rat cardiac TnT (TnTWT and TnTL71F), rat cardiac TnI, and rat cardiac TnC were used(Chandra et al. 1999; Gollapudi and Chandra 2012; Mamidi et al. 2013a). Both TnTWT and TnTL71Fweretagged with the c-myc epitope at the N-terminus. The expression vector was pSBETa and the expression cell was BL21 DE (STAR). In brief, TnTWT and TnTL71F were purified by ion-exchange chromatography on a DEAE-fast sepharose column (GE Healthcare Biosciences, Pittsburgh, PA) (Chandra et al. 1999; Chandra et al. 2006). TnIwas purified by ion exchange chromatography on a CM sepharose column (Gollapudi and Chandra 2012; Guo et al. 1994; Mamidi et al. 2013a). TnCwas purified by ion exchange chromatography on a DEAE-fast sepharose column (Gollapudi and Chandra 2012; Mamidi et al. 2013a; Pan and Johnson 1996), followed by phenyl sepharose column. Samples from eluted fractions were run on a 12.5% SDS gel to determine their purity. Pure protein fractions were pooled and dialyzed extensively against deionized water containing 15 mM β-mercaptoethanol, lyophilized, and stored at -80°C.

Western blot: Cardiac muscle fibers were solubilized in 2.5% SDS solution (10 µl/fiber) and an equal volume of protein gel loading dye (125 mM Tris-HCl (pH 6.8), 20% glycerol, 2% SDS, 0.01% bromophenol blue, and 50 mM β-mercaptoethanol) was added. Proteins were separated on an 8% SDS gel and transferred onto a polyvinylidenedifluoride membrane for Western blot analysis using a Trans-Blot Turbo Transfer System (Bio-Rad Laboratories Inc., Hercules, CA). The incorporation of TnTWT or TnTL71F was assessed using a monoclonal anti-TnT primary antibody (M401134, Fitzgerald Industries Int, Concord, MA), followed by a HRP-labeled anti-mouse secondary antibody (RPN 2132, Amersham Biosciences, Piscataway, NJ). Densitometric analysis was performed using the Image J software (acquired from NIH at:

Steady-state isometric force and ATPase activity: We measured steady-state isometric force and ATPase activity, as described previously (Chandra et al. 2007; de Tombe and Stienen 1995; Stienen et al. 1995). T-shaped aluminum clips were used to attach the muscle fiber between a motor arm (322C, Aurora Scientific Inc., Ontario, Canada) and a force transducer (AE 801, Sensor One Technologies Corp., Sausalito, CA). The sarcomere length (SL) of the muscle fibers was set to 2.3 µm under relaxing conditions (Chandra et al. 2007; de Tombe and Stienen 1995; Stienen et al. 1995). The SL was readjusted to 2.3 µm after two cycles of maximal activation and relaxation, if necessary. Detergent-skinned muscle fibers were exposed to various Ca2+ solutions in a constantly-stirred chamber. The concentration of Ca2+ in the test solutions ranged from pCa 4.3 to 9.0 (pCa = -log of [Ca2+]free). The composition of pCa solutions was calculated using methods described previously (Fabiato and Fabiato 1979). The composition of the maximal Ca2+ activating solution (pCa 4.3) was: (in mM) 31 potassium propionate, 5.95 Na2ATP, 6.61 MgCl2, 10 EGTA, 10.11 CaCl2, 50 BES, 5 sodium azide, and 10 phosphoenol pyruvate (PEP). The relaxing solution (pCa 9.0) contained: (in mM) 51.14 potassium propionate, 5.83 Na2ATP, 6.87 MgCl2, 10 EGTA, 0.024 CaCl2, 50 BES, 5 NaN3, and 10 PEP. The activating and relaxing solutions also contained 0.5 mg/ml pyruvate kinase (500 U/mg), 0.05 mg/ml lactate dehydrogenase (870 U/mg), 20 µM diadenosine pentaphosphate, 10 µM oligomycin, and a cocktail of protease inhibitors. The pH and ionic strength of maximal activating and relaxing solutions were adjusted to 7.0 and 180 mM, respectively. All measurements were made at 20°C.

We measured steady-state isometric ATPase activity using a protocol described previously (de Tombe and Stienen 1995; Gollapudi et al. 2012; Kirk et al. 2009; Rodgers et al. 2009; Stienen et al. 1995). Briefly, near-UV light (340 nm) was projected through the muscle chamber, then split (50:50) via a beam splitter and detected at 340 nm (sensitive to changes in NADH) and at 400 nm (insensitive to changes in NADH). ATPase activity was measured as follows: ATP regeneration from ADP was coupled to the breakdown of PEP to pyruvate and ATP catalyzed by pyruvate kinase, which was linked to the synthesis of lactate catalyzed by lactate dehydrogenase. The breakdown of NADH was proportional to the ATP consumption and was measured by changes in UV absorbance at 340 nm. The signal for NADH was calibrated by multiple injections of 250 pmol of ADP.

Measurement of rate of tension redevelopment, ktr: ktr was estimated at pCa 4.3 using a modification to the large slack/restretch maneuver originally described by Brenner and Eisenberg (Brenner and Eisenberg 1986). The modification was described in our earlier works (Ford and Chandra 2013; Gollapudi et al. 2012; Michael et al. 2014). Soon after reaching the steady-state isometric force, the motor arm was commanded to rapidly slacken the muscle fiber by 10% of the ML using a high-speed length-control device (322C, Aurora Scientific Inc., Ontario, Canada). After a brief shortening period of 25 ms at the decreased ML, the motor arm was commanded to rapidly (within 0.5 ms) swing past the original set point by a 10% stretch. The 10% stretch was applied to disengage any remaining bound XBs. A large release-restretch length transientimposed on a muscle fiber and the corresponding force response are shown in Fig. S1a and S1b, respectively. ktr was determined by fitting the following mono-exponential equation to the rising phase of the force response.

Where F is the force at time t, Fss is the steady-state isometric force, and Fres is the residual force from which force starts to redevelop.

Muscle fiber mechano-dynamics: Upon attainment of steady-state force at pCa 4.3, the muscle fiber was subjected various amplitude stretche/release perturbations (in the order of ± 0.5%, ± 1.0%, ± 1.5%, and ± 2.0% of the initial muscle length (ML)) and the resulting force responses were recorded, as described previously (Ford et al. 2010). A nonlinear recruitment distortion (NLRD) model was fit to the elicited force responses to estimate the following model parameters (Ford et al. 2010): the magnitude of the instantaneous muscle fiber stiffness brought about by a sudden change in ML (ED); the rate of dissipation of strain within bound XBs(c); the parameter describing the XB strain-mediated negative impact on other force-bearing XBs (γ); the rate by which new XBs are recruited into the force-bearing state (b). Below, we describe the relationship between model parameters and physiological phenomena.

ED: Phase 1: a sudden increase in ML (Fig. S2a) leads to an instantaneous increase in force, from Fss to F1(Fig. S2b). This rise in force results from the distortion of elastic elements in the strongly-bound XBs. Therefore, an increase in F1suggests an increase in the number of strongly-bound XBs. Conversely, a decrease in F1suggests a decrease in the number of strongly-bound XBs. Because ED is estimated as the slope of the relationship (F1-Fss) and change in muscle length (L), it is an approximate measure of the number of strongly-bound XBs (Ford et al. 2010).

c: Phase 2: force decays rapidly to a minimum (nadir; Fig. S2b), as the muscle fiber is held at the increased ML (Fig. S2a). Force decays rapidly because strained XBs detach and equilibrate into a non-force bearing state. The dynamic rate at which the force decays (c) is an index of XB detachment rate, g (Campbell et al. 2004; Ford et al. 2010).

γ:When the negative effect of strained XBs on the state of other force-bearing XBs is more pronounced, the magnitude of force decline is greater; that is, nadir is more pronounced and γ is augmented. Negative effect of strained XBs on other force-bearing XBs is transmitted along the thin filament via nonlinear (cooperative) effects because XBs do not interact directly with other XBs. Therefore, γ represents the allosteric/cooperative mechanisms by which strained XBs negatively impact other force-bearing XBs.

b: Phase 3: immediately after reaching a minimum (nadir), force rises gradually with a characteristic dynamic rate, b (Fig. S2b). This rise in force is due torecruitment of additional XBs into force-bearing state in response to an increase in ML.

Data Analysis: Steady-state and dynamic contractile parameters were analyzed using a two-way analysis of variance (ANOVA); one factor in this analysis was TnT (TnTWT and TnTL71F) and the other factor was MHC (α-MHC and β-MHC). When the interaction effect was significant, it suggested that the effect of TnT on a given contractile parameter was dissimilar in fibers containing α- versus β-MHC. When the interaction effect was not significant, we interpreted the main effect of TnT. To probe the cause for a significant interaction or a main effect, post-hoc multiple comparisons were made using Fisher’s uncorrected Least Significant Differences (LSD) method. Before analyzing the data, we verified that all NLRD model-derived parameters followed a normal distribution pattern within each group and that the variance in the parameter was uniform among groups. Measurements made from several muscle fibers in each group were averaged and presented as mean ± SEM. The criterion for statistical significance was set at P<0.05.

SUPPORTING RESULTS

TnTL71F-mediated impact on ED in α- and β-MHC fibers: Muscle length-perturbation experiments were performed to determine if EDcorroborated maximal tension (Fig. 1) in α- and β-MHC fibers. We have previously shown that EDis an approximation of the number of strongly-bound XBs (Ford et al. 2010), and that ED is positively correlated to maximal tension (Chandra et al. 2015; Gollapudi et al. 2015; Michael et al. 2014). ED was estimated as the slope of the linear relationship between changes in the instantaneous forces, ∆F, and the imposed ML changes, ∆L (Fig. S4a). Two-way ANOVA did not reveal a significant interaction effect (P=0.23) on ED, although, the main effect of TnT was significant (P<0.001). Post-hoc analysis revealed that TnTL71F significantly decreased ED by 27% in α-MHC fibers (P<0.001; Fig. S4b) and by 13% in β-MHC fibers (P<0.05; Fig S4b). Thus, EDcorroborated our observations on maximal tension; this substantiated that greater attenuation of maximal tension in α-MHC+TnTL71F fibers was due to greater decrease in the number of force-bearing XBs.

TnTL71F -mediated impact on ktr in α- and β-MHC fibers: To support our observations on the NLRD model parameter b (Fig. 3a), we also analyzed ktr, which is another index of XB turnover rate. Previous studies have shown correlation between b and ktr(Campbell et al. 2004). Two-way ANOVA revealed a significant interaction effect on ktr (P<0.05), suggesting that the TnTL71F-mediated effect was differently modulated in α- and β-MHC fibers. Post-hoc tests revealed that TnTL71F increased ktrby 15% in α-MHC fibers (P<0.001; Fig. S5a) but showed no effect on ktrin β-MHC fibers (P=0.71; Fig. S5a). Thus,ktrcorroborated our findings in b.

TnTL71F-mediated impact on tension cost in α- and β-MHC fibers: To support our observations on the NLRD model parameter c (Fig. 3b),we also assessed tension cost. Tension cost is strongly correlated to c and both are valuable indices of XB detachment rate, g(Campbell et al. 2004). Tension cost was estimated as the slope of the linear relationship between tension and ATPase activity at various pCa (Fig. S6a). Two-way ANOVA revealed a significant interaction effect on tension cost (P<0.05), suggesting that TnTL71F-mediated effect on tension cost was differently modulated by α- and β-MHC. Post-hoc analysis showed that TnTL71F increased tension cost by 17% in α-MHC fibers (P<0.01; Fig. S6b), but revealed no effect in β-MHC fibers (P=0.81; Fig. S6b). Our observations on tension cost corroboratec, suggesting that TnTL71F augments g in α-MHC fibers but not in β-MHC fibers.

REFERENCES

Brenner B, Eisenberg E (1986) Rate of force generation in muscle: correlation with actomyosin ATPase activity in solution Proceedings of the National Academy of Sciences of the United States of America 83:3542-3546

Campbell KB, Chandra M, Kirkpatrick RD, Slinker BK, Hunter WC (2004) Interpreting cardiac muscle force-length dynamics using a novel functional model Am J Physiol Heart Circ Physiol 286:H1535-1545 doi:10.1152/ajpheart.01029.2003

Chandra M, Montgomery DE, Kim JJ, Solaro RJ (1999) The N-terminal region of troponin T is essential for the maximal activation of rat cardiac myofilaments Journal of molecular and cellular cardiology 31:867-880 doi:10.1006/jmcc.1999.0928

Chandra M, Tschirgi ML, Ford SJ, Slinker BK, Campbell KB (2007) Interaction between myosin heavy chain and troponin isoforms modulate cardiac myofiber contractile dynamics American journal of physiology Regulatory, integrative and comparative physiology 293:R1595-1607 doi:10.1152/ajpregu.00157.2007

Chandra M, Tschirgi ML, Rajapakse I, Campbell KB (2006) Troponin T modulates sarcomere length-dependent recruitment of cross-bridges in cardiac muscle Biophysical journal 90:2867-2876 doi:10.1529/biophysj.105.076950

Chandra V, Gollapudi SK, Chandra M (2015) Rat cardiac troponin T mutation (F72L)-mediated impact on thin filament cooperativity is divergently modulated by alpha- and beta-myosin heavy chain isoforms Am J Physiol Heart Circ Physiol 309:H1260-1270 doi:10.1152/ajpheart.00519.2015

Chizzonite RA, Zak R (1984) Regulation of myosin isoenzyme composition in fetal and neonatal rat ventricle by endogenous thyroid hormones The Journal of biological chemistry 259:12628-12632

de Tombe PP, Stienen GJ (1995) Protein kinase A does not alter economy of force maintenance in skinned rat cardiac trabeculae Circulation research 76:734-741 doi:

Fitzsimons DP, Patel JR, Moss RL (1998) Role of myosin heavy chain composition in kinetics of force development and relaxation in rat myocardium J Physiol 513 ( Pt 1):171-183 doi:10.1111/j.1469-7793.1998.171by.x

Ford SJ, Chandra M (2013) Length-dependent effects on cardiac contractile dynamics are different in cardiac muscle containing alpha- or beta-myosin heavy chain Archives of biochemistry and biophysics 535:3-13 doi:10.1016/j.abb.2012.10.011

Ford SJ, Chandra M, Mamidi R, Dong W, Campbell KB (2010) Model representation of the nonlinear step response in cardiac muscle The Journal of general physiology 136:159-177 doi:10.1085/jgp.201010467

Gollapudi SK, Chandra M (2012) Cardiomyopathy-Related Mutations in Cardiac Troponin C, L29Q and G159D, Have Divergent Effects on Rat Cardiac Myofiber Contractile Dynamics Biochemistry research international 2012:824068 doi:10.1155/2012/824068

Gollapudi SK, Mamidi R, Mallampalli SL, Chandra M (2012) The N-terminal extension of cardiac troponin T stabilizes the blocked state of cardiac thin filament Biophysical journal 103:940-948 doi:10.1016/j.bpj.2012.07.035

Gollapudi SK, Tardiff JC, Chandra M (2015) The functional effect of dilated cardiomyopathy mutation (R144W) in mouse cardiac troponin T is differently affected by alpha- and beta-myosin heavy chain isoforms Am J Physiol Heart Circ Physiol 308:H884-893 doi:10.1152/ajpheart.00528.2014

Guo X, Wattanapermpool J, Palmiter KA, Murphy AM, Solaro RJ (1994) Mutagenesis of cardiac troponin I. Role of the unique NH2-terminal peptide in myofilament activation The Journal of biological chemistry 269:15210-15216 doi:None

Herron TJ, Korte FS, McDonald KS (2001) Loaded shortening and power output in cardiac myocytes are dependent on myosin heavy chain isoform expression Am J Physiol Heart Circ Physiol 281:H1217-1222 doi:None

Kirk JA et al. (2009) Left ventricular and myocardial function in mice expressing constitutively pseudophosphorylated cardiac troponin I Circulation research 105:1232-1239 doi:10.1161/CIRCRESAHA.109.205427

Krenz M, Sadayappan S, Osinska HE, Henry JA, Beck S, Warshaw DM, Robbins J (2007) Distribution and structure-function relationship of myosin heavy chain isoforms in the adult mouse heart The Journal of biological chemistry 282:24057-24064 doi:10.1074/jbc.M704574200

Locher MR, Razumova MV, Stelzer JE, Norman HS, Patel JR, Moss RL (2009) Determination of rate constants for turnover of myosin isoforms in rat myocardium: implications for in vivo contractile kinetics Am J Physiol Heart Circ Physiol 297:H247-256 doi:10.1152/ajpheart.00922.2008

Mamidi R, Mallampalli SL, Wieczorek DF, Chandra M (2013a) Identification of two new regions in the N-terminus of cardiac troponin T that have divergent effects on cardiac contractile function J Physiol 591:1217-1234 doi:10.1113/jphysiol.2012.243394

Mamidi R, Muthuchamy M, Chandra M (2013b) Instability in the central region of tropomyosin modulates the function of its overlapping ends Biophysical journal 105:2104-2113 doi:10.1016/j.bpj.2013.09.026

Metzger JM, Wahr PA, Michele DE, Albayya F, Westfall MV (1999) Effects of myosin heavy chain isoform switching on Ca2+-activated tension development in single adult cardiac myocytes Circulation research 84:1310-1317 doi:

Michael JJ, Chandra M (2016) Interplay Between the Effects of Dilated Cardiomyopathy Mutation (R206L) and the Protein Kinase C Phosphomimic (T204E) of Rat Cardiac Troponin T Are Differently Modulated by alpha- and beta-Myosin Heavy Chain Isoforms Journal of the American Heart Association 5 doi:10.1161/JAHA.115.002777

Michael JJ, Gollapudi SK, Chandra M (2014) Effects of pseudo-phosphorylated rat cardiac troponin T are differently modulated by alpha- and beta-myosin heavy chain isoforms Basic research in cardiology 109:442 doi:10.1007/s00395-014-0442-9

Pan BS, Johnson RG, Jr. (1996) Interaction of cardiotonic thiadiazinone derivatives with cardiac troponin C The Journal of biological chemistry 271:817-823 doi:doi: 10.1074/jbc.271.2.817

Rodgers BD et al. (2009) Myostatin represses physiological hypertrophy of the heart and excitation-contraction coupling J Physiol 587:4873-4886 doi:10.1113/jphysiol.2009.172544

Stienen GJ, Zaremba R, Elzinga G (1995) ATP utilization for calcium uptake and force production in skinned muscle fibres of Xenopus laevis J Physiol 482 ( Pt 1):109-122 doi:10.1113/jphysiol.1995.sp020503

Tschirgi ML, Rajapakse I, Chandra M (2006) Functional consequence of mutation in rat cardiac troponin T is affected differently by myosin heavy chain isoforms J Physiol 574:263-273 doi:10.1113/jphysiol.2006.107417