+ All Categories
Home > Documents > Final Circ Res Publication

Final Circ Res Publication

Date post: 16-Feb-2017
Category:
Upload: jerry-curran
View: 133 times
Download: 0 times
Share this document with a friend
30
Boyden, Sandor Gyorke, Hamid Band, Thomas J. Hund and Peter J. Mohler Polina, Joseph S. Alecusan, Patrick Wright, Jingdong Li, George E. Billman, Penelope A. Jerry Curran, Michael A. Makara, Sean C. Little, Hassan Musa, Bin Liu, Xiangqiong Wu, Iuliia Physiology EHD3-Dependent Endosome Pathway Regulates Cardiac Membrane Excitability and Print ISSN: 0009-7330. Online ISSN: 1524-4571 Copyright © 2014 American Heart Association, Inc. All rights reserved. is published by the American Heart Association, 7272 Greenville Avenue, Dallas, TX 75231 Circulation Research doi: 10.1161/CIRCRESAHA.115.304149 2014;115:68-78; originally published online April 23, 2014; Circ Res. http://circres.ahajournals.org/content/115/1/68 World Wide Web at: The online version of this article, along with updated information and services, is located on the http://circres.ahajournals.org/content/suppl/2014/04/23/CIRCRESAHA.115.304149.DC1.html Data Supplement (unedited) at: http://circres.ahajournals.org//subscriptions/ is online at: Circulation Research Information about subscribing to Subscriptions: http://www.lww.com/reprints Information about reprints can be found online at: Reprints: document. Permissions and Rights Question and Answer about this process is available in the located, click Request Permissions in the middle column of the Web page under Services. Further information Editorial Office. Once the online version of the published article for which permission is being requested is can be obtained via RightsLink, a service of the Copyright Clearance Center, not the Circulation Research in Requests for permissions to reproduce figures, tables, or portions of articles originally published Permissions: at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from at Ohio State University--Columbus on September 12, 2014 http://circres.ahajournals.org/ Downloaded from
Transcript
Page 1: Final Circ Res Publication

Boyden, Sandor Gyorke, Hamid Band, Thomas J. Hund and Peter J. MohlerPolina, Joseph S. Alecusan, Patrick Wright, Jingdong Li, George E. Billman, Penelope A.

Jerry Curran, Michael A. Makara, Sean C. Little, Hassan Musa, Bin Liu, Xiangqiong Wu, IuliiaPhysiology

EHD3-Dependent Endosome Pathway Regulates Cardiac Membrane Excitability and

Print ISSN: 0009-7330. Online ISSN: 1524-4571 Copyright © 2014 American Heart Association, Inc. All rights reserved.is published by the American Heart Association, 7272 Greenville Avenue, Dallas, TX 75231Circulation Research

doi: 10.1161/CIRCRESAHA.115.3041492014;115:68-78; originally published online April 23, 2014;Circ Res.

http://circres.ahajournals.org/content/115/1/68World Wide Web at:

The online version of this article, along with updated information and services, is located on the

http://circres.ahajournals.org/content/suppl/2014/04/23/CIRCRESAHA.115.304149.DC1.htmlData Supplement (unedited) at:

http://circres.ahajournals.org//subscriptions/

is online at: Circulation Research Information about subscribing to Subscriptions:

http://www.lww.com/reprints Information about reprints can be found online at: Reprints:

document. Permissions and Rights Question and Answer about this process is available in the

located, click Request Permissions in the middle column of the Web page under Services. Further informationEditorial Office. Once the online version of the published article for which permission is being requested is

can be obtained via RightsLink, a service of the Copyright Clearance Center, not theCirculation Researchin Requests for permissions to reproduce figures, tables, or portions of articles originally publishedPermissions:

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 2: Final Circ Res Publication

68

Highly evolved and differentiated for excitation–contraction coupling, cardiac myocytes express a specific profile of ion

channels, pumps, and transporters that maintain cardiomyocyte electric excitability. Collectively, these membrane proteins me-diate action potential (AP) formation and response, Ca-induced Ca release, and the secretion of natriuretic peptides.1,2 Equally

important is the set of hormone receptors localized to the sarco-lemmal membrane that regulate the activity and response of ion channels and pumps through specific second messenger path-ways. These highly evolved systems are tightly synchronized to tune cardiac output to meet the changing demands placed on the heart by variable stresses. Like other complex cells, cardiac

Integrative Physiology

© 2014 American Heart Association, Inc.

Circulation Research is available at http://circres.ahajournals.org DOI: 10.1161/CIRCRESAHA.115.304149

Rationale: Cardiac function is dependent on the coordinate activities of membrane ion channels, transporters, pumps, and hormone receptors to tune the membrane electrochemical gradient dynamically in response to acute and chronic stress. Although our knowledge of membrane proteins has rapidly advanced during the past decade, our understanding of the subcellular pathways governing the trafficking and localization of integral membrane proteins is limited and essentially unstudied in vivo. In the heart, to our knowledge, there are no in vivo mechanistic studies that directly link endosome-based machinery with cardiac physiology.

Objective: To define the in vivo roles of endosome-based cellular machinery for cardiac membrane protein trafficking, myocyte excitability, and cardiac physiology.

Methods and Results: We identify the endosome-based Eps15 homology domain 3 (EHD3) pathway as essential for cardiac physiology. EHD3-deficient hearts display structural and functional defects including bradycardia and rate variability, conduction block, and blunted response to adrenergic stimulation. Mechanistically, EHD3 is critical for membrane protein trafficking, because EHD3-deficient myocytes display reduced expression/localization of Na/Ca exchanger and L-type Ca channel type 1.2 with a parallel reduction in Na/Ca exchanger–mediated membrane current and Cav1.2-mediated membrane current. Functionally, EHD3-deficient myocytes show increased sarcoplasmic reticulum [Ca], increased spark frequency, and reduced expression/localization of ankyrin-B, a binding partner for EHD3 and Na/Ca exchanger. Finally, we show that in vivo EHD3-deficient defects are attributable to cardiac-specific roles of EHD3 because mice with cardiac-selective EHD3 deficiency demonstrate both structural and electric phenotypes.

Conclusions: These data provide new insight into the critical role of endosome-based pathways in membrane protein targeting and cardiac physiology. EHD3 is a critical component of protein trafficking in heart and is essential for the proper membrane targeting of select cellular proteins that maintain excitability. (Circ Res. 2014;115:68-78.)

Key Words: ankyrins cell biology Ehd3 protein electrophysiology ion channels mice protein transport

Original received April 7, 2014; revision received April 20, 2014; accepted April 23, 2014. In March 2014, the average time from submission to first decision for all original research papers submitted to Circulation Research was 12.63 days.

From The Dorothy M. Davis Heart & Lung Research Institute, The Ohio State University Wexner Medical Center, Columbus (J.C., M.A.M., S.C.L., H.M., B.L., X.W., I.P., J.S.A., P.W., J.L., G.E.B., S.G., T.J.H., P.J.M.); Departments of Internal Medicine (P.J.M.) and Physiology and Cell Biology (J.C., M.A.M., S.C.L., H.M., B.L., P.W., G.E.B., S.G., P.J.M.), The Ohio State University, Columbus; Department of Biomedical Engineering, The Ohio State University College of Engineering, Columbus (T.J.H.); Department of Pharmacology and Center for Molecular Therapeutics, Columbia University Medical Center, New York, NY (P.A.B.); and The Eppley Institute and UNMC-Eppley Cancer Center, University of Nebraska Medical Center, Omaha (H.B.).

The online-only Data Supplement is available with this article at http://circres.ahajournals.org/lookup/suppl/doi:10.1161/CIRCRESAHA. 115.304149/-/DC1.

Correspondence to Peter J. Mohler, PhD, The Dorothy M. Davis Heart & Lung Research Institute, The Ohio State University Wexner Medical Center, 473 W 12th Ave, Columbus, OH 43230. E-mail [email protected]; or Jerry Curran, PhD, The Dorothy M. Davis Heart & Lung Research Institute, The Ohio State University Wexner Medical Center, 473 W 12th Ave, Columbus, OH 43230. E-mail [email protected]

EHD3-Dependent Endosome Pathway Regulates Cardiac Membrane Excitability and Physiology

Jerry Curran, Michael A. Makara, Sean C. Little, Hassan Musa, Bin Liu, Xiangqiong Wu, Iuliia Polina, Joseph S. Alecusan, Patrick Wright, Jingdong Li, George E. Billman,

Penelope A. Boyden, Sandor Gyorke, Hamid Band, Thomas J. Hund, Peter J. Mohler

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 3: Final Circ Res Publication

Curran et al EHD3-Dependent Membrane Protein Trafficking 69

membrane protein residency is in constant flux and primarily regulated by 3 general processes including synthesis and traf-ficking of new proteins through the endoplasmic reticulum and Golgi to specific membrane domains, membrane protein inter-nalization and recycling, and ultimate membrane protein degra-dation. In metazoan cells, endosome-based protein machinery is indispensable for each of these functions. However, we know little to nothing about the in vivo components of the cardiac endosome system. The Eps15 homology domain (EHD)–containing protein family (EHD1–4) mediates endosome-based trafficking in nonexcitable and heterologous cells.3–5 Recently, we identified EHD3 in human heart.6 Furthermore, we identified that EHD3 levels are elevated in multiple forms of cardiovascu-lar disease.7 Based on our initial findings, we hypothesized that the endosome-based EHD3 protein plays critical roles in car-diac membrane protein trafficking and physiology at baseline and in disease and that EHD3 deficiency would result in defects in both cardiac electric and functional phenotypes.

Here, we define the in vivo physiological and mechanis-tic roles of EHD3 in the heart. EHD3-deficient (EHD3−/−) mice display enlarged hearts and abnormal cardiac function. Furthermore, EHD3−/− mice exhibit bradycardia, atrioventricu-lar conduction block, and heart rate (HR) variability and have a blunted response to β-adrenergic receptor (β-AR) stimula-tion. Adult ventricular myocytes isolated from EHD3−/− mouse hearts display a significant attenuation of AP duration (APD), increased total sarcoplasmic reticulum Ca concentration ([Ca]

SRT) and Ca sparks, a blunted β-AR response, and re-

duced expression and function of the L-type Ca channel type 1.2 (Ca

v1.2) and Na/Ca exchanger (NCX). Confocal studies

revealed improper localization of both the Cav1.2 and NCX,

consistent with the role of EHD3 for targeting select mem-brane proteins. Moreover, ankyrin-B, an intermediate binding partner between EHD3 and the NCX, is decreased in EHD3−/− mice, providing the underlying mechanism between EHD3 and membrane proteins. Finally, EHD3−/− defects are attributable to cardiac-intrinsic roles of EHD3 because mice with cardiac-spe-cific EHD3 deficiency demonstrate structural and electric phe-notypes. These new data define a critical role for EHD3 in select protein trafficking in the heart as well as indicate the impor-tance for subcellular protein targeting for cardiac excitability.

MethodsFor complete, expanded methods please refer to the Online Data Supplement. All animal studies were performed in accordance with the American Physiological Society Guiding Principles for Research Involving Animals and Human Beings and approved by The Ohio State University Institutional Animal Care and Use Committee. The investigation conforms to the Guide for the Care and Use of Laboratory Animals published by the US National Institutes of Health (NIH Publication No. 85-23, revised 1996).

EchocardiographyTransthoracic echocardiogram was performed on anesthetized wild-type (WT) and EHD3−/− mice as previously described7 to measure the in vivo function of the heart.

ElectrophysiologyWhole-cell recordings were obtained at room temperature with the use of standard patch clamp techniques, and cells from the WT and EHD3−/− groups had a similar membrane capacitance (Online Figure I).8

ECG ExperimentsECG recordings of ambulatory mice were obtained using subcuta-neously implanted radiotelemeters (DSI, St. Paul, MN).9 Recordings were obtained from mice that were conscious after exercise, after isoprenaline injection, and after exercise plus isoprenaline injection. For baseline HR analysis, continuous ECG data were collected from WT and EHD3−/− mice for 1 hour. Recordings were obtained every 48 hours at the same time of day. For stress tests, mice were run on a treadmill for a maximum of 45 minutes or until exhaustion and then immediately injected with isoprenaline (0.5 mg/kg). Nonsustained and sustained arrhythmias were identified using standard ECG analy-sis guidelines.10

In a separate group of mice than those implanted with radiotele-meters, surface ECG recordings were obtained under anesthesia with 1% to 2% isoflurane. Three needle electrodes were placed subcutane-ously in the standard limb configuration. For each mouse, 15 minutes of continuous data were sampled at 4 kHz with a PowerLab 4/30 interface (AD Instruments, Colorado Springs, CO). Analysis was per-formed offline using LabChart 7 Pro (AD Instruments).

SR Ca LoadSR Ca load and Ca handling were assessed from isolated ventricular myocytes as previously described.11 [Ca]

SRT was calculated through

the pseudoratio as previously described11 with [Ca]d assumed to be

120 nmol/L for all mice.

Generation of EHD3−/− MouseEHD3−/− mice were generated as previously described.12 DNA was isolated from tail clips of 10-day-old mice, and mice were genotyped by polymerase chain reaction. Three primers in a single duplex poly-merase chain reaction reaction amplified the WT allele (377 bp) and the deleted allele (488 bp, Figure 1A). To test the in vivo cardiac-intrinsic roles of EHD3, we used a conditional null mutant allele in which the 5′ untranslated region and exon 1 of the mouse EHD3 gene (Ehd3) were flanked by LoxP sites (Ehd3f/f) and therefore are deleted in the presence of Cre recombinase. We selectively eliminated EHD3 in cardiomyocytes by using alpha subtype myosin heavy chain (αMHC)-Cre knock-in mice13; homozygous conditional knockout mice are referred to as αMHC-Cre; Ehd3f/f or cKO. Mice displayed lack of EHD3 by polymerase chain reaction and immunoblot. Mice were born at expected Mendelian ratios and were healthy and fertile with body weights comparable with their WT littermates.

AntibodiesThe following antibodies were used to conduct this study: af-finity-purified rabbit polyclonal antibody directed at human EHD4 (SHRKSLPKAD), rabbit polyclonal anti-EHD1 (abcam, Cambridge, MA), mouse monoclonal anti-NCX1 (Swant, Bellinzona, Switzerland), rabbit polyclonal anti–ankyrin-B,14 mouse monoclonal anti-Ca

v1.2,15 rabbit polyclonal anti–β

1- and anti–β

2-AR (Santa Cruz

Nonstandard Abbreviations and Acronyms

[Ca]SRT total sarcoplasmic reticulum Ca concentration

AP action potential

APD action potential duration

β-AR β-adrenergic receptor

Cav1.2 L-type Ca channel type 1.2

cKO conditional knockout

EHD Eps15 homology domain

HR heart rate

ICa,L L-type Ca channel–mediated membrane current

INCX Na/Ca exchanger–mediated membrane current

NCX Na/Ca exchanger

SR sarcoplasmic reticulum

WT wild type

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 4: Final Circ Res Publication

70 Circulation Research June 20, 2014

Biotechnology, Dallas, TX), rabbit polyclonal anti-Nav1.5,16 actin

(Santa Cruz Biotechnology), rabbit polyclonal anti-SERCA (Santa Cruz), mouse monoclonal antiphospholamban (abcam, Cambridge, MA), and rabbit monoclonal anti–calsequestrin 2 (abcam).

Immunoblots and ImmunostainingImmunoblots of whole heart lysates were performed as described.17 Briefly, whole hearts were harvested from WT and EHD3−/− adult, age-matched littermates and immediately placed into ice cold ho-mogenization buffer (in mmol/L: 50 Tris-HCl, 10 NaCl, 320 sucrose, 5 EDTA, 2.5 EGTA; supplemented with 1:1000 protease inhibitor cocktail and 1:1000 phenylmethanesulfonyl fluoride [Sigma]). After quantification, tissue lysates were analyzed on Mini-PROTEAN tetra cell (BioRad) on a 4% to 15% precast TGX gel (BioRad). Gels were transferred to a nitrocellulose membrane using the Mini-PROTEAN tetra cell (BioRad). Membranes were blocked for 1 hour at room tem-perature using a 3% BSA solution or 5% milk solution and incubated with primary antibody overnight at 4°C. Densitometry analysis was done using ImageLab software (BioRad). For all experiments, protein values were normalized against an internal loading control (actin or GAPDH).

StatisticsAll values are presented as mean±SEM. When appropriate, data were analyzed using a 2-tailed Student t test. NCX–mediated membrane current (I

NCX) values were analyzed using a 1-tail Student t test with

the WT myocytes predicted to have the larger mean value based on previous work.6 P values <0.05 were considered significant.

ResultsEHD3−/− Mice Display Chamber Dilation and Reduced Ejection FractionTo evaluate the role of EHD3 in cardiac physiology, we first ex-amined the structure and function of the heart in EHD3−/− and WT mice (Figure 1A). By gross examination, whole heart mor-phology was significantly altered in EHD3−/− mice. Both atria and ventricles were larger in excised EHD3−/− hearts compared with those of WT littermates (Figure 1B), and we observed increased heart weight to body weight ratio in EHD3−/− mice compared with WT littermates (EHD3−/−: 7.98±0.26 mg/g; WT: 6.79±0.21 mg/g; P=0.004; Online Table I). We performed echocardiograms on age-matched WT and EHD3−/− mice to assess whether EHD3 deficiency directly affected the de-velopment, structure, or contractility of the adult heart. Both fractional shortening and ejection fraction were decreased in EHD3−/− mice (Figure 1F and 1G), although stroke volume was preserved (Figure 1H; likely attributable to a larger left ventricular diameter in EHD3−/− hearts [Figure 1C]). Both an-terior and posterior systolic wall thicknesses were decreased in EHD3−/− mice (Figure 1D and 1E); however, we observed no difference in diastolic wall thickness between genotypes. Chamber dilation and stroke volume phenotypes were present

Figure 1. Eps15 homology domain 3–deficient (EHD3−/−) mice display abnormal cardiac function. A, Genotyping of wild-type (WT), EHD3+/−

, and EHD3−/− mice assessed by polymerase chain reaction of DNA isolated from tail clips of 8- to 10-day-old mice. B, Representative whole hearts isolated from WT and EHD3−/− mice. C, Left ventricular (LV) systolic diameter is increased in EHD3−/− hearts. D, Systolic anterior wall thickness is increased in EHD3−/− hearts. E, Systolic posterior wall thickness is increased in EHD3−/− hearts. F, Fractional shortening is decreased in EHD3−/− hearts. G, Ejection fraction is decreased in EHD3−/− hearts. H, Stroke volume is preserved across genotypes. I and J, Representative echocardiographic images from WT and EHD3−/− mice; *P<0.05; n=7 WT; n=6 EHD3−/−.

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 5: Final Circ Res Publication

Curran et al EHD3-Dependent Membrane Protein Trafficking 71

in EHD3−/− mice as early as 4 weeks, although at this age we observed no difference in fractional shortening or ejection frac-tion between genotypes (Online Tables I and II).

EHD3−/− Mice Display Bradycardia, HR Variability, and Cardiac Conduction DefectsWe next assessed the in vivo role of EHD3 in cardiac electric signaling. ECG data were acquired from conscious, ambula-tory WT and EHD3−/− mice with subcutaneously implanted radiotelemeters. Notably, we observed a significant reduc-tion in baseline HR in EHD3−/− mice and WT littermates (Figure 2A). We examined HR variability by conducting a fast Fourier transform analysis of HR. In addition to brady-cardia, EHD3−/− mice displayed increased HR variability com-pared with controls, with the frequency of lower HRs more prominent in EHD3−/− mice (Figure 2B and 2C). Analysis of individual ECG traces revealed numerous electric pheno-types in EHD3−/− mice. Unlike WT mice (Figure 2D and 2E), EHD3−/− mice were prone to bursts of irregular HR reflected as increased variability in the R-R interval and sinus pause in se-lected segments of the ECG recordings (Figure 2C). EHD3−/−, but not WT, mice commonly displayed significant R-R inter-val variability for periods lasting >1 minute and as long as 5 minutes under baseline conditions. This phenotype became even more apparent when EHD3−/− mice were stimulated with isoprenaline injection (Figure 2E). With the exception of episodes of sinus pause or atrioventricular block, we did not observe significant differences in either P-R interval or QRS duration between WT and EHD3−/− mice (Online Figure II).

In addition to increased R-R variability, EHD3−/− mice con-sistently displayed type II atrioventricular conduction block,

with exacerbation of the phenotype after isoprenaline stimula-tion (Figure 2F). The type II block was characterized by 2 P waves without subsequent QRS complexes followed by a sin-gle extended R-R interval and resumption of normal rhythm (Figure 2F). This pattern of atrioventricular block was main-tained for as long as 15 minutes in EHD3−/− mice and was nev-er observed in WT littermates. Collectively, we conclude that normal cardiac automaticity and conduction require EHD3.

EHD proteins have been implicated in membrane receptor expression, internalization, and recycling in other cell sys-tems.5,18,19 We therefore assessed the impact of EHD3 deficiency on β-AR signaling in vivo. In conscious mice, the maximum HR in response to isoprenaline injection (either low or high dose) was equivalent in WT and EHD3−/− mice; however, the duration of this response was significantly shorter in the EHD3−/− mice (Online Figure IIIA and IIIB). To limit any variation in the data resulting from the physical handling of the mice, we performed surface ECG recordings on sedated mice with continuous moni-toring of HR (Online Figure IIIC). After injection of isoprena-line (0.5 mg/kg), WT mice had a significantly larger increase in HR compared with EHD3−/− mice (79±10 versus 46±5 bpm; P<0.05; Online Figure IIID). Outside the sinus pause and atrio-ventricular block, we observed no other evidence of arrhyth-mias (ie, premature ventricular complexes, or atrial flutter or fibrillation) in EHD3−/− mice at baseline, after isoprenaline in-jection, or after isoprenaline injection and exercise.

Based on these data, we hypothesized that EHD3−/− myo-cytes would display altered β

1-AR trafficking and membrane

expression. However, by immunoblot, we observed increased expression of β

1- and β

2-ARs in EHD3−/− hearts (Online Figure

IIIE–IIIH). Because immunoblots are unable to discriminate

Figure 2. Eps15 homology domain 3–deficient (EHD3−/−) mice display heart rate variability and conduction defects. A, EHD3−/− mice display bradycardia compared with wild-type (WT) littermates. B, Distribution of heart rate (HR) variability of WT and EHD3−/− mice. A and B, n=5 for each genotype, P<0.05. C, EHD3−/− mice were prone to bursts of irregular R-R interval variability not seen in WT mice. D, Selected 5-second trace of a WT mouse radiotelemetry demonstrating regular HR. E, Selected 5-second trace of an EHD3−/− mouse demonstrating a burst of high R-R interval variability. F, Trace of an EHD3−/− mouse demonstrating type II atrioventricular conduction block. *P<0.05 vs WT.

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 6: Final Circ Res Publication

72 Circulation Research June 20, 2014

between proteins embedded in the surface membrane and those confined to internal compartments (including endosomes), we used radiolabeled β

1-AR–specific antagonist, 3H-CGP-12177, to

assay potential changes in cell surface β1-AR density in intact,

isolated ventricular cardiomyocytes. Using this assay, we ob-served no significant difference in β

1-AR surface density (Online

Figure III I). Notably, however, by immunostaining and confocal analysis of WT and EHD3−/− adult ventricular cardiomyocytes, we observed a significant population of β

1-AR expression in the

perinuclear region of EHD3−/− myocytes (Online Figure IIIJ). Based on these data, we hypothesize that although anterograde membrane trafficking of the β

1-AR is EHD3 independent, β

1-AR

endosomal recycling may be compromised in EHD3−/− myocytes.

EHD3−/− Myocytes Display Abnormal Myocyte ExcitabilityWe directly evaluated the effect of EHD3 deficiency on car-diomyocyte membrane excitability. APD was significantly shorter in ventricular myocytes isolated from EHD3−/− hearts (Figure 3A, 3B, and 3D; n=7, both WT and EHD3−/−). At all intervals, APD was significantly shorter in EHD3−/− myocytes, although the shortening was most significant in the early phases of the AP (APD at 50% and 75% repolarization [APD

50, APD

75],

Figure 3A, 3B, and 3D). Furthermore, consistent with in vivo data, EHD3−/− myocytes displayed a blunted β-adrenergic re-sponse to isoprenaline treatment (100 nmol/L) compared with WT littermates (Figure 3C–3E).

EHD3−/− Myocytes Display Increased SR Ca Load and Ca Spark FrequencyGiven the abbreviated AP, but no reduction of the stroke volume in EHD3−/− mice, we hypothesized that EHD3−/− myocytes may have an increased [Ca]

SRT. This higher Ca load would increase

the Ca sensitivity of the SR Ca release channel, ryanodine re-ceptor, making it more prone to release even with a smaller Ca trigger, similar to the physiological mechanism underlying increased contractility during catecholamine stimulation.20 To test this hypothesis, we fielded stimulated isolated cardiac my-ocytes at 0.5 Hz and assessed [Ca]

SRT using caffeine-stimulated

Ca release.11 Figure 4 shows that no differences in cytosolic

Ca concentration transients were observed between WT and EHD3−/− myocytes (Figure 4A–4D), although there was a sig-nificant increase in [Ca]

SRT in the latter (186±25 versus 262±20

μmol/L; Figure 4E). Increased Ca sensitivity of the ryanodine receptor is further reflected by an increased spark frequency found in intact ventricular EHD3−/− myocytes (Figure 4F–4H). These data support the hypothesis that [Ca]

SRT is upregulated to

maintain contractility in EHD3−/− myocytes, indicating a role for EHD3 in Ca homeostasis.

EHD3 Is Required for NCX and Cav1.2 Membrane Targeting in HeartAs a first step toward determining the ionic basis for reduced APD, increased SR Ca2+ load, and increased spark frequency in EHD3−/− myocytes, we screened for likely candidates by performing parameter sensitivity analysis on a well-validat-ed mathematical model of the mouse ventricular AP (Online Figure IV). This analysis revealed that, among sarcolemmal ion channels/transporters, the L-type Ca2+ current and NCX had the greatest influence on APD and SR Ca2+ load, respec-tively, in a manner consistent with experimentally measured changes. More specifically, the model predicts that loss of L-type Ca2+ channel membrane targeting would produce the greatest decrease in APD, whereas loss of NCX would result in the greatest increase in SR Ca2+ load.

Based on WT and EHD3−/− APD and [Ca]SRT

data and subsequent mathematical predictions, we first evaluated the expression of NCX and Ca

v1.2 by immunoblot. Ca

v1.2 and

NCX expression were reduced by 24% and 17%, respectively (Figure 5A and 5B; P<0.05). Conversely, no differences in sarco/endoplasmic reticulum Ca2+ ATPase 2, phospholamban, or calsequestrin expression were identified between genotypes (Online Figure V). Together with computational analysis, these results provide a possible mechanism for abnormal AP and SR Ca2+ properties in EHD3−/− myocytes.

Decreased expression of NCX and Cav1.2 was paralleled

by decreased membrane expression of both NCX and Cav1.2

in EHD3−/− mice by immunostaining and confocal analysis of isolated ventricular cardiomyocytes. Representative images are shown in Figure 5C to 5H. In WT myocytes, the NCX

Figure 3. Eps15 homology domain 3–deficient (EHD3−/−) myocytes display shortened action potential duration (APD). A, Representative baseline AP waveform in wild-type (WT; black) and EHD3−/− (gray) mice. B, APD at baseline. C, Representative AP waveform after application of 100 nmol/L isoprenaline (Iso) in WT (black) and EHD3−/− (gray) myocytes. D, APD at 50%, 75%, and 90% repolarization (APD50, APD75, and APD90) (±Iso) in WT and EHD3−/− myocytes. E, Change in APD after Iso application. For data in figure, *P<0.05 vs WT; #P<0.05 vs WT+Iso; n=7 mice/genotype; **P<0.05 vs WT+Iso.

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 7: Final Circ Res Publication

Curran et al EHD3-Dependent Membrane Protein Trafficking 73

clearly localizes to both the sarcolemmal membrane and along the transverse-tubule network (Figure 5C and 5E). However, in EHD3−/− myocytes, the localization to these 2 domains is decreased with a striking perinuclear distribution that is not observed in WT cells (Figure 5D and 5F).

Cav1.2 localization was also disrupted by EHD3 deficien-

cy. In WT myocytes, Cav1.2 staining shows a typical striated

pattern (Figure 5G).21 This pattern was disrupted in EHD3−/− myocytes (Figure 5H). Although some striations were evident, Ca

v1.2-dependent staining was largely diffused and not local-

ized to any specific subcellular domain. Using an antibody specific for N-cadherin and the lipophilic membrane marker, di-8-ANEPPS, we observed no structural changes to interca-lated disc and the t-tubule system in EHD3−/− myocytes, in-dicating that myocyte ultrastructure is EHD3 independent (Figure 5I–5L).

Finally, to quantify potential functional differences, NCX (I

NCX) and Ca

v1.2 (I

Ca,L) currents were assessed in iso-

lated ventricular myocytes using whole-cell patch clamp. Data in Figure 5M to 5Q show that both I

Ca,L and I

NCX were

significantly reduced in EHD3−/− myocytes. INCX

was sig-nificantly decreased in EHD3−/− myocytes compared with that in WT myocytes with a 47% reduction in peak current (Figure 5M and 5N). At baseline, I

Ca,L was decreased 60% in

EHD3−/− myocytes (Figure 5O and 5Q). Treatment with iso-prenaline (100 nmol/L) increased I

Ca,L similarly in both WT

and EHD3−/− myocytes (Figure 5P and 5Q).In summary, our data obtained using 3 different approaches

support a mechanistic in vivo role of EHD3 for the membrane targeting of both NCX and the Ca

v1.2 in primary adult cardio-

myocytes. Importantly, loss of membrane protein expression phenotypes was EHD3 dependent (versus a compensatory response to a failing myocardium) as we observed defects in channel or transporter expression in EHD3−/− primary car-diomyocytes (Online Figure VI) isolated at postnatal day 1 (weeks before observed functional phenotypes).

Dysregulated Ankyrin-B Trafficking in EHD3−/− MyocytesOur findings implicate EHD3 for targeting select ion channels/transporters to the myocyte membrane. However, to further under-stand the mechanistic link between EHD3 and membrane protein targeting, we investigated the regulation of ankyrin-B, a membrane adapter protein previously linked with both NCX and EHD3.6,22–24 Notably, ankyrin-B expression is significantly decreased (≈40%) in EHD3−/− hearts (Figure 6A and 6B). Furthermore, EHD3−/− my-ocytes show altered ankyrin-B staining. We observed a decrease in overall intensity with increased perinuclear ankyrin-B staining in

Figure 4. Eps15 homology domain 3–deficient (EHD3−/−) myocytes display increased sarcoplasmic reticulum (SR) Ca load and Ca spark frequency. A and B, Representative line-scan images of 0.5-Hz Ca transients in wild-type (WT) and EHD3−/− myocytes. C and D, Representative Ca transients of WT and EHD3−/− myocytes, normalized to F0. E, Average SR Ca load in WT (186±25 μmol/L; n=9) and EHD3−/− (262±20 μmol/L; n=14) myocytes. F, Spark frequency is increased in EHD3−/− compared with WT myocytes (0.55±0.15 vs 0.1±0.05 sparks·100 μm−1·s−1; n=9 each). Representative confocal line-scan images of sparks in WT (G) and EHD3−/− (H) myocytes.*P<0.05 vs WT.

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 8: Final Circ Res Publication

74 Circulation Research June 20, 2014

Figure 5. Eps15 homology domain 3 (EHD3) regulates Na/Ca exchanger (NCX) and L-type Ca channel type 1.2 (Cav1.2) membrane targeting and function in heart. A, NCX expression is decreased by 17% in EHD3-deficient (EHD3−/−) hearts (left); representative immunoblot (right). Space between lanes denotes that data were collected from noncontiguous lanes of the same gel. B, Cav1.2 expression is decreased by 24% in EHD3−/− hearts (left); representative immunoblot (right); n=8 hearts/wild type (WT), n=9 hearts/EHD3−/− for immunoblots (*P<0.05 vs WT). C to L, Representative confocal images of WT and EHD3−/− isolated ventricular cardiomyocytes displaying localization of NCX, Cav1.2, N-cadherin, and di-8-ANEPPs (T-tubule marker). Bar=10 μm for all images. M and N, Whole-cell NCX–mediated membrane current (INCX) was decreased in EHD3−/− myocytes (n=10 myocytes/genotype). Specifically, peak INCX was decreased by 47% in EHD3−/− myocytes. O and P, EHD3−/− myocytes display reduced CaV1.2–mediated membrane current (ICa,L) compared with WT at baseline and after application of 100 nmol/L isoprenaline (Iso; n=6 myocytes/genotype, *P<0.05 vs WT). Q, Peak ICa,L is significantly decreased in EHD3−/− myocytes±Iso. #P<0.05 vs WT+Iso.

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 9: Final Circ Res Publication

Curran et al EHD3-Dependent Membrane Protein Trafficking 75

EHD3−/− myocytes compared with WT myocytes (Figure 6C–6J). Defects in ankyrin-B expression and localization are present from birth and are directly EHD3 dependent, because viral reintroduc-tion of green fluorescent protein-EHD3 is sufficient to rescue the expression of ankyrin-B in EHD3−/− myocytes (Online Figure VII). Based on these findings, we predict that ankyrin-B plays a key nodal role for EHD3-dependent membrane targeting in heart.

Cardiac-Specific EHD3−/− Mice Display Structural and Electric PhenotypesGlobal EHD3−/− mice display bradycardia, HR variability, conduction defects, and structural phenotypes (Figures 1 and 2). However, these parameters may be influenced by cardi-ac-extrinsic factors (ie, nervous system). To test the in vivo cardiac-intrinsic roles of EHD3 directly, we used a conditional null mutant allele in which exon 1 of the mouse EHD3 gene (Ehd3) was flanked by LoxP sites (Ehd3f/f) and therefore are deleted in the presence of Cre recombinase. We selectively eliminated EHD3 in cardiomyocytes by using αMHC-Cre knock-in mice13; homozygous conditional knockout mice are referred to as αMHC-Cre; Ehd3f/f or cKO.

Data presented in Figure 7 demonstrate a striking similarity between the EHD3−/− and the EHD3−/− cKO mice. Resting HR was equally depressed in the EHD3−/− cKO and EHD3−/− mice compared with WT (Figure 7A; P<0.05) with a shifted distribu-tion of HR similar to the EHD3−/− model (Figure 7B). Moreover, we observed similar conduction defects in the EHD3−/− cKO mice: a high HR variability at baseline (Figure 7C), a high incidence of sinoatrial node (SAN) pause (Figure 7D), and atrioventricular conduction block (Figure 7E) after isoprena-line treatment. SAN pause and atrioventricular block were also present at a lower rate in the absence of isoprenaline stimula-tion (not shown). These conduction disorders were never ob-served in the WT mice. Importantly, at the level of the single

myocyte, cKO mouse myocytes display significant loss of INCX

and I

Ca,L (Figure 7F–7I). Specifically, EHD3−/− cKO peak I

NCX

was reduced ≈43%, and ICa,L

was reduced ≈53% compared with WT (both P<0.001). Beyond electric phenotypes, the structural and function phenotype of EHD3−/− global and cKO mice as-sessed by echocardiography were similar in nearly all respects (Online Table III). Together, these new data indicate that cardiac EHD3 is critical for normal cardiac structural and electric phe-notypes. Furthermore, these data support cardiac-intrinsic roles for EHD3 in regulating normal cardiac structure and function.

DiscussionAnterograde and retrograde protein trafficking and endocytic protein recycling are often overlooked cellular systems. As an often membrane-centric discipline, we take for granted that these pathways are present and critical for cell function. Our lack of un-derstanding may stem from the particular difficulties in studying these systems because endosomes undergo an incredibly com-plex maturation process that makes investigating particular pro-cesses or steps within these pathways problematic, particularly in vivo. Relatively little is known about the particular proteins and enzymes that may be involved in cell membrane trafficking pro-cesses. Research during the past 15 years has only just started to isolate single proteins and enzymes that are involved in traffick-ing. Proteins such as Arf6 (ADP-ribosylation factor 6), SNARE (soluble NSF attachment protein) complexes, BIN1 (amphiphy-sin 2), EHD1–4 among others are now the focus of efforts to fur-ther understand protein trafficking in all cell types.4,25–27 Recently, a mutation in EHD3 was linked to major depressive disorder in humans, indicating that trafficking systems simply do not main-tain cellular health but may be the primary cause for medical disorders.28 We recently identified EHD3 in the heart and demon-strated an association of this protein with cardiac disease.7

Figure 6. Dysregulation of ankyrin-B expression and localization in Eps15 homology domain 3–deficient (EHD3−/−) heart. A, Ankyrin-B expression is reduced ≈40% in EHD3−/− hearts (n=9 wild type [WT], 8 EHD3−/−; *P<0.05 vs WT). B, Representative immunoblot of ankyrin-B expression in WT and EHD3−/− hearts. Space between lanes denotes that data were collected from noncontiguous lanes of the same gel. C to J, Representative confocal images detailing ankyrin-B trafficking dysregulation in EHD3−/− myocytes (*P<0.05 vs WT).

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 10: Final Circ Res Publication

76 Circulation Research June 20, 2014

This report is the first to detail the functional in vivo roles of EHD3 in the heart. The key findings of this investigation are as follows: (1) EHD3−/− hearts are significantly larger with preserved cardiac output; (2) EHD3−/− mice display bradycar-dia, HR variability, type II conduction block, and a blunted response to β-AR stimulation; (3) isolated EHD3−/− myocytes display a significantly abbreviated APD and a blunted response to β-AR stimulation; (4) EHD3−/− myocytes have a significant reduction in expression, targeting, and function of Ca

v1.2 and

NCX; (5) EHD3−/− myocytes display a significantly larger [Ca]

SRT and a higher frequency of sparks compared with WT;

(6) ankyrin-B function is directly altered by EHD3 loss; and

(7) EHD3−/− phenotypes can be directly linked with EHD3 ex-pressed in myocytes. Together, these data highlight the impor-tance of endosome-based pathways for normal cardiac function.

EHD proteins (EHD1–4) are key regulators of membrane pro-tein targeting in other tissue and cell types.29–31 This family of pro-teins has high homology among the members (ranging from 71% to 86%), with EHD1 and EHD3 sharing the highest homology.32 This degree of homology infers similar and potentially redun-dant cellular function. For example, in a Caenorhabditis elegans model lacking the EHD ortholog, RME-1, expression of human EHD1–4 is sufficient to rescue RME-1–dependent protein traf-ficking.4 These findings seem relevant to our study. For example,

Figure 7. Cardiac-specific Eps15 homology domain 3–deficient (EHD3−/−) mice phenocopy global EHD3−/− mouse electric phenotypes. A, EHD3−/− conditional knockout (cKO) mice display bradycardia compared with wild-type (WT) littermates. B, Distribution of heart rate variability of WT and EHD3−/− cKO mice. A and B, n=4 for each genotype, P<0.05. C, EHD3−/− cKO mice were prone to bursts of irregular R-R interval variability at baseline not seen in WT mice. D, Selected trace of cKO mouse radiotelemetry demonstrating sinus pause and rate variability after application of isoproterenol (Iso). E, Selected trace of cKO mouse demonstrating atrioventricular conduction block phenotypes. F and G, Whole-cell INCX was decreased in the EHD3 cKO. H and I, ICa,L is decreased at both baseline (H) and after treatment with isoproterenol (I). *P<0.05 vs WT. ICa,L indicates Cav1.2–mediated membrane current; and INCX, Na/Ca exchanger–mediated membrane current.

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 11: Final Circ Res Publication

Curran et al EHD3-Dependent Membrane Protein Trafficking 77

although select protein trafficking pathways are disrupted and EHD3−/− hearts display multiple in vivo phenotypes, the viabil-ity of EHD3−/− mice strongly support that additional pathways are present for ion channel and transporter membrane expression. Interestingly, EHD1 expression was significantly upregulated in the EHD3−/− mice (Online Figure VIII). We speculate that this up-regulation may represent a compensatory response of the heart to maintain protein trafficking. We also observed reduced expression of EHD4 in the EHD3−/− hearts (Online Figure VIII). Although nothing is currently known regarding the role of either EHD1 or EHD4 proteins in heart, these data suggest that the EHD family may collaborate to regulate endosome-based trafficking pathways.

EHD3−/− and EHD3−/− cKO mice display bradycardia, rate variability, and conduction defects. The SAN is the pacemaker of the heart, and the atrioventricular node is critical for the proper conduction of the AP from the atria to the ventricles. Both the SAN and atrioventricular node depend on voltage-gated ion channels and exchangers to maintain proper elec-tric activity. In particular, the SAN is known to rely on both the I

Ca,L and I

NCX for automaticity, whereas I

Ca,L activity in the

atrioventricular node is thought to be indispensable for con-duction through this junction.33–36 Notably, EHD3 is expressed in the SAN of both mice and canine hearts (Online Figure IX). Although additional studies will be necessary to identify specific roles of EHD3 in these critical cardiac cell types, our initial findings link EHD3 function with cardiac automaticity.

EHD3-dependent trafficking in the heart may represent a nodal control pathway for regulating protein trafficking in response to acute or chronic stress. Here, we demonstrate that EHD3-dependent mechanisms are broadly implicated in the subcellular trafficking and localization of many of the proteins involved in excitation–contraction coupling. NCX expression is increased in nearly all reports examining protein expression in heart failure, including in humans.37,38 Although initially compensatory, as the severity of heart failure progresses, the increased NCX expres-sion becomes maladaptive and arrhythmogenic leading to de-layed afterdepolarizations.39,40 We previously reported that EHD3 expression is increased in numerous causes of heart failure.7 The evidence in this report strongly indicates that EHD3 mediates NCX trafficking in the heart. Together, I

NCX and I

Ca,L make up a

substantial fraction of the whole-cell current during a typical AP. Based on modeling predictions, we expect that loss of I

CaL plays a

critical role in abbreviation of APD in EHD3-null myocytes, espe-cially in the early phases of AP development, whereas loss of I

NCX

determines changes in SR Ca2+ load. Although EHD3 is critical for NCX, Ca

v1.2, and ankyrin-B targeting, it is important to note

that EHD3 also likely targets additional membrane proteins and that observed phenotypes in the EHD3−/− heart may not be related to calcium-based signaling pathways. Future experiments will be critical to define the full spectrum of EHD3 targets in heart.

Finally, additional work will be necessary to define the struc-tural phenotypes observed in EHD3−/− hearts during develop-ment, as well as the relationship between structural and electric phenotypes. Despite chamber dilation, reduced ejection fraction, and reduced stroke volume, the myopathy phenotype in EHD3−/− mice is complex and may not simply represent a pure maladap-tive or physiological hypertrophy phenotype. In fact, despite observing increased expression of slow skeletal muscle troponin (consistent with maladaptive hypertrophy, P<0.05), we did not

observe elevated expression of atrial natriuretic peptide, brain natriuretic peptide, or β-myosin heavy chain expression mRNA in EHD3−/− hearts (P=NS). Future experiments will be critical to characterize the role of EHD3 fully in cardiac development as well as the specific roles of EHD3 in heart failure. Importantly, EHD3−/− NCX targeting phenotypes were directly attributable to EHD3 deficiency, and not a compensatory response to functional decompensation, because dysregulated NCX targeting was pres-ent from birth, and rhythm defects were observed as early as 4 weeks of age in EHD3−/− mice (before depressed cardiac func-tion; Online Figures VI and X). Thus, our data support that EHD3 plays roles in targeting proteins involved in both cardiac struc-tural and electric functions.

Sources of FundingThis work was supported by the National Institutes of Health (HL114252 to J. Curran; HL084583, HL083422, and HL114383 to P.J. Mohler; CA105489, CA87986, CA99163, and CA116552 to H. Band; HL096805 and HL114893 to T.J. Hund), Saving Tiny Hearts Society (P.J. Mohler), and the American Heart Association (P.J. Mohler).

DisclosuresNone.

References 1. Bers DM. Cardiac excitation-contraction coupling. Nature.

2002;415:198–205. 2. Thibault G, Amiri F, Garcia R. Regulation of natriuretic peptide secretion

by the heart. Annu Rev Physiol. 1999;61:193–217. 3. Caplan S, Naslavsky N, Hartnell LM, Lodge R, Polishchuk RS, Donaldson

JG, Bonifacino JS. A tubular EHD1-containing compartment involved in the recycling of major histocompatibility complex class I molecules to the plasma membrane. EMBO J. 2002;21:2557–2567.

4. George M, Ying G, Rainey MA, Solomon A, Parikh PT, Gao Q, Band V, Band H. Shared as well as distinct roles of EHD proteins revealed by bio-chemical and functional comparisons in mammalian cells and C. elegans. BMC Cell Biol. 2007;8:3.

5. Naslavsky N, Rahajeng J, Sharma M, Jovic M, Caplan S. Interactions be-tween EHD proteins and Rab11-FIP2: a role for EHD3 in early endosomal transport. Mol Biol Cell. 2006;17:163–177.

6. Gudmundsson H, Hund TJ, Wright PJ, et al. EH domain proteins regulate cardiac membrane protein targeting. Circ Res. 2010;107:84–95.

7. Gudmundsson H, Curran J, Kashef F, Snyder JS, Smith SA, Vargas-Pinto P, Bonilla IM, Weiss RM, Anderson ME, Binkley P, Felder RB, Carnes CA, Band H, Hund TJ, Mohler PJ. Differential regulation of EHD3 in human and mammalian heart failure. J Mol Cell Cardiol. 2012;52:1183–1190.

8. Le Scouarnec S, Bhasin N, Vieyres C, et al. Dysfunction in ankyrin-B-dependent ion channel and transporter targeting causes human sinus node disease. Proc Natl Acad Sci U S A. 2008;105:15617–15622.

9. DeGrande S, Nixon D, Koval O, Curran JW, Wright P, Wang Q, Kashef F, Chiang D, Li N, Wehrens XH, Anderson ME, Hund TJ, Mohler PJ. CaMKII inhibition rescues proarrhythmic phenotypes in the model of hu-man ankyrin-B syndrome. Heart Rhythm. 2012;9:2034–2041.

10. Mitchell GF, Jeron A, Koren G. Measurement of heart rate and Q-T inter-val in the conscious mouse. Am J Physiol. 1998;274:H747–H751.

11. Curran J, Hinton MJ, Ríos E, Bers DM, Shannon TR. Beta-adrenergic enhance-ment of sarcoplasmic reticulum calcium leak in cardiac myocytes is mediated by calcium/calmodulin-dependent protein kinase. Circ Res. 2007;100:391–398.

12. George M, Rainey MA, Naramura M, Foster KW, Holzapfel MS, Willoughby LL, Ying G, Goswami RM, Gurumurthy CB, Band V, Satchell SC, Band H. Renal thrombotic microangiopathy in mice with combined deletion of endo-cytic recycling regulators EHD3 and EHD4. PLoS One. 2011;6:e17838.

13. Agah R, Frenkel PA, French BA, Michael LH, Overbeek PA, Schneider MD. Gene recombination in postmitotic cells. Targeted expression of Cre recombinase provokes cardiac-restricted, site-specific rearrangement in adult ventricular muscle in vivo. J Clin Invest. 1997;100:169–179.

14. Abdi KM, Mohler PJ, Davis JQ, Bennett V. Isoform specificity of ankyrin-B: a site in the divergent C-terminal domain is required for intramolecular association. J Biol Chem. 2006;281:5741–5749.

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 12: Final Circ Res Publication

78 Circulation Research June 20, 2014

15. Balijepalli RC, Foell JD, Hall DD, Hell JW, Kamp TJ. Localization of cardiac L-type Ca(2+) channels to a caveolar macromolecular signaling complex is required for beta(2)-adrenergic regulation. Proc Natl Acad Sci U S A. 2006;103:7500–7505.

16. Hund TJ, Koval OM, Li J, Wright PJ, Qian L, Snyder JS, Gudmundsson H, Kline CF, Davidson NP, Cardona N, Rasband MN, Anderson ME, Mohler PJ. A β(IV)-spectrin/CaMKII signaling complex is essential for mem-brane excitability in mice. J Clin Invest. 2010;120:3508–3519.

17. Kline CF, Wright PJ, Koval OM, Zmuda EJ, Johnson BL, Anderson ME, Hai T, Hund TJ, Mohler PJ. βIV-Spectrin and CaMKII facilitate Kir6.2 regulation in pancreatic beta cells. Proc Natl Acad Sci U S A. 2013;110:17576–17581.

18. Posey AD Jr, Swanson KE, Alvarez MG, Krishnan S, Earley JU, Band H, Pytel P, McNally EM, Demonbreun AR. EHD1 mediates vesicle traffick-ing required for normal muscle growth and transverse tubule development. Dev Biol. 2014;387:179–190.

19. Guilherme A, Soriano NA, Furcinitti PS, Czech MP. Role of EHD1 and EHBP1 in perinuclear sorting and insulin-regulated GLUT4 recycling in 3T3-L1 adipocytes. J Biol Chem. 2004;279:40062–40075.

20. Györke I, Hester N, Jones LR, Györke S. The role of calsequestrin, triadin, and junctin in conferring cardiac ryanodine receptor responsiveness to lu-minal calcium. Biophys J. 2004;86:2121–2128.

21. Scriven DR, Dan P, Moore ED. Distribution of proteins implicated in excitation-contraction coupling in rat ventricular myocytes. Biophys J. 2000;79:2682–2691.

22. Li ZP, Burke EP, Frank JS, Bennett V, Philipson KD. The cardiac Na+-Ca2+ exchanger binds to the cytoskeletal protein ankyrin. J Biol Chem. 1993;268:11489–11491.

23. Cunha SR, Bhasin N, Mohler PJ. Targeting and stability of Na/Ca ex-changer 1 in cardiomyocytes requires direct interaction with the mem-brane adaptor ankyrin-B. J Biol Chem. 2007;282:4875–4883.

24. Cunha SR, Mohler PJ. Cardiac ankyrins: essential components for de-velopment and maintenance of excitable membrane domains in heart. Cardiovasc Res. 2006;71:22–29.

25. D’Souza-Schorey C, Li G, Colombo MI, Stahl PD. A regulatory role for ARF6 in receptor-mediated endocytosis. Science. 1995;267:1175–1178.

26. Weber T, Zemelman BV, McNew JA, Westermann B, Gmachl M, Parlati F, Söllner TH, Rothman JE. SNAREpins: minimal machinery for membrane fusion. Cell. 1998;92:759–772.

27. Hong TT, Smyth JW, Gao D, Chu KY, Vogan JM, Fong TS, Jensen BC, Colecraft HM, Shaw RM. BIN1 localizes the L-type calcium channel to cardiac T-tubules. PLoS Biol. 2010;8:e1000312.

28. Shi C, Zhang K, Wang X, Shen Y, Xu Q. A study of the combined effects of the EHD3 and FREM3 genes in patients with major depressive disorder. Am J Med Genet B Neuropsychiatr Genet. 2012;159B:336–342.

29. Daumke O, Lundmark R, Vallis Y, Martens S, Butler PJ, McMahon HT. Architectural and mechanistic insights into an EHD ATPase involved in membrane remodelling. Nature. 2007;449:923–927.

30. Galperin E, Benjamin S, Rapaport D, Rotem-Yehudar R, Tolchinsky S, Horowitz M. EHD3: a protein that resides in recycling tubular and vesicular membrane structures and interacts with EHD1. Traffic. 2002;3:575–589.

31. Miliaras NB, Wendland B. EH proteins: multivalent regulators of endocy-tosis (and other pathways). Cell Biochem Biophys. 2004;41:295–318.

32. Pohl U, Smith JS, Tachibana I, Ueki K, Lee HK, Ramaswamy S, Wu Q, Mohrenweiser HW, Jenkins RB, Louis DN. EHD2, EHD3, and EHD4 en-code novel members of a highly conserved family of EH domain-contain-ing proteins. Genomics. 2000;63:255–262.

33. Mangoni ME, Couette B, Bourinet E, Platzer J, Reimer D, Striessnig J, Nargeot J. Functional role of L-type Cav1.3 Ca2+ channels in cardiac pacemaker activity. Proc Natl Acad Sci U S A. 2003;100:5543–5548.

34. Lyashkov AE, Juhaszova M, Dobrzynski H, Vinogradova TM, Maltsev VA, Juhasz O, Spurgeon HA, Sollott SJ, Lakatta EG. Calcium cycling pro-tein density and functional importance to automaticity of isolated sinoatri-al nodal cells are independent of cell size. Circ Res. 2007;100:1723–1731.

35. Marger L, Mesirca P, Alig J, Torrente A, Dubel S, Engeland B, Kanani S, Fontanaud P, Striessnig J, Shin HS, Isbrandt D, Ehmke H, Nargeot J, Mangoni ME. Functional roles of Ca(v)1.3, Ca(v)3.1 and HCN channels in automaticity of mouse atrioventricular cells: insights into the atrioven-tricular pacemaker mechanism. Channels (Austin). 2011;5:251–261.

36. Gao Z, Rasmussen TP, Li Y, et al. Genetic inhibition of Na+-Ca2+ ex-changer current disables fight or flight sinoatrial node activity without af-fecting resting heart rate. Circ Res. 2013;112:309–317.

37. Sipido KR, Volders PG, Vos MA, Verdonck F. Altered Na/Ca exchange activity in cardiac hypertrophy and heart failure: a new target for therapy? Cardiovasc Res. 2002;53:782–805.

38. Sipido KR, Volders PG, Schoenmakers M, De Groot SH, Verdonck F, Vos MA. Role of the Na/Ca exchanger in arrhythmias in compensated hyper-trophy. Ann N Y Acad Sci. 2002;976:438–445.

39. Bers DM, Pogwizd SM, Schlotthauer K. Upregulated Na/Ca exchange is involved in both contractile dysfunction and arrhythmogenesis in heart failure. Basic Res Cardiol. 2002;97(suppl 1):I36–I42.

40. Köster OF, Szigeti GP, Beuckelmann DJ. Characterization of a [Ca2+]i-dependent current in human atrial and ventricular cardiomyocytes in the absence of Na+ and K+. Cardiovasc Res. 1999;41:175–187.

What Is Known?

• Ion channels and transporters require complex trafficking and reten-tion pathways to modulate cardiac cell excitability.

• In noncardiac cell types, endosome pathways play critical roles in membrane protein trafficking, internalization, and recycling.

• Eps15 homology domain (EHD) family gene products were recently identified in heart, associated with membrane protein trafficking in myocytes, and shown to be altered in multiple forms of heart disease.

What New Information Does This Article Contribute?

• Global deficiency of EHD3 in mice results causes in vivo defects in cardiac structure and electric function including chamber dilation, re-duced ejection fraction, bradycardia, and conduction defects.

• EHD3-deficient myocytes display defects in ion channel and trans-porter expression, membrane localization, and function. These defects are associated with loss of the membrane adapter protein ankyrin-B.

• Mice that selectively lack EHD3 in cardiomyocytes show similar cardiac phenotypes of global EHD3-deficient mice, demonstrating a key role of the EHD3-based endosome pathway in cardiovascular physiology.

Cardiac excitability is governed by the synchronized activities of a host of membrane-bound ion channels, transporters, and

receptors. Although the field has gained significant insight into the pathways governing membrane protein structure and bio-physical function in health and disease, little is known about the cellular mechanisms that regulate membrane protein expres-sion, trafficking, and internalization at baseline or in response to acute or chronic stress. Because these mechanisms are fun-damental for cardiac structural and electric remodeling in heart failure, new in vivo studies focused on these pathways are es-sential for generating new therapeutic targets for disease. Our results using new in vivo models demonstrate key roles for the cardiac endosome-based system for membrane protein target-ing. Furthermore, these data show that lack of the endosome EHD3 protein results in both cardiac structural and electric phe-notypes. Together, our findings provide new in vivo data that link endosome-based intracellular protein trafficking pathways with the expression, membrane targeting, and function of key cardiac membrane proteins. Because altered EHD3 levels have been pre-viously linked with multiple forms of heart failure, our findings may identify new markers and therapeutic avenues for the diag-nosis and treatment of cardiovascular disease.

Novelty and Significance

at Ohio State University--Columbus on September 12, 2014http://circres.ahajournals.org/Downloaded from

Page 13: Final Circ Res Publication

Supplemental Material

Methods

Mathematical modeling and sensitivity analysis Sensitivity analysis was performed, as

described1 on a mathematical model of the mouse ventricular myocyte action potential and

calcium transient to identify likely candidates for observed changes in AP and SR Ca2+ load2

Briefly, maximal conductances of sarcolemmal ion channels/transporters in the model were

perturbed one parameter at a time +15% and -15%. Action potential duration at 90%

repolarization (APD90) and maximal diastolic Ca2+ concentration in the junctional sarcoplasmic

reticulum ([Ca2+]JSR) were determined following steady-state (change in APD < 0.1%) pacing at

a cycle length of 500 ms. For each property (X) and parameter (p), sensitivity was calculated

according to equation [1] and expressed relative to maximal value for all parameters.

, 15% , 15%, 0.3

p pX p

con

X XS

X

[1]

Generation of EHD3-/- mouse. LoxP sequences were inserted to flank exon 1. Cre/loxP-

mediated recombination resulted in exon deletion. DNA was isolated from tail clips of 10 day

old mice, and mice were genotyped by PCR. Three primers in a single duplex PCR reaction

amplified the WT allele (377 bp) and the deleted allele (488 bp). Primers were as follows: 5’

CAA CAA GAG TGT CAG GAA ACC TGA ACT A-3’; 5’-CTG GGA AAC TGC AGA ACA TCA

GGG AAC A-3’; 5’-ATG AGG GAC TCA AGG GGC AAG TCC TGG A-3’. PCR products were

separated by agarose gel electrophoresis and imaged on a BioRad ChemiDoc XRS+ (BioRad,

Hercules, CA). EHD3-/- deficiency was confirmed by immunoblot using EHD3-specific

antibodies.

Page 14: Final Circ Res Publication

Echocardiography. Transthoracic echocardiogram was performed on WT and EHD3-/- mice to

measure the in vivo function of the heart using the Vevo 2100 (Visualsonics, Toronto, Ontario,

Canada). The mice were anesthetized using 2.0 % isoflurane in 95% O2 / 5% CO2 at a rate of ~

0.8 L/min. Anesthesia was maintained by administration of oxygen and ~1% isoflurane.

Electrode gel was placed on the EKG sensors of the heated platform and the mouse was placed

supine on the platform to monitor electrical activity of heart. A temperature probe was inserted

into the rectum of the mouse to monitor core temperature of ~ 37°C. The MS-400 transducer

was used to collect the contractile parameters of the heart in the short axis M-mode. Heart rates

were continuously monitored, and echocardiography data acquired at heart rates below 450

bpm was excluded.

Electrophysiology. Membrane currents were assessed by use of an Axopatch-200B amplifier

and a CV-203BU head stage (Axon Instruments). Experimental control, data acquisition, and

data analysis were accomplished with the use of software package PClamp 10 with the Digidata

1440A acquisition system (Axon Instruments). Patch pipettes were pulled from thin-walled glass

capillary tubes (Sutter Instruments). The electrode resistance ranged from 2 to 4 MΩ. The

external solution contained the following (in mmol/L): NaCl, 145; CsCl, 4; MgCl2, 1; CaCl2, 2;

HEPES, 5; glucose, 10 (pH 7.4, adjusting with NaOH). Ouabain (0.02 mmol/L), nifedipine (0.01

mmol/L) and niflumic acid (0.02 mmol/L) were added to the solution. The internal solution

contained the following (in mmol/L): CsCl, 65:NaCl, 10; MgCl2, 4; CaCl2, 6; tetraethyl ammonium

chloride, 20; HEPES, 10; Na2ATP,5; EGTA, 11 (pH 7.2, adjusting with CsOH). Membrane

currents were elicited with the use of standard voltage ramp protocol. From a holding potential

of -40 mV, a 100-ms step depolarization to +50 mV was followed by a descending voltage ramp

(from +50 mV to -110 mV at 100 mV/s). The protocol was applied every 12 seconds. INCX was

measured as the Ni-sensitive current. Ni2+ (5 mmol/L) was added to define the fraction of current

that derives from NCX (the difference between total current and post-Ni2+ current). Membrane

Page 15: Final Circ Res Publication

capacitance was read directly from the membrane test of Pclamp10 before compensating for

series resistance and membrane capacitance. For ICa,L measurements electrodes were

fashioned from borosilicate glass capillaries (World Precision Instruments) and were filled with

an internal solution containing: 150 cesium methane sulfonate (CsMeSO3), 5 CsCl, 10 HEPES,

10 EGTA, 1 MgCl2, and 4 MgATP (pH to 7.2 with NaOH). Pipettes typically had a resistance of

3-4 MΩ before series compensation. To form gigaohm seals and for initial break-in to the

whole-cell configuration, cells were perfused with normal Tyrode solution: 138 NaCl, 4 KCl, 2

CaCl2, 1 MgCl2, 0.33 NaH2PO4, 10 HEPES, 1.8 CaCl2, 0.5 MgCl2, 25 CsCl (pH adjusted to 7.4

with NMG). Data traces were acquired at a repetition interval of 2 s from -70 to +50 mV with a

holding potential of -80 mV.

SR Ca load and Ca spark measurements. Briefly, myocytes were loaded with Ca-dependent

fluorescent dye, fluo-4 AM (10 µM), for 30 min at RT. Myocytes were electrically field stimulated

at 0.5 Hz for at least 5 min before data acquisition to assure steady-state Ca handling.

Stimulation occurred in Tyrode solution (in mM): 140 NaCl, 4 KCl, 1 MgCl2, 10 glucose, 5

HEPES, 1 CaCl2 (pH adjusted to 7.4 with NaOH). A rapid solution switch to 0 Na/0 Ca NT

solution + 10 mM caffeine (140 LiCl substituted for NaCl) was applied for 2 s to empty the SR of

Ca. The difference between basal and peak cytosolic [Ca2+] in the presence of caffeine is

considered the total SR [Ca2+]. [Ca]SRT was calculated through the pseudo-ratio as previously

described3 with [Ca]d assumed to be 120 nM for all mice. through the pseudo-ratio as previously

described3 with [Ca]d assumed to be 120 nM for all mice. Ca sparks measurement were

acquired in intact ventricular myocytes loaded with 10 µM fluo-3 AM for 30 min. at RT. Imaging

was performed using an Olympus Fluoview 1000 confocal microscope equipped with a 60X 1.4

N.A. objective in line-scan mode at a rate of 2 or 5 ms per line. Fluo-3 was excited by 488-nm

line of an argon-ion laser, and fluorescence was acquired at wavelengths 500-530 nm. Ca

sparks were analyzed using IDL software as previously described.4

Page 16: Final Circ Res Publication

Surface Receptor Density. Surface membrane density assay of the β-AR was adapted from

Limas et al. and Yonemochi et al.5, 6 Isolated ventricular myocytes from both EHD3-/- and WT

mice were isolated as described. The resulting myocyte su spension was divided into two

equal aliquots were washed in ice cold assay solution (in mM: 137 NaCl; 5.4 KCl, 4 KCl; 0.16

NaH2PO4; 3 NaHCO3 20 HEPES, 10 glucose; supplemented with 650 mg taurine, pH 7.4). One

aliquot was used to assess β-AR specific signal; the second was used to assess non-specific

radioligand binding (background). To minimize the potential for surface receptor internalization,

myocytes were kept on ice for the duration of the assay. Intact myocytes were counted by

hemocytometer and a volume equal ~100,000 myocytes was added to a 5 mL BD Falcon tube.

Sample was volume up to 500 µL with assay buffer. Radiolabeled β-AR specific antagonist 3H-

CGP-12177 (Perkin Elmer) was then added to each tube at a final concentration of 15 nM.

Samples were incubated for 16 h at 4° C. To terminate radioligand binding, 4.5 mL of ice cold

assay buffer was added the samples. The samples were filtered over a Whatman GF/C filter

and washed three times with ice cold assay buffer. Radioligand binding was then quantified on

a scintillation counter. To account for non-specific binding of 3H-CGP-12177 to cell membranes,

simultaneous experiments were conducted with the addition of 10 µM propranolol, a general β-

AR agonist. The signal acquired in the presence of 3H-CGP-12177 and propranolol was

considered the general non-β-AR-specific signal (or background) and was subtracted from all

scintillation counts. All experiments for all treatments were conducted in triplicate and

averaged.

Page 17: Final Circ Res Publication

Supplemental References

1. Romero L, Pueyo E, Fink M, Rodriguez B. Impact of ionic current variability on human ventricular cellular electrophysiology. Am J Physiol Heart Circ Physiol. 2009;297:H1436-1445

2. Bondarenko VE, Szigeti GP, Bett GC, Kim SJ, Rasmusson RL. Computer model of action potential of mouse ventricular myocytes. Am J Physiol Heart Circ Physiol. 2004;287:H1378-1403

3. Curran J, Hinton MJ, Rios E, Bers DM, Shannon TR. Beta-adrenergic enhancement of sarcoplasmic reticulum calcium leak in cardiac myocytes is mediated by calcium/calmodulin-dependent protein kinase. Circulation research. 2007;100:391-398

4. Terentyev D, Viatchenko-Karpinski S, Gyorke I, Terentyeva R, Gyorke S. Protein phosphatases decrease sarcoplasmic reticulum calcium content by stimulating calcium release in cardiac myocytes. J Physiol. 2003;552:109-118

5. Limas CJ, Limas C. Rapid recovery of cardiac beta-adrenergic receptors after isoproterenol-induced "down"-regulation. Circ Res. 1984;55:524-531

6. Yonemochi H, Yasunaga S, Teshima Y, Takahashi N, Nakagawa M, Ito M, Saikawa T. Rapid electrical stimulation of contraction reduces the density of beta-adrenergic receptors and responsiveness of cultured neonatal rat cardiomyocytes. Possible involvement of microtubule disassembly secondary to mechanical stress. Circulation. 2000;101:2625-2630

Page 18: Final Circ Res Publication

Online Figure I

Online Figure I. Whole-cell recordings were obtained at RT with the use of standard patch-clamp techniques, and cells from the WT and EHD3-/- groups had a similar membrane capacitance .

Page 19: Final Circ Res Publication

Online Figure II

Online Figure II. EHD3-/- mice display normal PR and QT interval. Radiotelemetryin conscious, ambulatory mice reveal no differences in A) average PR interval (WT = 35.4 ± 0.5, EHD3-/- = 35.8 ± 0.4,) or B) average QRS complex duration (WT = 12.0 ±0.56, EHD3-/- = 10.9 ± 0.38 ms, n = 5 mice each).

Page 20: Final Circ Res Publication

Online Figure III. EHD3-/- mice display altered β-AR-dependent signaling. (A) Maximum heart rate achieved after low (left) or high dose (right) Iso injection in conscious, ambulatory WT and EHD3-/-

mice. (B) Duration of Iso response (time from peak HR to return to baseline) after low dose (left) or high dose (right) Iso injection in conscious, ambulatory mice. (C) Baseline heart rate in anesthetized mice before Iso injection. (D) Change in heart rate after Iso injection in anesthetized mice is blunted in EHD3-/- mice. For A-D, n=7 WT and n=8 EHD3-/- mice, p<0.05. (E-F) Both β1-AR and β2-AR protein expression are increased EHD3-/- mice (n=5 mouse myocyte preps/genotype; p<0.05). Representative immunoblots for E and F are shown in (G-H) (I) β1-AR surface receptor density was no different between WT and EHD3-/- mice (n=5 mouse myocyte preps/genotype). (J)Immunostaining of the β1-AR receptor in WT (left) and EHD3-/- myocytes (right). EHD3-/- myocytes display no overt difference in membrane protein localization, but show significant increase in peri-nuclear staining compared with WT myocytes.

Online Figure III

Page 21: Final Circ Res Publication

Online Figure IV. Sensitivity analysis to define likely candidates for decreased APD and increased SR Ca2+ load in EHD3-/- myocytes. Relative sensitivities of steady-state (A) APD90 and (B) [Ca2+]JSR to maximal conductances of sarcolemmal ion channels, transporters and pumps in a mathematical model of the mouse ventricular AP.The L-type Ca2+ channel (ICaL) and Na+/Ca2+ exchanger (INCX) had the greatest influence on APD90 and [Ca2+]JSR, respectively, in a manner consistent with observed defects (e.g. due to loss of membrane targeting would produce change in property consistent with observed defects in EHD-/-, designated by green bars). Abbreviations are as follows: ICaL - L-type Ca2+

channel; ICab – Background Ca2+ current; IK1 – Inward rectifier K+ current; IKss –Non-inactivating steady-state K+ current; IKur – Ultrarapid delayed rectifier K+

current; INa – Fast Na+ current; INab – Background Na+ current; INaK – Na+/K+

ATPase; INCX - Na+/Ca2+ exchanger; Ins – Nonspecific leak current; IpCa –Sarcolemmal Ca2+ pump; Ito – Transient outward K+ current.

Online Figure IV

Page 22: Final Circ Res Publication

Online Figure V. Summary data immunoblots assessing phospholamban, calsequestrin, and SERCA2a in whole heart lysates isolated from WT and EHD3-/- mice (n=5 each). Representative immunoblots showing protein of interest (top) and loading control (actin, bottom) are shown for each data set. Space between lanes denotes that data were collected from noncontiguous lanes of the same gel.

Online Figure V

Page 23: Final Circ Res Publication

Online Figure VI. Immunostaining of alpha-actinin and NCX in WT (A-C) and EHD3-/- (D-F) post-natal day 1 primary cardiomyocytes. Note peri-nuclear localization of NCX in EHD3-/- myocytes. Bar equals 10 microns.

Online Figure VI

Page 24: Final Circ Res Publication

Online Figure VII. Loss of ankyrin-B in EHD3-/- myocytes is rescued by viral transduction of GFP-EHD3. (A) Non-transduced EHD3-/- myocytes show lack of GFP signal as well as reduced ankyrin-B immunostaining. (B) GFP-EHD3 transduction results in peri-nuclear distribution of EHD3 as well as rescue of ankyrin-B expression. Bar equals 10 microns.

Online Figure VII

Page 25: Final Circ Res Publication

Online Figure VIII. EHD1 and EHD4 are differentially expressed in EHD3-/- mice. A) EHD1 is increased 30% in EHD3-/- hearts. B) EHD4 expression is decreased by 31% in EHD3-/-

hearts (n = 9 WT; 8 EHD3-/-, p < 0.05).

Online Figure VIII

Page 26: Final Circ Res Publication

Online Figure IX. EHD3 is expressed in sinoatrial node. A) EHD3 is expressed in mouse sinoatrial node cell lysates with SAN marker HCN4. B) EHD3 is expressed in canine sinoatrial node lysates with SAN marker HCN4. C) Control experiment to validate specificity of EHD3 antibody on EHD3-/- ventricular lysates.

Online Figure IX

Page 27: Final Circ Res Publication

Online Figure X. Sinus node dysfunction in 4 week old EHD3-/- mouse. SAN dysfunctionwas not observed in WT mice.

Online Figure X

Page 28: Final Circ Res Publication

Online Table I

ParameterWT hearts8 weeks

(n=5)

EHD3-/- hearts8 weeks

(n=6)

LV diastolic diameter (mm)

3.72 ± 0.09 4.05 ± 0.09*

LV systolic diameter(mm)

2.4 ± 0.11 2.95 ± 0.08*

LV diastolic volume (μL)

58.17 ± 3.52 72.60 ± 3.83*

LV systolic volume (µL)

20.51 ± 2.46 33.91 ± 2.28*

Anterior diastolic wallthickness (mm)

0.75 ± 0.035 0.74 ± 0.02

Anterior systolic wallthickness (mm)

1.19 ± 0.045 1.09 ± 0.02

Posterior diastolic wall thickness (mm)

0.65 ± 0.02 0.63 ± 0.02

Posterior systolic wall thickness (mm)

1.06 ± 0.04 0.89 ± 0.03*

Stroke Volume (μL) 38.47 ± 3.27 38.69 ± 1.89

Ejection Fraction 65.41 ± 2.14 53.59 ± 1.34*

Fractional Shortening 35.4 ± 1.23 27.34 ± 0.86*

Heart weigh/Body weight (mg/g)

6.79 ± 0.21 7.98 ± 0.26*

Online Table I. Echocardiographic parameters of WT and EHD3-/- mice. In all, seven WT and 6 EHD3-/- mice were assessed at 8 weeks of age. (*p<0.05 vs. WT).

Page 29: Final Circ Res Publication

ParameterWT hearts4 weeks

(n=5)

EHD3-/- hearts4 weeks

(n=5)

LV diastolic diameter (mm)

3.38 ± 0.03 3.76 ± 0.05*

LV systolic diameter(mm)

2.20 ± 0.05 2.61 ± 0.09*

LV diastolic volume (μL)

46.7 ± 1.04 60.7 ± 1.91*

LV systolic volume (µL)

16.35 ± 0.89 25.15 ± 2.26*

Anterior diastolic wallthickness (mm)

0.7 ± 0.03 0.75 ± 0.02

Anterior systolic wallthickness (mm)

1.03 ± 0.03 1.16 ± 0.02*

Posterior diastolic wall thickness (mm)

0.66 ± 0.04 0.64 ± 0.02

Posterior systolic wall thickness (mm)

0.95 ± 0.04 0.96 ± 0.032

Stroke Volume (μL) 30.35 ± 0.68 35.55 ± 0.99*

Ejection Fraction 65.08 ± 1.37 58.89 ± 2.73

Fractional Shortening 34.8 ± 1.02 30.8 ± 1.88

Online Table II

Online Table II. Echocardiographic parameters of WT and EHD3-/- mice assessed at 4 weeks of age. (*p<0.05 vs. WT).

Page 30: Final Circ Res Publication

ParameterWT hearts8 weeks

(n=5)

EHD3-/- MHC8 weeks

(n=5)

LV diastolic diameter (mm) 3.72 ± 0.09 4.46 ± 0.13*

LV systolic diameter (mm) 2.4 ± 0.11 3.4 ± 0.09*

LV diastolic volume (μL) 58.17 ± 3.52 90.90 ± 6.33*

LV systolic volume (µL)

20.51 ± 2.46 47.67 ± 2.98*

Anterior diastolic wallthickness (mm)

0.75 ± 0.035 0.9 ± 0.06

Anterior systolic wallthickness (mm)

1.19 ± 0.045 1.3 ± 0.02

Posterior diastolic wall thickness (mm)

0.65 ± 0.02 0.66 ± 0.03

Posterior systolic wall thickness (mm)

1.06 ± 0.04 0.91 ± 0.07

Stroke Volume (μL) 38.47 ± 3.27 43.23±4.96

Ejection Fraction 65.41 ± 2.14 47.39 ± 2.81*

Fractional Shortening 35.4 ± 1.23 23.7±1.71*

Heart weigh/Body weight (mg/g)

6.79 ± 0.21 8.22±0.21*

Online Table III

Online Table III. Echocardiographic parameters of WT and EHD3-/- cKO mice assessed at 8 weeks of age. (*p<0.05 vs. WT).


Recommended