Skip to main content
Erschienen in: BMC Cardiovascular Disorders 1/2014

Open Access 01.12.2014 | Research article

Resistance to cardiomyocyte hypertrophy in ae3 −/− mice, deficient in the AE3 Cl/HCO3 exchanger

verfasst von: Daniel Sowah, Brittany F Brown, Anita Quon, Bernardo V Alvarez, Joseph R Casey

Erschienen in: BMC Cardiovascular Disorders | Ausgabe 1/2014

Abstract

Background

Cardiac hypertrophy is central to the etiology of heart failure. Understanding the molecular pathways promoting cardiac hypertrophy may identify new targets for therapeutic intervention. Sodium-proton exchanger (NHE1) activity and expression levels in the heart are elevated in many models of hypertrophy through protein kinase C (PKC)/MAPK/ERK/p90RSK pathway stimulation. Sustained NHE1 activity, however, requires an acid-loading pathway. Evidence suggests that the Cl/HCO3 exchanger, AE3, provides this acid load. Here we explored the role of AE3 in the hypertrophic growth cascade of cardiomyocytes.

Methods

AE3-deficient (ae3 −/− ) mice were compared to wildtype (WT) littermates to examine the role of AE3 protein in the development of cardiomyocyte hypertrophy. Mouse hearts were assessed by echocardiography. As well, responses of cultured cardiomyocytes to hypertrophic stimuli were measured. pH regulation capacity of ae3 −/− and WT cardiomyocytes was assessed in cultured cells loaded with the pH-sensitive dye, BCECF-AM.

Results

ae3 −/− mice were indistinguishable from wild type (WT) mice in terms of cardiovascular performance. Stimulation of ae3 −/− cardiomyocytes with hypertrophic agonists did not increase cardiac growth or reactivate the fetal gene program. ae3 −/− mice are thus protected from pro-hypertrophic stimulation. Steady state intracellular pH (pHi) in ae3 −/− cardiomyocytes was not significantly different from WT, but the rate of recovery of pHi from imposed alkalosis was significantly slower in ae3 −/− cardiomyocytes.

Conclusions

These data reveal the importance of AE3-mediated Cl/HCO3 exchange in cardiovascular pH regulation and the development of cardiomyocyte hypertrophy. Pharmacological antagonism of AE3 is an attractive approach in the treatment of cardiac hypertrophy.
Begleitmaterial
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​1471-2261-14-89) contains supplementary material, which is available to authorized users.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

DS: First author of the manuscript and completed the majority of the experimental procedures. BFB: Assisted with manuscript preparation. AQ: Collected the qRT-PCR data and generated/optimized the respective primers. BVA: Collection of HW: BW data. JRC: Conceived and supervised the experiments and assisted with manuscript preparation. All authors read and approved the final manuscript.

Background

Cardiovascular diseases remain a major cause of death worldwide despite progress in disease outcomes of patients [1]. Heart failure (HF) is the common end-stage of many cardiovascular disorders, with a prevalence of 5.8 million in the USA and about 23 million worldwide [2, 3]. Annually, 550,000 new cases of HF arise in the U.S.A. The intricate molecular events resulting in heart failure remain incompletely understood, but enlargement of cardiac contractile cells (cardiomyocyte hypertrophy) in response to various stimuli is central to the progression to heart failure [4].
Cardiac cells are terminally differentiated cells that respond to increased stress by increasing their size rather than mitotically dividing to increase their number [5]. Cardiovascular events that increase myocardial stress (workload) chronically induce hypertrophic growth. Pressure overload, myocardial infarction, obesity, pregnancy or exercise can independently trigger molecular mechanisms culminating in increased cardiomyocyte size. Cardiac hypertrophy occurs to normalize the elevated demand on the myocardium, and can be physiological or pathological depending on the source of the initiating stimuli [6]. Physiological hypertrophy prevails in healthy individuals during increased physical activities or in pregnant women. Pathological hypertrophy results from prolonged elevated blood pressure (pressure overload), ischemia accompanied by changes in Ca++ handling, or genetic abnormalities. Initially, pathological hypertrophic growth compensates for the decline in contractile function, but ultimately the myocardium becomes decompensated from sustained exposure to the initiating stimuli. Understanding the distinct pathways mediating cardiac hypertrophic development has potential to identify new drug targets for the management of heart failure.
Intracellular pH (pHi) regulation is paramount in maintaining normal cardiac function [7, 8]. Plasma membrane transporters involved in maintaining pHi at physiological levels in the heart include the Na+/H+ exchanger (NHE1), Na+/HCO3 co-transporters (NBC), and Cl/HCO3 exchangers [9, 10]. Cytosolic acidification or hormonal stimulation activate NHE1, which facilitates electroneutral Na+/H+ exchange, to alkalinize the cytosol [11]. Accumulating evidence suggests that NHE1 expression level and activity increase in hypertrophy [12, 13]. In the hypertrophied myocardium of the spontaneously hypertensive rats (SHR), there was an increased activation of NHE1 [14] and NHE1 inhibition reduced cardiac hypertrophy and interstitial fibrosis [15]. Transgenic mice expressing activated NHE1 exchanger had enlargement of the heart and increased sensitivity to hypertrophic stimulation [16]. Since NHE1 activation induces acid extrusion, alkalinization should accompany NHE1 activation. NHE1 activation was not, however, accompanied by increased pHi, although cytosolic Na+ was elevated [14]. Moreover, under alkaline conditions, NHE1 activity is self-inhibited, which suggests that an acidifying mechanism running counter to NHE1 is necessary for sustained NHE1 activation [1720]. Indeed, Cl/HCO3 exchange mediated by AE3 provides this acidifying pathway [7, 8, 10].
The heart expresses three Cl/HCO3 exchanger isoforms: AE1, AE2 and AE3 [10, 21]. Another cardiac Cl/HCO3 , SLC26a6, [2224], may represent the Cl/OH exchanger (CHE) that has been reported in the heart [25]. Two AE3 variants, AE3 full length (AE3fl) and cardiac AE3 (AE3c) are expressed in the heart; AE3fl is also expressed in the brain and retina [2628]. Phenylephrine (PE) and angiotensin II (ANGII), acting on their G-protein-coupled receptors (GPCRs), activate AE3fl via protein kinase C (PKC). Interestingly, PKC can indirectly activate NHE1 via MAPK-dependent mechanisms [29]. Moreover, carbonic anhydrase II (CAII), another modulator of the PE-dependent hypertrophic growth, interacts with both NHE1 and AE3 to provide their respective transport substrates, H+ and HCO3 [30, 31].
CAII activation was recently found to be important in the induction of cardiomyocyte hypertrophy. In isolated rat cardiomyocytes, inhibition of CAII catalytic activity reduced phenylephrine (PE) and angiotensin II (ANGII) induced cardiomyocyte hypertrophy [32]. Additionally, infection of neonatal rat cardiomyocytes with adenoviral constructs encoding catalytically inactive CAII mutant, CAII-V143Y, reduced the response of the cardiomyocytes to hypertrophic stimuli, suggested to arise from a dominant negative mode of action [33]. Cardiomyocytes from CAII-deficient mice had physiological hypertrophy, but were unresponsive to hypertrophic stimulation [33]. Finally, expression of CAII and CAIV was elevated in the hypertrophic ventricles from failing human hearts, indicating that elevation of carbonic anhydrases is a feature of heart failure in people [34]. Taken together, these findings show that CAII plays a role in the development of cardiomyocyte hypertrophy.
Several reports revealed that CAII physically and functionally interacts with Cl/HCO3 anion exchangers to enhance the transport activity of anion exchangers forming a bicarbonate transport metabolon [31, 3537], although some reports have questioned the physiological relevance of this physical and functional linkage [3840]. CAII also interacts physically and functionally with NHE1 to increase the exchange activity [30, 41]. These observations suggest that simultaneous activation of AE3, CAII and NHE1 occurs upon pro-hypertrophic stimulation by the PKC-coupled agonists, PE, ANGII or endothelin I (ET-I). This pathological activated complex has been termed the hypertrophic transport metabolon (HTM) [34].
Accumulating evidence suggests a significant role of AE3 in cardiac function. AE3 Cl/HCO3 exchange activity is involved in cardiac contractility by altering cardiac Ca++ handling [42]. Moreover, disruption of the ae3 gene in mice resulted in an exacerbated cardiac function and precipitated heart failure in hypertrophic cardiomyopathy mice [43]. Pacing of AE3 null hearts abrogated frequency-dependent inotropy, which, suggests that AE3 is required in mediating force-frequency response induced by acute biochemical stress [44]. Taken together, these findings suggest that the AE3 Cl/HCO3 exchanger is critical in heart growth and function but the exact mechanism remains unknown.
In the present study, we examined the role of AE3 in cardiomyocyte hypertrophy, using AE3-deficient (ae3 −/− ) mice. Cardiac growth parameters and fetal gene reactivation were measured in the presence of pro-hypertrophic stimulation in cardiomyocytes from ae3 −/− mice. We also examined the role of AE3 in cardiomyocyte steady state pHi, using the ae3 −/− mice. Our results indicate that ae3 deletion prevents cardiomyocyte hypertrophy and reduces the rate of pHi recovery in cardiomyocytes, reinforcing the importance of AE3 in cardiovascular pH regulation and the development of cardiomyocyte hypertrophy.

Methods

Animal care and use

All procedures involving animals were performed in accordance with the guidelines established by the Canadian Council on Animal Care and the University of Alberta Animal Care and Use Committee.

ae3 null mice

Experiments were performed using ae3 null mice in a C57BL/6 background. The ae3 null strain has been previously described and characterized [42]. Age-matched WT mice from separate breedings were used as controls.

Heart weight to body weight ratio

Mice were weighed and anesthetized with sodium pentobarbital (50 mg/kg) by intraperitoneal injection. Upon reaching surgical plane, hearts were removed after performing midsection thoracotomy and rinsed in 4°C phosphate buffer saline (PBS: 140 mM NaCl, 3 mM KCl, 6.5 mM Na2HPO4, 1.5 mM KH2PO4, pH 7.4). Ventricles were separated from atria and blood vessels, blotted dry and the ventricular weight was measured. Heart weight to body weight ratio (HW/BW) was then calculated by dividing the weight of the ventricles by the weight of the whole animal.

Hematoxylin-eosin staining of heart sections

Hematoxylin and eosin (HE) staining was performed on longitudinal and transverse sections of wildtype and ae3 null adult mouse heart, using previously described protocols [45, 46]. Briefly, paraffin-embedded hearts were sectioned into 3 μm slices, which were trimmed and floated onto a water bath at 42°C, containing 50 mg/l of gelatin while gently stretching the cut sections to avoid wrinkles. Poly-L-lysine coated microscope slides were dipped under the meniscus of the water bath and a tissue slice was carefully mounted onto it. Sections were then air-dried for 16 h at 20°C, after which the slides were placed on edge in an oven and baked for 15 min at 60°C. Sections were deparaffinized by successively immersing them for 5 min with agitation in xylene, 100% ethanol and 70% ethanol, and rehydrated in Tris-EDTA buffer (1 mM EDTA, 0.05% Tween 20, 10 mM Tris, pH 9.0) for 1 min. Slides were rinsed in distilled water for 1 min with agitation. Slides were agitated for 30 s in Mayer’s hematoxylin solution (1.0 g/l hematoxylin (Sigma), sodium iodate (0.2 g/l), aluminum ammonium sulfate · 12 H2O (50 g/l), chloral hydrate (50 g/l) and citric acid (1 g/l) and rinsed in water for 1 min. Slides were stained in 1% eosin Y solution (1% eosin Y aqueous solution, Fisher) for 30 s with agitation. Sections were dehydrated by successively immersing it twice in 95% ethanol and twice in 100% ethanol for 30 s each. Ethanol was extracted twice in xylene, followed by addition of two drops of mounting medium (Canada Balsam, Sigma), after which the sections were covered with a coverslip. Images of transverse and longitudinal sections were collected, using a Nikon digital camera (DXM 200) mounted on top of a Nikon Eclipse E600 microscope.

Echocardiography

Echocardiographic assessment of cardiac performance in male WT and ae3 −/− mice was performed by the Cardiovascular Research Centre Core Facility (University of Alberta). Mice were subjected to mild anesthesia by isoflurane inhalation and echocardiography parameters were measured using a Vevo 770 High-Resolution Imaging System with a 30-MHz transducer (RMV-707B; Visual Sonics, Toronto). M-mode images and a four chamber view allowed for the calculation of wall measurements, ejection fraction, fractional shortening, and mitral velocities (E and A). Mitral valve tissue motion (E’) was measured, by tissue Doppler echocardiography of the mitral septal annulus.

Blood pressure measurements

Non-invasive blood pressure measurements were performed by the Cardiovascular Research Centre Core Facility (University of Alberta). Mice were comfortably restrained in a 26°C warming chamber (IITC Life Science) for ~15 min prior to taking blood-pressure (BP) measurements. Tail-cuff sensors were secured on the tail to occlude the blood flow, and connected to the recording device. Systolic and diastolic pressure, heart rate, blood volume and flow were obtained, using CODA6 software (Kent Scientific Corporation, Connecticut, USA).

Isolation and culture of adult mouse cardiomyocytes

Cardiomyocytes from adult male mouse hearts were isolated and cultured with modifications of published protocols [21, 47]. Briefly, adult mice were euthanized with sodium pentobarbital (50 mg/kg body mass) by intraperitoneal injection. Upon reaching surgical plane, midsection thoracotomy was performed and hearts were quickly excised and placed in 4°C perfusion solution, containing in mM: 120 NaCl, 5.4 KCl, 1.2 MgSO4, 5.6 glucose, 10 2,3-butanedione monoxime (BDM) (Sigma), 5 taurine (Sigma), 1.2 NaH2PO4, 10 HEPES, pH 7.4. Extra-cardiac tissues were removed and hearts were subjected to retrograde perfusion with perfusion solution to remove excess blood. Perfusion was switched for 15 min to perfusion solution at 37°C, supplemented with 0.5 mg/ml collagenase type B (Roche), 0.5 mg/ml collagenase type D (Roche), 0.02 mg/ml protease XIV (Roche) and 50 μM CaCl2. Ventricles, partially digested at this stage, were removed, cut into several pieces, and digested further in the same enzyme digestion solution by gentle trituration with a transfer pipette. Once the ventricles were completely digested, enzymatic digestion was terminated by addition of Digestion stop buffer I (perfusion solution, containing 10% (v/v) fetal bovine serum (FBS) (GIBCO) and 50 μM CaCl2). Lysates settled under gravity for 10 min at room temperature and pellets were resuspended in myocyte stopping buffer II (perfusion solution, containing 5% (v/v) FBS and 50 μM CaCl2). Samples were transferred to 60 mm tissue culture dish and calcium levels were increased by addition of CaCl2 to obtain final concentrations of 62, 112, 212, 500 and 1000 μM, sequentially at 4 min intervals at 20°C. Cells were transferred to a 14 ml culture tube and allowed to sediment for 10 min under gravity. Cells were resuspended in myocyte culture medium (Dulbecco’s Modified Eagle’s Medium/Nutrient Mixture F12-Ham (Sigma), supplemented with 10 mM BDM, 5% (v/v) FBS, 1% penicillin (GIBCO), 10 mM BDM, and 2 mM L-glutamine (GIBCO)). Myocytes were plated at a density of (0.5-1) x 104 cells/cm2 onto 35 mm culture dishes, pre-coated for 2 h with 10 μg/ml mouse laminin (Invitrogen) in PBS. Cells were incubated at 37°C in a 5% CO2 incubator for 1 h at which point medium was replaced with Dulbecco’s Modified Eagle’s Medium/Nutrient Mixture F12-Ham, containing 10 mM BDM, 1% penicillin, 2 mM L-glutamine, 0.1 mg/ml bovine serum albumin, and 1x ITS Liquid Media Supplement (Sigma). The entire culture procedure was carried out in a sterile laminar flow hood.

Assessment of cardiomyocyte hypertrophic growth

Cardiomyocytes were isolated and cultured from adult mouse hearts as described above. Following 18 h of culture, myocytes were treated with solvent carrier (control), 10 μM phenylephrine (Sigma) or 1 μM angiotensin II (Sigma) for another 24 h. Hypertrophy was assessed by analysis of cell surface area of cardiomyocytes pre- and post-treatment with the hypertrophic agonists. Images of characteristic rod-shaped cardiomyocytes were collected with a QICAM fast-cooled 12-bit colour camera (QImaging Corporation). Cell surface areas were measured, using Image-Pro Plus software (Media Cybernetics). Each treatment group contained 100–200 cells of ~ ten different experiments. Cell surface area (% relative to control) = Surface area (post-treatment)/Surface area (pre-treatment) X 100.

Real-time quantitative reverse transcription PCR (qRT-PCR)

Cardiomyocytes were prepared and treated as above. At the end of the culture period, medium was aspirated and cardiomyocytes were harvested in 350 μl Buffer RLT (Qiagen). RNA was extracted from the lysates with an RNeasy Plus Mini Kit, as per manufacturer’s instructions (Qiagen). RNA samples (100 ng) were reverse transcribed, following the manufacturer’s instructions for SuperScript II™ reverse transcriptase (Invitrogen). qRT-PCR was performed in a Rotorgene 3000 real time thermal cycler (Corbett Research), using a reaction mix containing: 5 μl template cDNA, 12 μl 2x Rotor-Gene SYBR Green PCR Master Mix (Rotor-Gene SYBR Green PCR Kit, Qiagen), and 1 μM of each primer. Data were obtained and analyzed, using Rotor Gene 6.0.14 software. Cycle threshold (Ct) values were obtained for carbonic anhydrase II (CAII), NHE1, atrial natriuretic peptide (ANP), β-MHC and glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Primers (Table 1) were designed, using Primer3 (http://​Frodo.​wi.​mit.​edu/​primer3/​).
Table 1
Sequences of primers used in qRT-PCR
Target gene
Primer sequence
Anp
F:
5′-TCCAGGCCATATTGGAGCAAATCC-3′
R:
5′-TCCAGGTGGTCTAGCAGGTTCTTG-3′
β-mhc
F:
5′-GAGACGGAGAATGGCAAGAC-3′
R:
5′-AAGCGTAGCGCTCCTTGAG-3′
Caii
F:
5′-CTCTGCTGGAATGTGTGACCT-3′
R:
5′-GCGTACGGAAATGAGACATCTGC-3′
Nhe1
F:
5′-TTTTCACCGTCTTTGTGCAG-3′
R:
5′-TGTGTGGATCTCCTCGTTGA-3′
Gapdh
F:
5′-CCTCGTCCCGTAGACAAAAT-3′
 
R:
5′-TGATGGCAACAATCTCCACT-3′
Anp, atrial natriuretic peptide; β-mhc , beta myosin heavy chain; Caii, carbonic anhydrase II; F, Forward; Nhe1, sodium proton exchanger isoform 1; Gapdh, glyceraldehyde 3-phosphate dehydrogenase; R, Reverse.

Immunoblotting

Cardiomyocytes, isolated and cultured from adult mouse hearts, were subjected to drug intervention as described above. Twenty-four h following treatment, medium was aspirated and myocytes were washed with 4°C PBS. Cells were lysed with SDS-PAGE sample buffer (10% (v/v) glycerol, 2% (w/v) SDS, 2% 2-mercaptoethanol, 0.001% (w/v) bromophenol blue, 65 mM Tris, protease inhibitors (1 μg/ml) pH 6.8) and lysates were heated 5 min at 65°C. Protein concentrations were determined by the bicinchoninic acid (Pierce Biotechnology) assay [48], and 20 μg of protein was resolved by SDS-PAGE on 10% acrylamide gels. Proteins were transferred onto PVDF membranes by electrophoresis for 1 h at 100 V in transfer buffer (10% (v/v) methanol, 25 mM Tris, and 192 mM glycine). PVDF membranes were blocked for 30 min with 5% (w/v) nonfat dry milk/ 0.1% (v/v) Tween 20 in TBS (137 mM NaCl, 20 mM Tris, pH 7.5). Immunoblots were incubated with rabbit polyclonal anti-CAII antibody (Santa Cruz Biotechnology; 1:1000), rabbit anti-human SLC26a6 (1:1000) [49], or rabbit polyclonal anti-NHE1 antibody (1:1000) [32] in TBST-M for 16 h at 4°C. Immunoblots were washed with TBST (TBS, containing 0.1% (v/v) Tween 20) and incubated with donkey anti-rabbit IgG conjugated to horseradish peroxidase (Santa Cruz Biotechnology; 1:2000) or mouse anti-goat IgG conjugated to horseradish peroxidase (Santa Cruz Biotechnology; 1:2000) for 1 h at room temperature. Immunoblots were washed in TBST and visualized, using ECL reagents (Perkin Elmer) and a Kodak Imaging Station 440CF (Kodak, Rochester, NY). Proteins were quantified by densitometry, using Kodak Molecular Imaging software (version 4.0.3; Kodak). Immunoblots were stripped by incubating in 10 ml of stripping buffer (2% (w/v) SDS, 10 mM 2-mercaptoethanol, 62.5 mM Tris, pH 6.8) at 50°C for 10 min with occasional shaking, followed by three washes with TBST. Membranes were incubated with mouse monoclonal anti-β-actin antibody (Santa Cruz Biotechnology; 1:2000) for 1 h at 20°C, washed with TBST, and incubated with sheep anti-mouse IgG conjugated to horseradish peroxidase (Santa Cruz Biotechnology; 1:3000) for 1 h. Immunoblots were washed and visualized again, as described above.

Protein synthesis assays

Cardiomyocytes were prepared and subjected to hypertrophic stimulation as above. Radiolabeled phenylalanine ([3H]-Phe, 1 μCi/ml, (Perkin Elmer)) was added immediately after drug intervention and cells were incubated for another 24 h. Proteins were precipitated, using trichloroacetic acid (TCA) as described previously with some modifications [50]. Medium was carefully aspirated and 500 μl of 0.5% (v/v) Triton X-100, containing protease inhibitor cocktail (Roche) were added. Lysates were transferred into 1.5 ml microcentrifuge tubes and TCA (100%: 500 g in 350 ml H2O) was added to each tube to a final concentration of 40% (v/v). Samples were incubated at 4°C for 30 min, after which proteins were sedimented by centrifugation at 12 682 x g for 15 min at 4°C. Pellets were resuspended by adding 200 μl acetone (−20°C) and sedimented again by centrifugation, as above, for 10 min. The acetone wash was repeated one more time, and the pellets air dried for 20 min at room temperature. Pellets were resuspended in 200 μl of 0.2 M NaOH, 1% (w/v) SDS. Scintillation fluid (Perkin Elmer, 3.5 ml) was added to each sample and the radioactivity of [3H]-Phe was counted in a Beckman LS6500 liquid scintillation counter.

Measurement of pHiin adult mouse cardiomyocytes

The protocol was as described previously with minor modifications [51, 52]. Briefly, cardiomyocytes were isolated as described above and cultured on laminin-coated glass coverslips. Approximately 2 h later, cells were loaded with 2 μM BCECF-AM (Sigma-Aldrich, Canada) for 30 min at 37°C. Coverslips were placed in an Attofluor cell chamber (Invitrogen, Canada), then transferred onto the stage of a Leica DMIRB microscope. Perfusion with HCO3 Ringer’s buffer solution (in mM: 128.3 NaCl, 4.7 KCl, 1.35 CaCl2, 20.23 NaHCO3, 1.05 MgSO4 and 11 glucose, pH 7.4) was initiated at 3.5 ml/min. Solutions were bubbled with 5% CO2-balanced air. Intracellular alkalosis was induced by switching to a HCO3 Ringer’s solution, containing, 20 mM trimethylamine (TMA) (Sigma) and perfusion was continued for 3 min. Perfusion was switched back to the HCO3 Ringer’s buffer solution. pHi of individual cardiomyocytes was measured by photometry at excitation wavelengths of 502.5 nm and 440 nm with a Photon Technologies International (PTI, Lawrenceville, NJ, USA) Deltascan monochromator. Emission wavelength, 528.7 nm, was selected, using a dichroic mirror and narrow range filter (Chroma Technology Corp., Rockingham, VT, USA) and was measured with a PTI D104 photometer. At the end of experiments, pHi was clamped by the high K+/Nigericin technique [53] in calibration solutions containing, 140 mM KCl, 1 mM MgCl2, 2 mM EGTA, 11 mM glucose, 20 mM BDM, 10 mM HEPES. Three pH standards spanned a range of 6.5-7.5. Steady-state pHi was measured from the pHi value prior to induction of alkalosis. Rate of pHi recovery was measured by linear regression from first min of recovery from imposed alkalosis.

Statistical analysis

Data are expressed as mean ± S.E.M. Statistical analyses were performed using paired t-tests or ANOVA where appropriate. P < 0.05 was considered significant.

Results

Cardiac development in ae3 −/− mice

To examine the role of AE3 in heart development, we characterized age-matched wildtype (WT) and ae3 −/− mice (~three months old). Breeding yielded Mendelian ratios of litters of both WT and ae3 −/− mice, indicating that loss of the AE3 does not affect prenatal survival. While the body weights of age-matched WT and ae3 −/− mice were not significantly different (Figure 1A), the difference in heart weights between the groups was statistically significant (Figure 1B). As a result, the HW/BW ratio, an index of cardiac hypertrophy, was reduced in the ae3 −/− mice relative to their WT counterparts (Figure 1C and D), suggesting that AE3 has a role in heart growth. Despite the observed apparent smaller size of the ae3 null hearts, gross morphology of hearts stained with hematoxylin/eosin transverse and longitudinal sections revealed no differences in the wall size and the chamber diameter of WT (left panel) and ae3 null (right panel) mouse hearts (Figure 2B-C).

Blood pressure and echocardiography

Cardiovascular performance of age-matched WT and ae3 −/− mice was assessed by echocardiography. No major variations in cardiovascular functional parameters between WT and ae3 −/− mice (~three months old males) were observed, except for a significant decrease in the mitral value E/A ratio in ae3 −/− mice (Table 2). While this would suggest that more blood is entering the ventricle during the atrial systolic phase than during ventricular relaxation, other parameters of diastolic cardiac function (E/E’ ratio, IVRT) were unaffected. ae3 −/− mice therefore likely do not exhibit diastolic dysfunction. Systemic blood pressure measurements were also performed using the non-evasive tail cuffing approach. No significant difference in the systemic blood pressure of WT and ae3 −/− mice was observed (Table 2). Overall, these observations suggest that loss of AE3 does not affect cardiovascular performance under basal conditions, consistent with previous findings [42, 44].
Table 2
Echocardiographic and blood pressure analysis of WT and ae3 −/− mice
Parameter
WT
ae3 −/−
Ejection fraction, %
65 ± 3
67 ± 7
Fractional shortening, %
35 ± 3
37 ± 7
IVSd, mm
0.70 ± 0.01
0.75 ± 0.04
LVIDd, mm
3.5 ± 0.1
3.4 ± 0.1
LVPWd, mm
0.67 ± 0.03
0.70 ± 0.03
IVSs, mm
1.08 ± 0.03
1.11 ± 0.09
LVIDs, mm
2.3 ± 0.1
2.1 ± 0.3
LVPWs, mm
1.05 ± 0.07
1.05 ± 0.12
MV E Velocity, mm.s−1
730 ± 50
710 ± 50
MV A Velocity, mm.s−1
390 ± 30
520 ± 40
MV E/A ratio
1.9 ± 0.1
1.4 ± 0.1
E/E’
24 ± 3
21 ± 4
IVRT, ms
16 ± 0
17 ± 2
TEI index
0.63 ± 0.03
0.61 ± 0.02
Systolic pressure, mm Hg
102 ± 7
100 ± 9
Diastolic pressure, mm Hg
66 ± 6
64 ± 8
Flow, ml.min−1
11 ± 1
10 ± 1
Volume of blood, ml
36 ± 6
31 ± 5
Heart rate, beats.min−1
699 ± 1
721 ± 49
All experiments were performed on male mice. Values are expressed as means ± S.E.M. (n = 3 per group for echocardiography; n = 7 for tail cuff parameters, with three replicates over three separate days (bottom five rows of the table)); P < 0.05. IVSd, diastolic intraventricular septal wall thickness; LVIDd, diastolic left ventricular internal diameter; LVPWd, diastolic left ventricular posterior wall thickness; IVSs, systolic intraventricular septal wall thickness; LVIDs, systolic left ventricular internal diameter; LVPWs, systolic left ventricular posterior wall thickness; MV E velocity, peak E wave mitral valve velocity; MV A velocity, peak A wave mitral valve velocity; MV E/A, ratio of peak E wave mitral valve velocity to peak A wave mitral valve velocity; E/E’, ratio of peak E wave mitral valve velocity to E wave mitral valve tissue motion; IVRT, isovolumic relaxation time.

Cardiomyocyte growth upon pro-hypertrophic stimulation

Cardiomyocyte hypertrophy is characterized by an increase in cardiomyocyte surface area, resulting in an overall increase in heart size. Cardiomyocytes were isolated from adult WT and ae3 −/− mice and the cell surface assessed by morphometry. The cell surface area of ae3 null cardiomyocytes was 20% ± 4% (n = 6) lower than WT (Figure 3). To determine the response of cardiomyocytes to pro-hypertrophic stimulation, adult cardiomyocytes were cultured and treated with PE and ANGII 18 h later. Cell surface area was measured 24 h following treatment with hypertrophic agonists. PE and ANGII induced a 20–25 ± 4% (n = 10) increase in the cell surface area of WT cardiomyocytes, but cardiomyocytes from ae3 −/− hearts were not susceptible to pro-hypertrophic stimulation by these agents (Figure 4). This suggests that AE3 has a role in the hypertrophic signaling pathway downstream of PE and ANGII.

Expression of hypertrophic marker genes

Cardiac hypertrophic development is associated with increased expression of marker genes, including ANP [54], β-myosin heavy chain (β-MHC) [55] and α-skeletal actin [56]. mRNA and protein levels of these markers are elevated in hypertrophic hearts [57]. Expression levels of ANP and β-MHC, were assessed by qRT-PCR in cardiomyocytes subjected to pro-hypertrophic stimulation. Transcript abundance of ANP and β-MHC were not significantly different between untreated cardiomyocytes from WT and ae3 −/− mice (Figure 5A and B). Stimulation with PE and ANGII, however, led to 8–10 fold increase of ANP and β-MHC expression levels in cardiomyocytes from WT mice. In contrast, ANP and β-MHC increased only 1–2 fold in cardiomyocytes from ae3 −/− mice. These data suggest that pro-hypertrophic stimulation induces a much greater upregulation of hypertrophic marker genes in cardiomyocytes from WT mice than ae3 −/− cardiomyocytes.

Expression of HTM genes in mouse cardiomyocytes

AE3 is implicated as part of a functional complex with NHE1 and CAII, referred to as the hypertrophic transport metabolon (HTM), whose activation has been proposed to induce cardiac hypertrophy [32, 33]. To assess the possibility of functional compensation for loss of AE3 by altered expression of these partner proteins, we examined mRNA expression of NHE1 and CAII by qRT-PCR. Baseline expression level of NHE1 in ae3 −/− mice was significantly elevated compared to the WT mice (Figure 6A). Pro-hypertrophic stimulation with PE or ANGII did not influence the NHE1 transcript abundance in the WT or ae3 −/− cardiomyocytes (Figure 6A). CAII transcript abundance, however, was higher in cardiomyocytes from ae3 −/− mice than from WT mice (Figure 6B). ae3 ablation thus induces a compensatory increase in CAII expression. Interestingly, pro-hypertrophic stimulation markedly increased CAII transcript levels in both WT and ae3 −/− cardiomyocytes (Figure 6B).
Since the baseline expression level of CAII was elevated in ae3 −/− cardiomyocytes and was further enhanced by hypertrophic stimulants in WT cardiomyocytes, we evaluated CAII protein expression. Cardiomyocytes isolated from ae3 −/− and WT male adult mice hearts were probed for CAII on immunoblots (Additional file 1: Figure S1A). CAII migrated at ~27 kDa, consistent with the expected molecular weight of CAII. Both the steady-state level of CAII protein (Additional file 1: Figure S1B) and mRNA level for CAII (Figure 6B) were higher in ae3 −/− cardiomyocytes than WT cardiomyocytes. Importantly, however, there was a large quantitative difference in the response, where CAII protein level rose by about 50%, whereas CAII message rose was about eight-fold higher in the ae3 −/− mice than WT. PE and ANGII increased CAII levels in the WT cardiomyocytes. Contrastingly, CAII levels in ae3 −/− cardiomyocytes were not significantly affected by pro-hypertrophic stimulation (Additional file 1: Figure S1B).

Protein synthesis in pro-hypertrophically-stimulated cardiomyocytes

Cardiac hypertrophy is also characterized by an increase in protein synthesis to accommodate cardiomyocyte enlargement [58]. Protein synthesis was measured by determining the amount of radioactive [3H]-Phe incorporated into proteins in cardiomyocytes, isolated and cultured as described above and treated with PE and ANGII. [3H]-Phe incorporation into proteins in the presence of pro-hypertrophic stimuli was increased in cardiomyocytes from WT, but not ae3 −/− mice. We conclude that ae3 −/− mice do not respond to pro-hypertrophic agonists with increased protein synthesis (Figure 7).

pHiregulation in mouse cardiomyocytes

AE3 contributes to cardiomyocyte pHi regulation [5961], but there are no AE3-specific inhibitors that would enable delineation of the role of AE3 in pHi regulation. We thus examined the rate of recovery of pHi from an imposed intracellular alkalinization in freshly isolated cardiomyocytes. Cardiomyocytes were cultured on laminin-coated glass coverslips for 2 h, then incubated with the pH-sensitive fluorescence dye, BCECF-AM. Cells were perfused with a HCO3 -containing Ringer’s buffer until steady-state pHi was reached and the perfusion was switched to the HCO3 -containing Ringer’s buffer, containing TMA. The presence of TMA induced a rapid intracellular alkalinization (Figure 8A) until a new steady-state pHi was reached. pHi regulatory transporters spontaneously restored cell pHi. Steady-state pHi was measured as the pHi 60 s prior to switching to the TMA-supplemented HCO3 -containing Ringer’s buffer. The rate of pHi recovery from imposed intracellular alkalinization was measured by linear regression of the initial minute of pHi recovery. Baseline pHi was identical in cardiomyocytes from WT and ae3 −/− mice (Figure 8B). The mean peak pH following alkalization for WT group was 7.65 ± 0.05 and 7.60 ± 0.06 for ae3 −/− mice. Thus, under basal physiological conditions, loss of AE3 does not significantly affect the steady-state pHi in cardiomyocytes. The rate of recovery of pHi from alkalosis was, however, significantly slower in ae3 −/− cardiomyocytes than WT (Figure 8C).

Expression of pHi regulators in ae3 −/− cardiomyocytes

We next assessed the expression level of the other pHi regulatory transporters at the protein level. On immunoblots the proteins migrated at the expected sizes, NHE1 at ~100 kDa, and SLC26a6 at ~80 kDa (Additional file 2: Figure S2 and Additional file 3: Figure S3). Consistent with a previous study [44], expression of NHE1 was elevated in ae3 −/− cardiomyocytes compared to WT, but remained unchanged in the heterozygotes (Additional file 2: Figure S2A-B). Expression of SLC26a6 protein was not affected by deletion of the ae3 gene (Additional file 3: Figure S3A-B).

Discussion

Pathological cardiac hypertrophy renders the heart susceptible to cardiac failure. Accumulating evidence implicates NHE1 as a key candidate mediating pathological hypertrophy. Prolonged NHE1 activation produces intracellular alkalinization. Sustained NHE1 activation can only occur in the presence of a counter acidifying mechanism. The present study examined the possibility that the Cl/HCO3 anion exchange mediated by AE3 is responsible for the acidification mechanism. Our studies, using AE3 deficient mice, support a role for AE3 in cardiovascular pH regulation and the development of hormonally-induced cardiomyocyte hypertrophy. Pharmacological antagonism of AE3 is thus a possible therapeutic direction in the prevention of maladaptive cardiac hypertrophy.

Role of AE3 in cardiomyocyte hypertrophy

The role of AE3 in cardiac physiology is incompletely characterized, but several lines of evidence suggest that AE3 Cl/HCO3 exchange is required to maintain pHi homeostasis [10, 61]. Consistent with this, we found that the rate of recovery from an alkaline load was reduced in ae3 −/− cardiomyocytes, relative to WT.
The hypertrophic transport metabolon is a proposed pathological pathway in which AE3, NHE1 and CAII are coordinately activated and promote hypertrophic growth [32, 33]. Specifically, pro-hypertrophic agonists, including PE, ANGII and endothelin are coupled to PKC activation. NHE1 and AE3 are both activated by agonists coupled to PKC activation [59, 62, 63]. Co-activation of these respective alkalinizing and acidifying transporters has the net effect of loading the cell with NaCl, with no change of cytosolic pH. Elevated cytosolic Na+ in turn reduces the efficacy of Ca++-efflux by the plasma membrane Na+/Ca++ exchanger, resulting in a rise in cytosolic Ca++. Ca++ is a pro-hypertrophic second messenger [64, 65] and also may promote hypertrophy in a feed-forward cascade by stimulation of PKC [6]. Accumulating evidence suggests that both NHE1 and AE3 activities are influenced by physical and functional interactions with CAII [32, 41], which provides substrates for these transporters. That said, very recent data suggest that intracellular carbonic anhydrase does not activate cardiomyocyte Na+/H+ exchange [66]. Prevention of PE-induced cardiomyocyte hypertrophy upon CAII blockade [32, 33] may thus arise from reduced AE3-mediated cytosolic acidification, thereby decreasing driving force for NHE-mediated alkalinization. We propose that CAII, NHE1 and AE3 form a hypertrophic transport metabolon, where hypertrophy is promoted by the pathological activation of AE3 and NHE1, stimulated by interactions with CAII.
The functional relationship between AE3, CAII and NHE1 was further supported by our analysis of protein expression in ae3 −/− mice. Increased CAII transcript abundance and protein expression in ae3 −/− mice compared to WT mice suggest that there is compensation for a loss of AE3. This finding parallels results in retinal tissue from ae3 −/− mice, where there was increased CAII expression [67]. Functional complementarity of AE3 and CAII is further supported by the significant increase in AE3 transcript abundance in Car2 (caii −/− ) mice, compared to WT mice [33]. The upregulation of NHE1 transcript abundance and increased protein expression provide further support for the HTM. Taken together, these data support a functional interaction between AE3 and CAII, where there is compensation of one for the loss of the other.

Loss of AE3 prevents cardiomyocyte hypertrophy

This study lends support to the idea that AE3 is the Cl/HCO3 exchanger isoform working in conjunction with NHE1 to promote cardiomyocyte hypertrophy. Non-specific inhibition of Cl/HCO3 exchangers, using stilbene derivatives, prevented hypoxia-induced acidification in rat ventricular myocytes, as well as increases in intracellular Cl and Ca++ concentrations [68, 69], suggesting a role of Cl/HCO3 exchangers in cardiac pathology. When subjected to ischemia and reperfusion, hearts isolated from ae3 −/− mice revealed no effect on cardiac performance demonstrated as contractility, ventricular developed pressure or end diastolic pressure relative to wildtype [42]. Double knock-out ae3/nkcc1 (Na+-K+-2Cl co-transporter) mice, however, had elevated ischemia/reperfusion injury, which resulted in impaired cardiac contractility and overall cardiac performance [42]. These findings were attributed to impaired Ca++ handling in the double knock-out cardiomyocytes [42] compared to the single mutants. In a hypertrophic cardiomyopathy mouse model carrying a Glu180Gly mutation in α-tropomyosin (TM180), disruption of ae3 did not prevent or reverse the hypertrophic phenotype [43]. The TM180/ae3 double knockout mice had reduced cardiac function and compromised Ca++ regulation, which accounted for the rapid decline to heart failure [43].
Taken together, these two studies suggest that AE3 loss is not cardioprotective, which contrasts with findings of the present study, which found that loss of AE3 renders cardiomyocytes less susceptible to pro-hypertrophic stimulation. Our data showed that the marked rise in cell surface area, protein synthesis, and fetal gene reactivation observed in response to hypertrophic stimulation in cardiomyocytes from WT mice was not present in ae3 −/− mice. In the context of cardiomyocyte hypertrophy mediated by hormonal stimuli, our data demonstrate that ablation of AE3 affords protection against cardiomyocyte hypertrophy.
The discrepancy could arise from the model of hypertrophy and the cardiac pathology being investigated. Hypertrophic cardiomyopathy is a genetic disorder, which occurs as a result of mutations in the genes that encode cardiac contractile proteins [70]. This anomaly manifests as sudden cardiac death, arrhythmias, hypertrophy and heart failure [70]. Overall, hypertrophic cardiomyopathy results in impairment of Ca++ sensitivity by the myofibrils. In the present study however, we employed a model of hypertrophy induced by PE or ANGII, which involves interaction of these ligands with their cell surface receptors, GPCR [71]. The resultant intracellular response leads to increased cytosolic Ca++ overload, which mediates a cascade of signaling pathway involving activation of PKC, which ultimately induces cardiomyocyte hypertrophy [72].
One other possible explanation for the discrepancy between the present findings and earlier studies of ae3 −/− mice is the assay of cardiac hypertrophy. Earlier investigations probed whole animal physiology, whereas we studied isolated cardiomyocytes in culture. Isolated cardiomyocytes are an established model, which may be more sensitive in detecting alterations of heart cell growth than assessments of functional alterations of heart function sued in earlier studies.
Interestingly, AE3 also forms a complex with CAXIV in the myocardium [73]. In the hypertrophic myocardium of SHR rats, exacerbated AE3/CAXIV activation was proposed to elicit hypertrophic growth of the heart [73]. Thus, the etiology, signaling pathway and pathophysiology of hypertrophic cardiomyopathy are distinct from that mediated by hormonal factors. These differences could account for the disparity between the influence of AE3 on hypertrophy in the present report and that shown in previous models of cardiovascular disease [42, 43]. Since hypertrophic interventions by PE and ANGII failed to induce hypertrophy in the ae3 null cardiomyocytes in our study, the PKC-coupled hypertrophic cascade appears to require AE3. Recently, it was demonstrated that when paced, AE3 null hearts had an impaired force-dependent inotropy characterized by an elevation of protein kinase B (PKB, Akt) phosphorylation and a downregulation of AMPK activity [44]. This observation may provide additional support of the present findings that revealed that ae3 null cardiomyocytes are less susceptible to develop hypertrophy in upon pro-hypertrophic stimulation. Akt phosphorylation is central to the signaling pathway mediated by growth factors that induce physiological hypertrophy [57]. Thus, its activation under stressful conditions may trigger a physiological growth response to counter the pathological signaling cascade stimulated the source of stress. The lack of response by the AE3 knock-out cardiomyocytes to hypertrophic stimulants seen here could arise in part from effects on Akt phosphorylation.

Phenotype of ae3 −/− mice

ae3 −/− mice have been reported to have no apparent defects, and the results of our analysis of the cardiac function of ae3 −/− mice is comparable to these previous studies [42, 43, 74]. Combined analysis of echocardiographic measurements of ventricular wall dimensions, chamber diameter and cardiac function between the two genotypes further suggests that loss of AE3 does not affect hypertrophy or cardiovascular performance. A significant decrease in the HW/BW ratio in ae3 −/− mice is the result of a reduction in heart size, arising from a decrease in cardiomyocyte size. This suggests a critical role for AE3 in heart development.

Role of AE3 in control of cardiomyocyte pHi

Since cardiomyocyte steady-state pH was the same in ae3 −/− and WT mice, loss of Cl/HCO3 exchange activity by ae3 deletion is likely compensated for by another protein. The principal Cl/HCO3 exchanger of cardiomyocytes is SLC26a6 [23], making it the most likely acid loading transporter to compensate for loss of AE3. Since we did not see an increase in SLC26a6 expression, SLC26a6 activation may occur through post-translational mechanisms. We did note an increase of CAII expression in ae3 −/− mice. Since CAII increases the rate of CO2/HCO3 inter-conversion, it increases the rate of observed Cl/HCO3 for SLC26a6 [75]. Thus increased CAII expression might in part compensate for loss of AE3. We also noted a much larger increase in CAII message than we saw for CAII protein. Since proteins carry out cellular functions, not mRNA, we consider the protein change to be a more reliable measure of cell response than the mRNA increase. The reason for a muted increase in CAII protein level in comparison to the mRNA rise remains unclear.
The importance of AE3 in intracellular pH regulation was, however, evident in the reduced rate of pHi recovery from imposed intracellular alkalosis in cardiomyocytes from ae3 −/− mice compared to WT. Endothelin 1 stimulation of myocardial Cl/HCO3 exchange activity in isolated rat papillary muscle is almost totally attributable to the AE3 isoform, on the basis of inhibition by an anti-AE3 antibody [61]. This provides a possible explanation for the resistance of ae3 −/− mice to pro-hypertrophic stimuli; ae3 −/− cardiomyocytes have reduced acidifying activity to counter enhanced NHE1 activity associated with pro-hypertrophic stimulation.

Conclusions

We explored the role of AE3 in the development of cardiomyocyte hypertrophy and cardiovascular pH regulation, using AE3 deficient mice. Cardiomyocytes from ae3 −/− mice were protected from increases in cell surface area, protein synthesis, and fetal gene reactivation in response to hypertrophic stimulation. Steady-state cardiomyocyte pHi in ae3 −/− mice was comparable to WT, but slower to recover from imposed intracellular alkalosis. Our findings demonstrate that AE3 is important in hypertrophic signaling pathways activated by PE and ANGII, possibly acting through the hypertrophic transport metabolon. Pharmacologically targeting AE3 activity in the event of hypertrophy is an attractive strategy to treat heart failure patients.

Acknowledgements

We thank Dr. Gary Shull (University of Cincinnati) for providing breeding stock of ae3 null mice. J.R.C. was a Scientist of the Alberta Heritage Foundation for Medical Research (AHFMR). BVA is an Established Investigator of CONICET (Argentina). The Heart and Stroke Foundation of Alberta provided operating support for this work.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made.
The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
The Creative Commons Public Domain Dedication waiver (https://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

DS: First author of the manuscript and completed the majority of the experimental procedures. BFB: Assisted with manuscript preparation. AQ: Collected the qRT-PCR data and generated/optimized the respective primers. BVA: Collection of HW: BW data. JRC: Conceived and supervised the experiments and assisted with manuscript preparation. All authors read and approved the final manuscript.
Anhänge

Electronic supplementary material

Literatur
1.
Zurück zum Zitat Roger VL, Go AS, Lloyd-Jones DM, Adams RJ, Berry JD, Brown TM, Carnethon MR, Dai S, de Simone G, Ford ES, Fox CS, Fullerton HJ, Gillespie C, Greenlund KJ, Hailpern SM, Heit JA, Ho PM, Howard VJ, Kissela BM, Kittner SJ, Lackland DT, Lichtman JH, Lisabeth LD, Makuc DM, Marcus GM, Marelli A, Matchar DB, McDermott MM, Meigs JB, Moy CS, et al: Heart disease and stroke statistics–2011 update: a report from the American Heart Association. Circulation. 2011, 123 (4): e18-e209.CrossRefPubMed Roger VL, Go AS, Lloyd-Jones DM, Adams RJ, Berry JD, Brown TM, Carnethon MR, Dai S, de Simone G, Ford ES, Fox CS, Fullerton HJ, Gillespie C, Greenlund KJ, Hailpern SM, Heit JA, Ho PM, Howard VJ, Kissela BM, Kittner SJ, Lackland DT, Lichtman JH, Lisabeth LD, Makuc DM, Marcus GM, Marelli A, Matchar DB, McDermott MM, Meigs JB, Moy CS, et al: Heart disease and stroke statistics–2011 update: a report from the American Heart Association. Circulation. 2011, 123 (4): e18-e209.CrossRefPubMed
3.
Zurück zum Zitat Bui AL, Horwich TB, Fonarow GC: Epidemiology and risk profile of heart failure. Nat Rev Cardiol. 2011, 8 (1): 30-41.CrossRefPubMed Bui AL, Horwich TB, Fonarow GC: Epidemiology and risk profile of heart failure. Nat Rev Cardiol. 2011, 8 (1): 30-41.CrossRefPubMed
4.
Zurück zum Zitat Frey N, Katus HA, Olson EN, Hill JA: Hypertrophy of the heart: a new therapeutic target?. Circulation. 2004, 109 (13): 1580-1589.CrossRefPubMed Frey N, Katus HA, Olson EN, Hill JA: Hypertrophy of the heart: a new therapeutic target?. Circulation. 2004, 109 (13): 1580-1589.CrossRefPubMed
5.
Zurück zum Zitat Li M, Naqvi N, Yahiro E, Liu K, Powell PC, Bradley WE, Martin DI, Graham RM, Dell’Italia LJ, Husain A: c-kit is required for cardiomyocyte terminal differentiation. Circ Res. 2008, 102 (6): 677-685.CrossRefPubMedPubMedCentral Li M, Naqvi N, Yahiro E, Liu K, Powell PC, Bradley WE, Martin DI, Graham RM, Dell’Italia LJ, Husain A: c-kit is required for cardiomyocyte terminal differentiation. Circ Res. 2008, 102 (6): 677-685.CrossRefPubMedPubMedCentral
6.
Zurück zum Zitat Karmazyn M: Therapeutic potential of Na-H exchange inhibitors for the treatment of heart failure. Expert Opin Investig Drugs. 2001, 10 (5): 835-843.CrossRefPubMed Karmazyn M: Therapeutic potential of Na-H exchange inhibitors for the treatment of heart failure. Expert Opin Investig Drugs. 2001, 10 (5): 835-843.CrossRefPubMed
7.
Zurück zum Zitat Cingolani HE, Ennis IL: Sodium-hydrogen exchanger, cardiac overload, and myocardial hypertrophy. Circulation. 2007, 115 (9): 1090-1100.CrossRefPubMed Cingolani HE, Ennis IL: Sodium-hydrogen exchanger, cardiac overload, and myocardial hypertrophy. Circulation. 2007, 115 (9): 1090-1100.CrossRefPubMed
8.
Zurück zum Zitat Cingolani HE: Na+/H+ exchange hyperactivity and myocardial hypertrophy: are they linked phenomena?. Cardiovasc Res. 1999, 44 (3): 462-467.CrossRefPubMed Cingolani HE: Na+/H+ exchange hyperactivity and myocardial hypertrophy: are they linked phenomena?. Cardiovasc Res. 1999, 44 (3): 462-467.CrossRefPubMed
9.
Zurück zum Zitat Vandenberg JI, Metcalfe JC, Grace AA: Intracellular pH recovery during respiratory acidosis in perfused hearts. Am J Physiol. 1994, 266: C489-C497.PubMed Vandenberg JI, Metcalfe JC, Grace AA: Intracellular pH recovery during respiratory acidosis in perfused hearts. Am J Physiol. 1994, 266: C489-C497.PubMed
10.
Zurück zum Zitat Vaughan-Jones RD, Spitzer KW: Role of bicarbonate in the regulation of intracellular pH in the mammalian ventricular myocyte. Biochem Cell Biol. 2002, 80 (5): 579-596.CrossRefPubMed Vaughan-Jones RD, Spitzer KW: Role of bicarbonate in the regulation of intracellular pH in the mammalian ventricular myocyte. Biochem Cell Biol. 2002, 80 (5): 579-596.CrossRefPubMed
11.
Zurück zum Zitat Fliegel L: Molecular biology of the myocardial Na+/H+ exchanger. J Mol Cell Cardiol. 2008, 44 (2): 228-237.CrossRefPubMed Fliegel L: Molecular biology of the myocardial Na+/H+ exchanger. J Mol Cell Cardiol. 2008, 44 (2): 228-237.CrossRefPubMed
12.
Zurück zum Zitat Xue J, Mraiche F, Zhou D, Karmazyn M, Oka T, Fliegel L, Haddad GG: Elevated myocardial Na+/H+ exchanger isoform 1 activity elicits gene expression that leads to cardiac hypertrophy. Physiol Genomics. 2010, 42 (3): 374-383.CrossRefPubMedPubMedCentral Xue J, Mraiche F, Zhou D, Karmazyn M, Oka T, Fliegel L, Haddad GG: Elevated myocardial Na+/H+ exchanger isoform 1 activity elicits gene expression that leads to cardiac hypertrophy. Physiol Genomics. 2010, 42 (3): 374-383.CrossRefPubMedPubMedCentral
13.
Zurück zum Zitat Takewaki S, Kuro-o M, Hiroi Y, Yamazaki T, Noguchi T, Miyagishi A, Nakahara K, Aikawa M, Manabe I, Yazaki Y, Nagai R: Activation of Na+-H+ antiporter (NHE-1) gene expression during growth, hypertrophy and proliferation of the rabbit cardiovascular system. J Mol Cell Cardiol. 1995, 27 (1): 729-742.CrossRefPubMed Takewaki S, Kuro-o M, Hiroi Y, Yamazaki T, Noguchi T, Miyagishi A, Nakahara K, Aikawa M, Manabe I, Yazaki Y, Nagai R: Activation of Na+-H+ antiporter (NHE-1) gene expression during growth, hypertrophy and proliferation of the rabbit cardiovascular system. J Mol Cell Cardiol. 1995, 27 (1): 729-742.CrossRefPubMed
14.
Zurück zum Zitat Perez NG, Alvarez BV, de Hurtado MC C, Cingolani HE: pHi regulation in myocardium of the spontaneously hypertensive rat. Compensated enhanced activity of the Na+-H+ exchanger. Circ Res. 1995, 77 (6): 1192-1200.CrossRefPubMed Perez NG, Alvarez BV, de Hurtado MC C, Cingolani HE: pHi regulation in myocardium of the spontaneously hypertensive rat. Compensated enhanced activity of the Na+-H+ exchanger. Circ Res. 1995, 77 (6): 1192-1200.CrossRefPubMed
15.
Zurück zum Zitat Cingolani HE, Camilion De Hurtado MC: Na+-H+ exchanger inhibition: a new antihypertrophic tool. Circ Res. 2002, 90 (7): 751-753.CrossRefPubMed Cingolani HE, Camilion De Hurtado MC: Na+-H+ exchanger inhibition: a new antihypertrophic tool. Circ Res. 2002, 90 (7): 751-753.CrossRefPubMed
16.
Zurück zum Zitat Mraiche F, Oka T, Gan XT, Karmazyn M, Fliegel L: Activated NHE1 is required to induce early cardiac hypertrophy in mice. Basic Res Cardiol. 2011, 106 (4): 603-616.CrossRefPubMed Mraiche F, Oka T, Gan XT, Karmazyn M, Fliegel L: Activated NHE1 is required to induce early cardiac hypertrophy in mice. Basic Res Cardiol. 2011, 106 (4): 603-616.CrossRefPubMed
17.
Zurück zum Zitat Engelhardt S, Hein L, Keller U, Klambt K, Lohse MJ: Inhibition of Na+-H+ exchange prevents hypertrophy, fibrosis, and heart failure in beta(1)-adrenergic receptor transgenic mice. Circ Res. 2002, 90 (7): 814-819.CrossRefPubMed Engelhardt S, Hein L, Keller U, Klambt K, Lohse MJ: Inhibition of Na+-H+ exchange prevents hypertrophy, fibrosis, and heart failure in beta(1)-adrenergic receptor transgenic mice. Circ Res. 2002, 90 (7): 814-819.CrossRefPubMed
18.
Zurück zum Zitat Kang TC, An SJ, Park SK, Hwang IK, Suh JG, Oh YS, Bae JC, Won MH: Alterations in Na+/H+ exchanger and Na+/HCO3 − cotransporter immunoreactivities within the gerbil hippocampus following seizure. Brain Res Mol Brain Res. 2002, 109 (1–2): 226-232.CrossRefPubMed Kang TC, An SJ, Park SK, Hwang IK, Suh JG, Oh YS, Bae JC, Won MH: Alterations in Na+/H+ exchanger and Na+/HCO3 cotransporter immunoreactivities within the gerbil hippocampus following seizure. Brain Res Mol Brain Res. 2002, 109 (1–2): 226-232.CrossRefPubMed
19.
Zurück zum Zitat Karmazyn M: Pharmacology and clinical assessment of cariporide for the treatment coronary artery diseases. Expert Opin Investig Drugs. 2000, 9 (5): 1099-1108.CrossRefPubMed Karmazyn M: Pharmacology and clinical assessment of cariporide for the treatment coronary artery diseases. Expert Opin Investig Drugs. 2000, 9 (5): 1099-1108.CrossRefPubMed
20.
Zurück zum Zitat Karmazyn M, Sostaric JV, Gan XT: The myocardial Na+/H+ exchanger: a potential therapeutic target for the prevention of myocardial ischaemic and reperfusion injury and attenuation of postinfarction heart failure. Drugs. 2001, 61 (3): 375-389.CrossRefPubMed Karmazyn M, Sostaric JV, Gan XT: The myocardial Na+/H+ exchanger: a potential therapeutic target for the prevention of myocardial ischaemic and reperfusion injury and attenuation of postinfarction heart failure. Drugs. 2001, 61 (3): 375-389.CrossRefPubMed
21.
Zurück zum Zitat Alvarez BV, Kieller DM, Quon AL, Robertson M, Casey JR: Cardiac hypertrophy in anion exchanger 1-null mutant mice with severe hemolytic anemia. Am J Physiol Heart Circ Physiol. 2007, 292 (3): H1301-H1312.CrossRefPubMed Alvarez BV, Kieller DM, Quon AL, Robertson M, Casey JR: Cardiac hypertrophy in anion exchanger 1-null mutant mice with severe hemolytic anemia. Am J Physiol Heart Circ Physiol. 2007, 292 (3): H1301-H1312.CrossRefPubMed
22.
Zurück zum Zitat Lohi H, Lamprecht G, Markovich D, Heil A, Kujala M, Seidler U, Kere J: Isoforms of SLC26A6 mediate anion transport and have functional PDZ interaction domains. Am J Physiol Cell Physiol. 2003, 284 (3): C769-C779.CrossRefPubMed Lohi H, Lamprecht G, Markovich D, Heil A, Kujala M, Seidler U, Kere J: Isoforms of SLC26A6 mediate anion transport and have functional PDZ interaction domains. Am J Physiol Cell Physiol. 2003, 284 (3): C769-C779.CrossRefPubMed
23.
Zurück zum Zitat Alvarez BV, Kieller DM, Quon AL, Markovich D, Casey JR: Slc26a6: a cardiac chloride/hydroxyl exchanger and predominant chloride/bicarbonate exchanger of the heart. J Physiol. 2004, 561: 721-734.CrossRefPubMedPubMedCentral Alvarez BV, Kieller DM, Quon AL, Markovich D, Casey JR: Slc26a6: a cardiac chloride/hydroxyl exchanger and predominant chloride/bicarbonate exchanger of the heart. J Physiol. 2004, 561: 721-734.CrossRefPubMedPubMedCentral
24.
Zurück zum Zitat Kim HJ, Myers R, Sihn CR, Rafizadeh S, Zhang XD: Slc26a6 functions as an electrogenic Cl−/HCO3 − exchanger in cardiac myocytes. Cardiovasc Res. 2013, 100: 383-391.CrossRefPubMed Kim HJ, Myers R, Sihn CR, Rafizadeh S, Zhang XD: Slc26a6 functions as an electrogenic Cl/HCO3 exchanger in cardiac myocytes. Cardiovasc Res. 2013, 100: 383-391.CrossRefPubMed
25.
Zurück zum Zitat Niederer SA, Swietach P, Wilson DA, Smith NP, Vaughan-Jones RD: Measuring and modeling chloride-hydroxyl exchange in the Guinea-pig ventricular myocyte. Biophys J. 2008, 94 (6): 2385-2403.CrossRefPubMed Niederer SA, Swietach P, Wilson DA, Smith NP, Vaughan-Jones RD: Measuring and modeling chloride-hydroxyl exchange in the Guinea-pig ventricular myocyte. Biophys J. 2008, 94 (6): 2385-2403.CrossRefPubMed
26.
28.
Zurück zum Zitat Casey JR, Sly WS, Shah GN, Alvarez BV: Bicarbonate homeostasis in excitable tissues: role of AE3 Cl−/HCO3 − exchanger and carbonic anhydrase XIV interaction. Am J Physiol. 2009, 297: C1091-C1102.CrossRef Casey JR, Sly WS, Shah GN, Alvarez BV: Bicarbonate homeostasis in excitable tissues: role of AE3 Cl/HCO3 exchanger and carbonic anhydrase XIV interaction. Am J Physiol. 2009, 297: C1091-C1102.CrossRef
29.
Zurück zum Zitat Chen S, Khan ZA, Karmazyn M, Chakrabarti S: Role of endothelin-1, sodium hydrogen exchanger-1 and mitogen activated protein kinase (MAPK) activation in glucose-induced cardiomyocyte hypertrophy. Diabetes Metab Res Rev. 2007, 23 (5): 356-367.CrossRefPubMed Chen S, Khan ZA, Karmazyn M, Chakrabarti S: Role of endothelin-1, sodium hydrogen exchanger-1 and mitogen activated protein kinase (MAPK) activation in glucose-induced cardiomyocyte hypertrophy. Diabetes Metab Res Rev. 2007, 23 (5): 356-367.CrossRefPubMed
30.
Zurück zum Zitat Li X, Alvarez B, Casey JR, Reithmeier RA, Fliegel L: Carbonic anhydrase II binds to and enhances activity of the Na+/H+ exchanger. J Biol Chem. 2002, 277 (39): 36085-36091.CrossRefPubMed Li X, Alvarez B, Casey JR, Reithmeier RA, Fliegel L: Carbonic anhydrase II binds to and enhances activity of the Na+/H+ exchanger. J Biol Chem. 2002, 277 (39): 36085-36091.CrossRefPubMed
31.
Zurück zum Zitat Sterling D, Reithmeier RA, Casey JR: A transport metabolon. Functional interaction of carbonic anhydrase II and chloride/bicarbonate exchangers. J Biol Chem. 2001, 276 (51): 47886-47894.PubMed Sterling D, Reithmeier RA, Casey JR: A transport metabolon. Functional interaction of carbonic anhydrase II and chloride/bicarbonate exchangers. J Biol Chem. 2001, 276 (51): 47886-47894.PubMed
32.
Zurück zum Zitat Alvarez BV, Xia Y, Sowah D, Soliman D, Light P, Karmazyn M, Casey JR: A carbonic anhydrase inhibitor prevents and reverts cardiomyocyte hypertrophy. J Physiol. 2007, 579: 127-145.CrossRefPubMed Alvarez BV, Xia Y, Sowah D, Soliman D, Light P, Karmazyn M, Casey JR: A carbonic anhydrase inhibitor prevents and reverts cardiomyocyte hypertrophy. J Physiol. 2007, 579: 127-145.CrossRefPubMed
33.
Zurück zum Zitat Brown BF, Quon A, Dyck JR, Casey JR: Carbonic anhydrase II promotes cardiomyocyte hypertrophy. Can J Physiol Pharmacol. 2012, 90 (12): 1599-1610.CrossRefPubMed Brown BF, Quon A, Dyck JR, Casey JR: Carbonic anhydrase II promotes cardiomyocyte hypertrophy. Can J Physiol Pharmacol. 2012, 90 (12): 1599-1610.CrossRefPubMed
34.
Zurück zum Zitat Alvarez BV, Quon AL, Mullen J, Casey JR: Quantification of carbonic anhydrase gene expression in ventricle of hypertrophic and failing human heart. BMC Cardiovasc Disord. 2013, 13: 2-CrossRefPubMedPubMedCentral Alvarez BV, Quon AL, Mullen J, Casey JR: Quantification of carbonic anhydrase gene expression in ventricle of hypertrophic and failing human heart. BMC Cardiovasc Disord. 2013, 13: 2-CrossRefPubMedPubMedCentral
35.
Zurück zum Zitat Becker HM, Deitmer JW: Carbonic anhydrase II increases the activity of the human electrogenic Na+/HCO3 − cotransporter. J Biol Chem. 2007, 282: 13508-13521.CrossRefPubMed Becker HM, Deitmer JW: Carbonic anhydrase II increases the activity of the human electrogenic Na+/HCO3 cotransporter. J Biol Chem. 2007, 282: 13508-13521.CrossRefPubMed
36.
Zurück zum Zitat Gonzalez-Begne M, Nakamoto T, Nguyen HV, Stewart AK, Alper SL, Melvin JE: Enhanced formation of a HCO3 − transport metabolon in exocrine cells of Nhe1 −/− mice. J Biol Chem. 2007, 282 (48): 35125-35132.CrossRefPubMed Gonzalez-Begne M, Nakamoto T, Nguyen HV, Stewart AK, Alper SL, Melvin JE: Enhanced formation of a HCO3 transport metabolon in exocrine cells of Nhe1 −/− mice. J Biol Chem. 2007, 282 (48): 35125-35132.CrossRefPubMed
37.
Zurück zum Zitat Pushkin A, Abuladze N, Gross E, Newman D, Tatishchev S, Lee I, Fedotoff O, Bondar G, Azimov R, Ngyuen M, Kurtz I: Molecular mechanism of kNBC1-carbonic anhydrase II interaction in proximal tubule cells. J Physiol. 2004, 559 (Pt 1): 55-65.CrossRefPubMedPubMedCentral Pushkin A, Abuladze N, Gross E, Newman D, Tatishchev S, Lee I, Fedotoff O, Bondar G, Azimov R, Ngyuen M, Kurtz I: Molecular mechanism of kNBC1-carbonic anhydrase II interaction in proximal tubule cells. J Physiol. 2004, 559 (Pt 1): 55-65.CrossRefPubMedPubMedCentral
38.
Zurück zum Zitat Al-Samir S, Papadopoulos S, Scheibe RJ, Meissner JD, Cartron JP, Sly WS, Alper SL, Gros G, Endeward V: Activity and distribution of intracellular carbonic anhydrase II and their effects on the transport activity of anion exchanger AE1/SLC4A1. J Physiol. 2013, 591: 4963-4982.CrossRefPubMedPubMedCentral Al-Samir S, Papadopoulos S, Scheibe RJ, Meissner JD, Cartron JP, Sly WS, Alper SL, Gros G, Endeward V: Activity and distribution of intracellular carbonic anhydrase II and their effects on the transport activity of anion exchanger AE1/SLC4A1. J Physiol. 2013, 591: 4963-4982.CrossRefPubMedPubMedCentral
39.
Zurück zum Zitat Lu J, Daly CM, Parker MD, Gill HS, Piermarini PM, Pelletier MF, Boron WF: Effect of human carbonic anhydrase II on the activity of the human electrogenic Na+/HCO3 − cotransporter NBCe1-A in Xenopus oocytes. J Biol Chem. 2006, 281: 19241-19250.CrossRefPubMed Lu J, Daly CM, Parker MD, Gill HS, Piermarini PM, Pelletier MF, Boron WF: Effect of human carbonic anhydrase II on the activity of the human electrogenic Na+/HCO3 cotransporter NBCe1-A in Xenopus oocytes. J Biol Chem. 2006, 281: 19241-19250.CrossRefPubMed
40.
Zurück zum Zitat Piermarini PM, Kim EY, Boron WF: Evidence against a direct interaction between intracellular carbonic anhydrase II and pure C-terminal domains of SLC4 bicarbonate transporters. J Biol Chem. 2006, 282: 1409-1421.CrossRefPubMed Piermarini PM, Kim EY, Boron WF: Evidence against a direct interaction between intracellular carbonic anhydrase II and pure C-terminal domains of SLC4 bicarbonate transporters. J Biol Chem. 2006, 282: 1409-1421.CrossRefPubMed
41.
Zurück zum Zitat Li X, Liu Y, Alvarez BV, Casey JR, Fliegel L: A novel carbonic anhydrase II binding site regulates NHE1 activity. Biochemistry. 2006, 45 (7): 2414-2424.CrossRefPubMed Li X, Liu Y, Alvarez BV, Casey JR, Fliegel L: A novel carbonic anhydrase II binding site regulates NHE1 activity. Biochemistry. 2006, 45 (7): 2414-2424.CrossRefPubMed
42.
Zurück zum Zitat Prasad V, Bodi I, Meyer JW, Wang Y, Ashraf M, Engle SJ, Doetschman T, Sisco K, Nieman ML, Miller ML, Lorenz JN, Shull GE: Impaired cardiac contractility in mice lacking both the AE3 Cl−/HCO3 − exchanger and the NKCC1 Na+-K+-2Cl− cotransporter: effects on Ca2+ handling and protein phosphatases. J Biol Chem. 2008, 283 (46): 31303-31314.CrossRefPubMedPubMedCentral Prasad V, Bodi I, Meyer JW, Wang Y, Ashraf M, Engle SJ, Doetschman T, Sisco K, Nieman ML, Miller ML, Lorenz JN, Shull GE: Impaired cardiac contractility in mice lacking both the AE3 Cl/HCO3 exchanger and the NKCC1 Na+-K+-2Cl cotransporter: effects on Ca2+ handling and protein phosphatases. J Biol Chem. 2008, 283 (46): 31303-31314.CrossRefPubMedPubMedCentral
43.
Zurück zum Zitat Al Moamen NJ, Prasad V, Bodi I, Miller ML, Neiman ML, Lasko VM, Alper SL, Wieczorek DF, Lorenz JN, Shull GE: Loss of the AE3 anion exchanger in a hypertrophic cardiomyopathy model causes rapid decompensation and heart failure. J Mol Cell Cardiol. 2011, 50 (1): 137-146.CrossRefPubMed Al Moamen NJ, Prasad V, Bodi I, Miller ML, Neiman ML, Lasko VM, Alper SL, Wieczorek DF, Lorenz JN, Shull GE: Loss of the AE3 anion exchanger in a hypertrophic cardiomyopathy model causes rapid decompensation and heart failure. J Mol Cell Cardiol. 2011, 50 (1): 137-146.CrossRefPubMed
44.
Zurück zum Zitat Prasad V, Lorenz JN, Lasko VM, Nieman ML, Al Moamen NJ, Shull GE: Loss of the AE3 Cl-/HCO3 − exchanger in mice affects rate-dependent inotropy and stress-related AKT signaling in heart. Front Physiol. 2013, 4: 399-CrossRefPubMedPubMedCentral Prasad V, Lorenz JN, Lasko VM, Nieman ML, Al Moamen NJ, Shull GE: Loss of the AE3 Cl-/HCO3 exchanger in mice affects rate-dependent inotropy and stress-related AKT signaling in heart. Front Physiol. 2013, 4: 399-CrossRefPubMedPubMedCentral
45.
Zurück zum Zitat Fischer AH, Jacobson KA, Rose J, Zeller R: Hematoxylin and eosin staining of tissue and cell sections. CSH Protoc. 2008, 2008: pdb prot4986-PubMed Fischer AH, Jacobson KA, Rose J, Zeller R: Hematoxylin and eosin staining of tissue and cell sections. CSH Protoc. 2008, 2008: pdb prot4986-PubMed
46.
Zurück zum Zitat Fischer AH, Jacobson KA, Rose J, Zeller R: Cutting sections of paraffin-embedded tissues. CSH Protoc. 2008, 2008: pdb prot4987-PubMed Fischer AH, Jacobson KA, Rose J, Zeller R: Cutting sections of paraffin-embedded tissues. CSH Protoc. 2008, 2008: pdb prot4987-PubMed
47.
Zurück zum Zitat Sambrano GR, Fraser I, Han H, Ni Y, O’Connell T, Yan Z, Stull JT: Navigating the signalling network in mouse cardiac myocytes. Nature. 2002, 420 (6916): 712-714.CrossRefPubMed Sambrano GR, Fraser I, Han H, Ni Y, O’Connell T, Yan Z, Stull JT: Navigating the signalling network in mouse cardiac myocytes. Nature. 2002, 420 (6916): 712-714.CrossRefPubMed
48.
Zurück zum Zitat Smith PK, Krohn RI, Hermanson GT, Mallia AK, Gartner FH, Provenzano MD, Fujimoto EK, Goeke NM, Olson BJ, Klenk DC: Measurement of protein using bicinchoninic acid. Anal Biochem. 1985, 150: 76-85.CrossRefPubMed Smith PK, Krohn RI, Hermanson GT, Mallia AK, Gartner FH, Provenzano MD, Fujimoto EK, Goeke NM, Olson BJ, Klenk DC: Measurement of protein using bicinchoninic acid. Anal Biochem. 1985, 150: 76-85.CrossRefPubMed
49.
Zurück zum Zitat Lohi H, Kujala M, Kerkela E, Saarialho-Kere U, Kestila M, Kere J: Mapping of five new putative anion transporter genes in human and characterization of SLC26A6, a candidate gene for pancreatic anion exchanger. Genomics. 2000, 70 (1): 102-112.CrossRefPubMed Lohi H, Kujala M, Kerkela E, Saarialho-Kere U, Kestila M, Kere J: Mapping of five new putative anion transporter genes in human and characterization of SLC26A6, a candidate gene for pancreatic anion exchanger. Genomics. 2000, 70 (1): 102-112.CrossRefPubMed
50.
Zurück zum Zitat Link A, Labaer J: Trichloroacetic Acid (TCA) Precipitation of Proteins. Cold Spring HarbProtoc. 2011, 8: 993-994. Link A, Labaer J: Trichloroacetic Acid (TCA) Precipitation of Proteins. Cold Spring HarbProtoc. 2011, 8: 993-994.
51.
Zurück zum Zitat de Hurtado MC C, Alvarez BV, Perez NG, Cingolani HE: Role of an electrogenic Na+-HCO3 − cotransport in determining myocardial pHi after an increase in heart rate. Circ Res. 1996, 79 (4): 698-704.CrossRef de Hurtado MC C, Alvarez BV, Perez NG, Cingolani HE: Role of an electrogenic Na+-HCO3 cotransport in determining myocardial pHi after an increase in heart rate. Circ Res. 1996, 79 (4): 698-704.CrossRef
52.
Zurück zum Zitat de Hurtado MC C, Alvarez BV, Perez NG, Ennis IL, Cingolani HE: Angiotensin II activates Na+-independent Cl−-HCO3 − exchange in ventricular myocardium. Circ Res. 1998, 82 (4): 473-481.CrossRef de Hurtado MC C, Alvarez BV, Perez NG, Ennis IL, Cingolani HE: Angiotensin II activates Na+-independent Cl-HCO3 exchange in ventricular myocardium. Circ Res. 1998, 82 (4): 473-481.CrossRef
53.
Zurück zum Zitat Thomas JA, Buchsbaum RN, Zimniak A, Racker E: Intracellular pH measurements in Ehrlich ascites tumor cells utilizing spectroscopic probes generated in situ. Biochemistry. 1979, 18 (11): 2210-2218.CrossRefPubMed Thomas JA, Buchsbaum RN, Zimniak A, Racker E: Intracellular pH measurements in Ehrlich ascites tumor cells utilizing spectroscopic probes generated in situ. Biochemistry. 1979, 18 (11): 2210-2218.CrossRefPubMed
54.
Zurück zum Zitat Tsutamoto T, Wada A, Maeda K, Hisanaga T, Maeda Y, Fukai D, Ohnishi M, Sugimoto Y, Kinoshita M: Attenuation of compensation of endogenous cardiac natriuretic peptide system in chronic heart failure: prognostic role of plasma brain natriuretic peptide concentration in patients with chronic symptomatic left ventricular dysfunction. Circulation. 1997, 96 (2): 509-516.CrossRefPubMed Tsutamoto T, Wada A, Maeda K, Hisanaga T, Maeda Y, Fukai D, Ohnishi M, Sugimoto Y, Kinoshita M: Attenuation of compensation of endogenous cardiac natriuretic peptide system in chronic heart failure: prognostic role of plasma brain natriuretic peptide concentration in patients with chronic symptomatic left ventricular dysfunction. Circulation. 1997, 96 (2): 509-516.CrossRefPubMed
55.
Zurück zum Zitat Miyata S, Minobe W, Bristow MR, Leinwand LA: Myosin heavy chain isoform expression in the failing and nonfailing human heart. Circ Res. 2000, 86 (4): 386-390.CrossRefPubMed Miyata S, Minobe W, Bristow MR, Leinwand LA: Myosin heavy chain isoform expression in the failing and nonfailing human heart. Circ Res. 2000, 86 (4): 386-390.CrossRefPubMed
56.
Zurück zum Zitat Suurmeijer AJ, Clement S, Francesconi A, Bocchi L, Angelini A, Van Veldhuisen DJ, Spagnoli LG, Gabbiani G, Orlandi A: Alpha-actin isoform distribution in normal and failing human heart: a morphological, morphometric, and biochemical study. J Pathol. 2003, 199 (3): 387-397.CrossRefPubMed Suurmeijer AJ, Clement S, Francesconi A, Bocchi L, Angelini A, Van Veldhuisen DJ, Spagnoli LG, Gabbiani G, Orlandi A: Alpha-actin isoform distribution in normal and failing human heart: a morphological, morphometric, and biochemical study. J Pathol. 2003, 199 (3): 387-397.CrossRefPubMed
57.
Zurück zum Zitat Bernardo BC, Weeks KL, Pretorius L, McMullen JR: Molecular distinction between physiological and pathological cardiac hypertrophy: experimental findings and therapeutic strategies. Pharmacol Ther. 2010, 128 (1): 191-227.CrossRefPubMed Bernardo BC, Weeks KL, Pretorius L, McMullen JR: Molecular distinction between physiological and pathological cardiac hypertrophy: experimental findings and therapeutic strategies. Pharmacol Ther. 2010, 128 (1): 191-227.CrossRefPubMed
58.
Zurück zum Zitat Kovacic S, Soltys CL, Barr AJ, Shiojima I, Walsh K, Dyck JR: Akt activity negatively regulates phosphorylation of AMP-activated protein kinase in the heart. J Biol Chem. 2003, 278 (41): 39422-39427.CrossRefPubMed Kovacic S, Soltys CL, Barr AJ, Shiojima I, Walsh K, Dyck JR: Akt activity negatively regulates phosphorylation of AMP-activated protein kinase in the heart. J Biol Chem. 2003, 278 (41): 39422-39427.CrossRefPubMed
59.
Zurück zum Zitat Alvarez BV, Fujinaga J, Casey JR: Molecular Basis for angiotensin II-induced increase of chloride/bicarbonate exchange in the myocardium. Circ Res. 2001, 89: 1246-1253.CrossRefPubMed Alvarez BV, Fujinaga J, Casey JR: Molecular Basis for angiotensin II-induced increase of chloride/bicarbonate exchange in the myocardium. Circ Res. 2001, 89: 1246-1253.CrossRefPubMed
60.
Zurück zum Zitat Yannoukakos D, Stuart-Tilley A, Fernandez HA, Fey P, Duyk G, Alper SL: Molecular cloning, expression and chromosomal localization of two isoforms of AE3 anion exchanger from human heart. Circ Res. 1994, 75: 603-614.CrossRefPubMed Yannoukakos D, Stuart-Tilley A, Fernandez HA, Fey P, Duyk G, Alper SL: Molecular cloning, expression and chromosomal localization of two isoforms of AE3 anion exchanger from human heart. Circ Res. 1994, 75: 603-614.CrossRefPubMed
61.
Zurück zum Zitat de Cingolani GE C, Ennis IL, Morgan PE, Alvarez BV, Casey JR, de Hurtado MC C: Involvement of AE3 isoform of Na+-independent Cl−/HCO3 − exchanger in myocardial pHi recovery from intracellular alkalization. Life Sci. 2006, 78: 3018-3026.CrossRef de Cingolani GE C, Ennis IL, Morgan PE, Alvarez BV, Casey JR, de Hurtado MC C: Involvement of AE3 isoform of Na+-independent Cl/HCO3 exchanger in myocardial pHi recovery from intracellular alkalization. Life Sci. 2006, 78: 3018-3026.CrossRef
62.
Zurück zum Zitat Farias F, Morgan P, Chiappe de Cingolani G, Camilion de Hurtado MC: Involvement of the Na+-independent Cl−/HCO3 − exchange (AE) isoform in the compensation of myocardial Na+/H+ isoform 1 hyperactivity in spontaneously hypertensive rats. Can J Physiol Pharmacol. 2005, 83 (5): 397-404.CrossRefPubMed Farias F, Morgan P, Chiappe de Cingolani G, Camilion de Hurtado MC: Involvement of the Na+-independent Cl/HCO3 exchange (AE) isoform in the compensation of myocardial Na+/H+ isoform 1 hyperactivity in spontaneously hypertensive rats. Can J Physiol Pharmacol. 2005, 83 (5): 397-404.CrossRefPubMed
63.
Zurück zum Zitat Moor AN, Fliegel L: Protein kinase-mediated regulation of the Na+/H+ exchanger in the rat myocardium by mitogen-activated protein kinase-dependent pathways. J Biol Chem. 1999, 274 (33): 22985-22992.CrossRefPubMed Moor AN, Fliegel L: Protein kinase-mediated regulation of the Na+/H+ exchanger in the rat myocardium by mitogen-activated protein kinase-dependent pathways. J Biol Chem. 1999, 274 (33): 22985-22992.CrossRefPubMed
64.
Zurück zum Zitat Frey N, McKinsey TA, Olson EN: Decoding calcium signals involved in cardiac growth and function. Nat Med. 2000, 6 (11): 1221-1227.CrossRefPubMed Frey N, McKinsey TA, Olson EN: Decoding calcium signals involved in cardiac growth and function. Nat Med. 2000, 6 (11): 1221-1227.CrossRefPubMed
66.
Zurück zum Zitat Villafuerte FC, Swietach P, Youm JB, Ford K, Cardenas R, Supuran CT, Cobden PM, Rohling M, Vaughan-Jones RD: Facilitation by intracellular carbonic anhydrase of Na+-HCO3 - co-transport but not Na+/H+ exchange activity in the mammalian ventricular myocyte. J Physiol. 2013, 592: 991-1007.CrossRefPubMed Villafuerte FC, Swietach P, Youm JB, Ford K, Cardenas R, Supuran CT, Cobden PM, Rohling M, Vaughan-Jones RD: Facilitation by intracellular carbonic anhydrase of Na+-HCO3 - co-transport but not Na+/H+ exchange activity in the mammalian ventricular myocyte. J Physiol. 2013, 592: 991-1007.CrossRefPubMed
67.
Zurück zum Zitat Alvarez BV, Gilmour G, Mema SC, Shull GE, Casey JR, Sauvé Y: Blindness results from deficiency in AE3 chloride/bicarbonate exchanger. PLoS One. 2007, 2: e839-CrossRefPubMedPubMedCentral Alvarez BV, Gilmour G, Mema SC, Shull GE, Casey JR, Sauvé Y: Blindness results from deficiency in AE3 chloride/bicarbonate exchanger. PLoS One. 2007, 2: e839-CrossRefPubMedPubMedCentral
68.
Zurück zum Zitat Kawasaki H, Otani H, Mishima K, Imamura H, Inagaki C: Involvement of anion exchange in the hypoxia/reoxygenation-induced changes in pHi and [Ca2+]i in cardiac myocyte. Eur J Pharmacol. 2001, 411 (1–2): 35-43.CrossRefPubMed Kawasaki H, Otani H, Mishima K, Imamura H, Inagaki C: Involvement of anion exchange in the hypoxia/reoxygenation-induced changes in pHi and [Ca2+]i in cardiac myocyte. Eur J Pharmacol. 2001, 411 (1–2): 35-43.CrossRefPubMed
69.
Zurück zum Zitat Lai ZF, Nishi K: Intracellular chloride activity increases in guinea pig ventricular muscle during simulated ischemia. Am J Physiol. 1998, 275 (5 Pt 2): H1613-H1619.PubMed Lai ZF, Nishi K: Intracellular chloride activity increases in guinea pig ventricular muscle during simulated ischemia. Am J Physiol. 1998, 275 (5 Pt 2): H1613-H1619.PubMed
70.
Zurück zum Zitat Coats CJ, Elliott PM: Genetic biomarkers in hypertrophic cardiomyopathy. Biomark Med. 2013, 7 (4): 505-516.CrossRefPubMed Coats CJ, Elliott PM: Genetic biomarkers in hypertrophic cardiomyopathy. Biomark Med. 2013, 7 (4): 505-516.CrossRefPubMed
71.
Zurück zum Zitat Xia Y, Karmazyn M: Obligatory role for endogenous endothelin in mediating the hypertrophic effects of phenylephrine and angiotensin II in neonatal rat ventricular myocytes: evidence for two distinct mechanisms for endothelin regulation. J Pharmacol Exp Ther. 2004, 310 (1): 43-51.CrossRefPubMed Xia Y, Karmazyn M: Obligatory role for endogenous endothelin in mediating the hypertrophic effects of phenylephrine and angiotensin II in neonatal rat ventricular myocytes: evidence for two distinct mechanisms for endothelin regulation. J Pharmacol Exp Ther. 2004, 310 (1): 43-51.CrossRefPubMed
72.
Zurück zum Zitat Barry SP, Davidson SM, Townsend PA: Molecular regulation of cardiac hypertrophy. Int J Biochem Cell Biol. 2008, 40 (10): 2023-2039.CrossRefPubMed Barry SP, Davidson SM, Townsend PA: Molecular regulation of cardiac hypertrophy. Int J Biochem Cell Biol. 2008, 40 (10): 2023-2039.CrossRefPubMed
73.
Zurück zum Zitat Vargas LA, Alvarez BV: Carbonic anhydrase XIV in the normal and hypertrophic myocardium. J Mol Cell Cardiol. 2012, 52 (3): 741-752.CrossRefPubMed Vargas LA, Alvarez BV: Carbonic anhydrase XIV in the normal and hypertrophic myocardium. J Mol Cell Cardiol. 2012, 52 (3): 741-752.CrossRefPubMed
74.
Zurück zum Zitat Hentschke M, Wiemann M, Hentschke S, Kurth I, Hermans-Borgmeyer I, Seidenbecher T, Jentsch TJ, Gal A, Hubner CA: Mice with a targeted disruption of the Cl−/HCO3 − exchanger AE3 display a reduced seizure threshold. Mol Cell Biol. 2006, 26 (1): 182-191.CrossRefPubMedPubMedCentral Hentschke M, Wiemann M, Hentschke S, Kurth I, Hermans-Borgmeyer I, Seidenbecher T, Jentsch TJ, Gal A, Hubner CA: Mice with a targeted disruption of the Cl/HCO3 exchanger AE3 display a reduced seizure threshold. Mol Cell Biol. 2006, 26 (1): 182-191.CrossRefPubMedPubMedCentral
Metadaten
Titel
Resistance to cardiomyocyte hypertrophy in ae3 −/− mice, deficient in the AE3 Cl−/HCO3 −exchanger
verfasst von
Daniel Sowah
Brittany F Brown
Anita Quon
Bernardo V Alvarez
Joseph R Casey
Publikationsdatum
01.12.2014
Verlag
BioMed Central
Erschienen in
BMC Cardiovascular Disorders / Ausgabe 1/2014
Elektronische ISSN: 1471-2261
DOI
https://doi.org/10.1186/1471-2261-14-89

Weitere Artikel der Ausgabe 1/2014

BMC Cardiovascular Disorders 1/2014 Zur Ausgabe

Update Kardiologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.