gene expression in cardiac hypertrophy

7

Click here to load reader

Upload: kenneth-r-boheler

Post on 31-Aug-2016

215 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Gene expression in cardiac hypertrophy

W: 1990. Ischemic preconditioning reduces infarct size in swine myocardium. Circ Res

66:1133-1142.

Schwarz ER, Mohri M, Sack S, An-as M: 199 1. The role of adenosine and its Al-receptor in ischemic preconditioning [abst]. Circula- tion 84(Suppl2):II-19 1.

Shiki K, Hearse DJ: 1987. Preconditioning of ischemic myocardium: reperfusion-induced arrhythmias. Am J Physiol 253:H1470- H1476.

Thornton J, Downey JM: 1991a. Gi proteins are involved in preconditioning’s protective effect [abst]. Circulation 84(Suppl 2):II- 192.

Thornton J, Downey JM: 1991b. Blockade of ATP-sensitive potassium channels does not prevent preconditioning in the rabbit heart [abst]. Circulation 84(Suppl2):11-432.

Thornton J, Strlplin S, Liu GS, et al.: 1990. Inhibition of protein synthesis does not block myocardial protection afforded by preconditioning. Am J Physiol259:H1822- H1825.

Tsuchida A, Miura T, Iimura 0: 1991. Role of adenosine receptor activation in infarct size limitation by preconditioning in the heart [abst]. Circulation 84(Suppl2):II-191.

Turrens JF, Thornton J, Barnard ML, Snyder S, Liu G, Downey, JM: 1992. Protection from reperfusion injury by preconditioning hearts does not involve increased anti- oxidant defenses. Am J Physiol 262:H585- H.589.

Van Winkle DM, Davis RF: 1991. Ischemic preconditioning of myocardiunm: effect of the adenosine agonist phenylisopropyl aden- nosine (PIA) [abst]. Circulation 84(Suppl 2):II-305.

Van WinkIe DM, Thornton J, Downey DM, Downey JM: 1991. The natural history of preconditioning: cardioprotection depends on duration of transient ischemia and time to subsequent ischemia. Coron Artery Dis 2:613-619.

Vegh A, Sekeres L, Parratt JR: 1990. Protec- tive effects of preconditioning of the isch- aemic myocardium involve cyclo-oxygen- ase products. Cardiovasc Res 24:1020- 1023.

Wyatt DA, Edmunds MC, Rubio R, Beme RM, Lasley RD, Mentzer RM Jr: i989. Adenosine stimulates glycolytic flux in isolated per- fused rat hearts by Al-adenosine receptors. Am J Physiol257:H1952-H1957.

Yellon D, Pasini E, Ferrari R, Downey J, Latchman D: 1990. Whole body heat stress protects the isolated perfused rabbit heart [abst]. Circulation 82(Suppl3):III-463.

Yellon DM, Illiodromitis E, Latchman DS, et al.: 1992. Whole body heat shock fails to limit infarct size in reperfused rabbit heart. Cardiovasc Res 26:342-346. TCM

Gene Expression in Cardiac Hypertrophy Kenneth R. Boheler and Ketty Schwartz

Cardiac hypertrophy due to hemodynamic overload is a harbinger of

morbidity and mortality in humans. The development of hypertrophy

involves both quantitative and qualitative changes in gene expression

that are thought to produce an enlarged organ more capable of meeting

its new functional requirements. The genes are normal, but the way in

which they are regulated is modified. Analysis of these changes and the

mechanisms involved are essential if we are to understand the role that

hypertrophy plays in the pathogenesis of heart failure. (Trends Cardiovasc Med 1992;2:176-182)

Cardiac hypertmphy due to chronic hemo-

dynamic overload is the common end

result of most cardiac disorders and a

consistent feature of cardiac failure.

Cardiac myocytes are terminally differ-

entiated cells that in utero divide but

shortly after birth become nonprolifera-

tive (although recent data suggest that

some myocyte hyperplasia is present in

the adult and aging rat myocardium

[Anversa et al. 19911). Not only the size

but also the phenotype of the nondivid-

ing myocytes changes with chronic hemo-

dynamic overload. The changes are qual-

itative (phenotypic conversions charac-

terized by isoform switches) and quanti-

tative (characterized by modulation of

individual gene expressions), and in some

instances lead to the reexpression of a

phenotype analogous to that seen in the

fetal or senescent heart (reviewed in

Lomprk et al. [1991]). The ensemble of

these pressure-overload-induced changes,

termed mechanogenic transduction, pro-

duces an enlarged organ considered to

be better adapted to the new functional

demands. An important feature of this mechanogenic transduction is that it

Kenneth R. Boheler and Ketty Schwartz are at INSERM U127, H6pital Lariboisitie, 75010 Paris, France.

does not involve genetic defects: the

genes are normal, but the way in which

they are regulated is modified. Cardiac

hypertrophy is thus due to enhanced

growth of a fixed population of myocytes

with some altered expressions of normal

genes. This review describes the molecu-

lar phenotype of the hypertrophied heart,

with particular emphasis on the homol-

ogy with development and aging and on

the regulatorial levels that control gene

expressions in the heart.

l Qualitative Changes in Gene Expression

Qualitative changes in gene expression

are due to differential expression of

multigene families (Figure 1). The first

identified and most thoroughly studied

example of isoform switching involves

the myosin heavy-chain family (reviewed

in Lornpre et al. [1991]). The thick filament of the cardiac sarcomere can contain two possible myosin heavy- chain isoforms, a-MHC and P-MHC. An effect of a permanent hemodynamic overload is activation of the P-MHC gene and a repression of the a-MHC one. This

change occurs in all tissues and species tested so far, but the potential for an

increase in P-MHC depends upon the

176 01992, Elsevier Science Publishing Co., lOSO-1738/92/$5.00 TCM Vol. 2, No. 5, 1992

Page 2: Gene expression in cardiac hypertrophy

Fibronectin (fetal)

Collagen

(Types I, 111,

Creatine Kinase BB

Creatine Kinase MB Lactate Dehydrogenase

P-Myosin Heavy Chain

Atrial Light Chain 1

(M-LDH)

Skeletal a-actin

Smooth muscle actin Troponin T (fetal isoform)

fi Tropomyosin

Myocytes

ti

a3 NalK ATPase

Figure 1. Qualitative changes characterized by isofonn switching that occur during early to say whether these fetal genes hemodynamic overload-induced cardiac hypertrophy. share common regulatory mechanisms.

initial phenotype: it is high in rat ventri- cles and in human atria, which normally contain -90% a-MHC, and it is small in human ventricles, which contain mainly P-MHC. The change from a-MHC to P-MHC (or isomyosin Vl, the aa homo- dimer, to V3, the pp homodimer) results in a slower rate of ATP cycling by myosin and a lower velocity of contraction in the hypertrophied fiber. The result is an improved economy of force development that is usually considered adaptive. Be- cause P-MHC is predominant in rat fetal ventricles, the idea was developed that reactivation of a fetal program occurs with hemodynamic overloading. P-MHC also increases with aging (O’Neill et al. 1991), consistent with the idea that the continuous loss of myocytes that occurs in aged ventricles generates a greater workload on the remaining myocytes, which may serve as a chronic mechani- cal stimulus.

Actin-stimulated myosin ATPase ac- tivity and crossbridge cycling generate force and lead to contraction. The activa- tion may involve two different sarco- merit isoforms of actin, a-cardiac and a-skeletal, and pressure overload in- duces changes in their expressions. The

a-skeletal isoactin gene in rats is transi- torily upregulated and, since it is also active in utero, represents the second example of a fetal program reactivation by hemodynamic overload (Schwartz et al. 1986; Izumo et al. 1988). a-Skeletal actin is not increased in the senescent myocardium (Carrier et al. 1992), and thus the analogy between compensated hypertrophy and aging is apt for P-MHC and less so with a-skeletal actin. In humans, in contrast to rodent heart, a-skeletal actin is upregulated during development and is the major isofonn of adult control, hypertrophied, and failing hearts (Boheler et al. 1991). Another actin isoform, a-smooth, normally ex- pressed in vascular smooth muscles, is the first muscular isoactin expressed during cardiac myogenesis. It was re- cently found to be reactivated perma- nently with pressure-overload hypertro- phy, and the high levels of smooth muscle actin mRNAs in the hypertro- phied myocardium strongly suggest that this isoform is present in the myocytes (Black et al. 1991). The functional and structural effects of reinducing the two fetal actin isogenes, a-skeletal and vas- cular smooth, are unknown, and it is too

The expression of other multigene families controlling cardiac contraction is also modified in cardiac hypertrophy, and as more probes become available, the number will probably increase in the near future. The two other types of myosin subunits, myosin light chains 1 and 2, both exist as two isofotms, the atria1 and ventricular isotypes. The atria1 isoform ALCl increases in several forms of ventricular hypertrophy, whereas the ventricular types VLC 1 and VLC2 appear in pressure-overloaded human atria (Cum- mins 1982). The physiologic significance of such transitions is unclear, except that the transition of ALC2 to VIC2 in the atria may account for the increased calcium sensitivity of skinned atria1 fi- bers from patients with dilated cardio- myopathy, and may indirectly reduce the myosin ATPase activity and maximum velocity of shortening of the fiber, result- ing in an improved economy for atria1 contraction (Wankerl et al. 1990). There is also evidence of qualitative changes in some other contractile proteins of the thin filament. The mRNA encoding the p isoform of tropomyosin, present during fetal and early postnatal life in rats, is transitorily activated at the onset of pressure-overload hypertrophy (Izumo et al. 1988). One subunit of the troponin complex, TnT, exists as two or more

TCM Vol. 2, No. 5, I992 01992, Elsevier Science Publishing Co., 1050-1738/92/$5.00 177

Page 3: Gene expression in cardiac hypertrophy

cardiac isoforms, which most probably, and in contrast to all other proteins previously mentioned, are produced by alternative splicing of a single gene. The fetal isotype is not upregulated with pressure overload in rats (Saggin et al. 1988), at least in the compensated stage, whereas in humans there appears to be a partial reactivation of fetal cardiac TnT expression in failing hearts (Anderson et al. 1991). If confirmed, this would be the first molecular basis of a thin filament dysregulation in heart failure and would partially provide an explanation to the unanswered question of why the myofi- brillar ATPase activities of human fail- ing hearts are lower than in controls in spite of the same l3-MHC content.

The isofotm switches produced in response to increased afterload are not limited to components of the contractile apparatus. There are also changes in tissue content of isoenzymes that control intermediary metabolism. Increases of fetal type isoenzymes (BB+MB) of creatine kinase and of the M-LDH isoform of lactate dehydrogenase are believed to produce an increased glycolytic capacity that could be a favorable adaptation ensuring that ATP resynthesis via the creatine kinase reaction remains (Ing- wall et al. 198.Q although it should also be noted that this creatine kinase isofonn switch was not observed in hypertensive rats (Fontanet et al. 1991). There are also changes in membrane proteins, which, as for the contractile proteins, depend on the animal species. A twofold increase in the accumulation of the isomRNA en- coding the a3 isoform of Na+/K+-ATPase (the cardiac glycoside receptor) was found in rat heart hypertrophy, but it is un- likely that this increase totally accounts for the altered properties of the enzyme (Charlemagne et al., unpublished data, 1992). In the failing human heart, Na+/K+- ATPase isoform expression is not altered and [sH]ouabain binding is nearly nor- mal, suggesting that adaptation to digi- talization in the failing human heart results from other mechanisms (Allen et al. 1992).

What about the nonmyocyte compart- ments? Many studies have shown that fibrillar collagen and fibronectin accu- mulate within the interstitium and the adventitia of the intramyocardial arter- ies together with a medial thickening of these vessels. This remodeling is thought to be the main cause of abnormalities in

myocardial stiffness and in vasomotor reactivity of intramyocardial resistance vessels (reviewed in Weber and Brilla [ 199 11). It is now clear that this compart- ment undergoes profound molecular changes; the distribution and density of collagen types I and III change tran- siently in experimental hypertrophy, and that of type IV changes permanently. The fetal forms of cellular fibronectin produced by alternative splicing of a single gene encoding multiple fibronectin isotypes accumulate in the wall of coro- nary arteries and in focal areas of the myocardium early after rat aortic sten- osis (Samuel et al. 1991). It is hypothe- sized that these fetal variants influence proliferation and migration of nonmus- cular cells as well as adhesion of my- ocytes of the extracellular matrix pro- teins that play an active role in cardiac remodeling.

From the above observations, it is therefore clear that when isoform switches occur, the form normally expressed in fetal life is either transitorily or perma- nently reinduced by hemodynamic over- load, and that reprogramming of cardiac gene expression is a common feature of all cardiac cell populations, including striated myocytes, smooth muscle cells, and fibroblasts.

l Quantitative Changes in Gene Expression

The central feature of myocyte hy- pet-trophy is an increased protein con- tent per cell, but this global increase may also involve a selective induction, repres- sion, or nonactivation of any one gene (Figure 2).

The first set of genes induced by pressure overload includes several tran- scription factors or proto-oncogenes, repre- senting what is now called the imme- diate early gene program. These proto- oncogenes are all transiently upregulated, but the kinetics vary with the experi- mental model. In the classic model of suprarenal aortic constriction that is usually considered to produce a rather moderate hemodynamic overload, ac- cumulation of c-fis and c-myc tran- scripts are detected 30 min to 2 h after constriction, peak at 8 h, and return to baseline by 48 h after surgery (Komuro et al. 1988; Izumo et al. 1988). In the model of thoracic aortic constriction that produces an acute pressure over-

load, c-fos and c-myc expression as well as that of other proto-oncogenes (c-Jon, junJ3, and nur77), peaks as early as 15 min after constriction and returns to basal values within 2-3 h (Rockman et al. 1991). Fos is localized in myocytes, whereas Myc is expressed primarily in nonmuscle cells, showing a cellular seg- regation of these proto-oncogene prod- ucts (Snoeckx et al. 1991). Their role in the development of cardiac hypertrophy is completely unknown although poten- tially critical. Heterodimers of Fos/Jun or homodimers of Jun act cooperatively to induce flexure at consensus AP- 1 sites in transcriptional promotor regions for a number of genes and may react with other factors to regulate differential gene expression (Kerppola and Curran 1991). As for Myc, it belongs to a larger family of proteins (helix-loophelix) that has been shown to be important in skeletal muscle differentiation (Tapscott and Wein- traub 1991). In heart, its role is entirely unclear, but there is evidence suggesting that it may be important early in cardiac development or cell division. Indeed, transgenic mice expressing c-myc tran- scripts undergo a proliferative response early on, which ends during postnatal development (Jackson et al. 1990). One of the challenges in the field of cardiac hypertrophy is thus to determine whether the immediate early gene program is necessary for or only coincidental with the activation of other cardiac genes.

Other ubiquitous genes involved in tissue response to stress and tissular growth are selectively upregulated after pressure overload. There is a transient and early expression of three heat-shock proteins (HSP70, HSP68, and HSP58) that could confer some degree of protec- tion to the actively growing myocytes (Delcayre et al. 1988). There is also evidence for production of trophic sub- stances and other transcription factors within the heart. Experimental aortic banding increases the accumulation of mRNAs encoding transforming growth factor pl (TGFPl), insulinlike growth factor-I, and early growth response factor 1 (Egr-1), a serum-inducible zinc finger protein. Basic fibroblast growth factor (bFGF) immunoreactivity also increases markedly in the hypertrophied myocytes, and it is worth noting that basic and acidic FGF exert antithetical and re- ciprocal effects on skeletal a-actin tran- scription that is only observed in cardiac

178 01992, Elsevier Science Publishing Co., 1050-1738/92/$5.00 TCM Vol. 2, No. 5, 1992

Page 4: Gene expression in cardiac hypertrophy

TroDhic Trigeers Adrenergic agonists Angiotensin II Endothelin

TGFf.3

bFGF, aFGF

echanical M Stretch Load

Atrial

rRNA

SERCA2a 1

Phospholamban 4

Natriuretic Factor ‘?

T Muscarinic

Receptors 1

/&y

PI-Adrenergic

Receptors L

Myocytes

Heat-Shock Proteins

HSP 70 1‘

HSP 68 t

HSP 58 t

Figure 2. Quantitative changes during hemodynamic overload-induced cardiac hypertrophy, including those in the immediate early gene program (in italics) and other more long-term changes in gene expression. A nonexhaustive list of some potential mechanical and trophic triggers is also presented.

myocytes (reviewed in Schneider and Parker [1991] and Chien et al. [1991]). Proof is still required to determine whether these factors (or other members of the vast family of growth factors) play a direct role in vivo, but it is now clear that cardiac muscle cells in vitro are targets for regulation by peptide growth factors, and an autocrine or paracrine model of myocardial hypertrophy is strongly sug- gested by these findings.

The ventricular expression of atria1 natriuretic factor (ANF) in response to a number of diverse hypertrophic stimuli including pressure and volume overload is one of the best characterized examples of selective long-term induction. During embryonic development, the ANF gene is expressed in both the ventricles and the atria, but shortly after birth its expres- sion is largely limited to the atria. With cardiac hypertrophy, this gene is reex- pressed in the ventricles of mice, rats, and humans, indicating that this induc- tion is a generalized response to pressure overload (reviewed in Mercadier and Michel [1990]). Since this peptide hormone plays a significant role in regulating blood pressure and natriu-

resis, its reinduction may contribute to a reduction in pre- and afterload and thus normalize the working conditions of the cardiac pump. From the point of view of gene expression, this induction repre- sented the third example of a reactivated fetal program.

In contrast, failure to upregulate the expression of individual genes can lead to a relative decrease of the level of gene products in the hypertrophied myocar- dium. The Ca2+ATPuse of the sarcoplas- mic reticulum is a product of a multi- gene family and, in ventricular and atria1 tissue, a single isoform predominates, the slow skeletal/cardiac form SERCAZa. With pressure overload, there is neither an induction of a new isofonn nor an apparent activation of this gene. The result is a net decrease in its transcripts in rats, rabbits and humans, paralleled by a decrease in its protein levels in rat and rabbit (Komuro et al. 1989; Nagai et al. 1989; de la Bastie et al. 1990; Mercad- ier et al. 1990). The mRNA decrease occurs relative to total rRNA content and total MHC RNA, indicating a potential imbalance between the contractile and relaxation properties of the heart, and is

probably implicated in delayed rates of Ca2+ transients and myocardial relaxa- tion (see Gwathmey et al. [ 19913. Thus, it appears that, in hemodynamically overloaded hearts, contraction is regu- lated via isoform switches and relaxation via quantitative modulation of a single isoform. mRNA expressions of phospho- lamban, one of the modulators of the activity of SERCA2a, decrease in parallel to that of SERCA2a, which may play a role in the differential sensitivity of hypertrophied and normal hearts to p- adrenergic action (Nagai et al. 1989). Other examples of apparently nonacti- vated genes include the Na channel (see Chien et al. [1991]) and the P1-adrener- gic and muscarinic receptors. For the receptors, it is not yet known whether this is due to nonactivation of the corre- sponding genes or to other types of mechanisms, since the results have come mostly from binding studies (reviewed in Bristow et al. [1990] and Delcayre and Swynghedauw [1991]).

All of the above data show that altera- tions in gene expressions in the hypertro- phied and failing hearts are selective. Evaluation of these alterations in human hearts has been limited in the past by the availability of cardiac tissue, but quanti- fication of gene expressions in endomyo- cardial biopsy specimens by polymerase chain reaction amplifications will cer-

TCM Vol. 2, No. 5, 1992 01992, Elsevier Science Publishing Co., IOSO-1738/92/$5.00 179

Page 5: Gene expression in cardiac hypertrophy

tainly soon advance our knowledge in this field (Feldman et al. 1991).

l Regulatory Levels

Since protein expressions depend upon both transcriptional and translational processes, differentiation between these events is necessary if we are to under- stand the regulatory levels that control gene expressions. The difficulties in study- ing these processes in vivo caused most early studies to be conducted in vitro, mostly through the use of beating or arrested perfused hearts exposed to vari- ous perfusion pressures. Increased aor- tic pressure per se is sufficient to acceler- ate protein synthesis in heart via acceleration of both peptide chain initia- tion and elongation, indicating increases in the translational capacity of preex- isting ribosomes (reviewed in Morgan and Baker [ 19911). There is also en- hanced protein synthesis through pre- translational events due to an increase in ribosome content resulting from en- hanced RNA polymerase I activity and rRNA synthesis. Direct evidence of pretranslational regulation of individual genes has come from in vivo studies in rats and rabbits, where the synthesis of P-MHC and its mRNA levels changed in parallel after pressure overload (Izumo et al. 1987; Nagai et al. 1987). Several promising in vivo approaches toward the elucidation of mechanisms that control gene expression in response to hemody- namic overload have been recently devel- oped. We have succeeded in isolating transcriptionally active nuclei from hearts of neonatal and adult rats, and found that the transcript accumulations for a- and P-MHC and cardiac and skeletal a-actins correspond directly to the relative transcriptional activity for each respective gene in 23-day-old ani- mals (Boheler et al. 1992). Based on these results, we now speculate that the changes in P-MHC expression seen in vivo with overload will be regulated primarily through transcriptionally me- diated mechanisms, but further proof is needed. Another approach is direct injec- tion of DNA into the myocardium (re- viewed in Ieinwand and Leiden [1991]). Unlike most cells, cardiac myocytes share with skeletal fibers the potential of tak- ing up and expressing foreign DNA. Since reporter plasmids containing rat a-myosin heavy-chain promoter regions

seem to be regulated after injection as the endogenous MHC gene, the model of gene transfer should prove useful in studying cardiac gene regulations in normal and pathologic conditions. Fi- nally the major difficulties of aortic constriction in transgenic mice (hemo- dynamic analysis and surgery) were both recently and successfully resolved, which will enable the delineation of the hemody- mimic responsive elements for a given gene (Rockman et al. 1991). The first data already suggest that distinct sets of regulatory sequences mediate atria1 spe- cific and inducible expression of ANF during in vivo hypertrophy.

How is it possible for the cell to have a differential response and an induction of a large pool of RNAs, yet a specific regulation of a subset of RNAs? One in vivo example of how selective induc- tions or nonactivations may occur in- volves G protein expression. G protein subunits form complexes that couple the functions of receptors (for example, adrenergic) to that of effecters (for instance, adenylate cyclase or phosphol- ipase C) and whose modulation may induce significant functional changes in myocardial cells (reviewed in Neer and Clapham [ 19921). With isoproterenol infusions into rats, the concentrations of the inhibitory subunits Gi,* and Gia_s transcripts increase while those of the stimulator-y subunit G,, remain largely unchanged (Eschenhagen et al. 1991). Recent results from in vitro transcrip- tion assays indicate that the gene for

Gia-2 is transcriptionally activated to levels that correspond well with its accumulated mRNA; however, the gene for G,, is not activated (Muller et al., unpublished data, 1992). The difference in this activation probably resides in the promoter regions for these two genes, where Gicr.z unlike G,, contains a CAMP responsive element that is responsive to the P-adrenergic stimulation.

l Potential Triggers

The use of in vitro models and cell culture systems has led to the greatest insights into the possible triggers in the development of cardiac hypertrophy. Although they cannot mimic exactly the conditions of the intact heart, they are very useful in describing individual pathways by which a change in hemody- namic load is transduced into a change

in cardiac mass and cardiac phenotype. These systems will not be described here in detail because they have been reviewed at length elsewhere (Cooper et al. 1989; Simpson et al. 1989; Mor- gan and Baker 1991; Chien et al. 1991; Schneider et al. 1991). They include cultured myocytes isolated from neo- nate and adult hearts, isolated papil- lary muscles, skinned fibers, and beat- ing or arrested isolated perfused hearts subjected to various pressure and vol- ume-loading conditions. In vivo studies have also been extensively used, includ- ing chronic or acute administration of adrenergic agonists or antagonists, thy- roxine, angiotensin II, converting en- zyme inhibitors, and the unloading of papillary muscles. Very schematically, two types of triggers have been found: mechanical triggers via stretch, load, cell deformation, and contraction; and trophic triggers such as peptide- derived growth factors (including fibro- blast growth factors, transforming growth factors, angiotensin IX, and endothelin), thyroxine, adrenocorticoids, insulin, growth hormone, or adrenorecep- tor activation.

What are the intracellular mech- anisms coupled to these initiating sig- nals? The data are largely circumstan- tial but highly suggestive of the involvement of protein-kinase-C- and/ or protein-kinase-A-dependent pathways and of ionic messengers like sodium and calcium. The best evidence, how- ever, comes from in vitro studies and activation of protein kinase C (PKC). Stimulation of PKC by phorbol ester addition (phorbol myristate acetate) is capable of partially mimicking the a,-agonist-mediated induction of the immediate early gene program (c-fos, c-jun, Egr-1) and of skeletal a-actin, P-MHC, and myosin light-chain-2 gene expressions in cultured neonatal cells even in the presence of cycloheximide, a protein synthesis blocker (Simpson et al. 1989; Dunnmon et al. 1990). Further- more, transcription of the P-MHC isogene and other reporter plasmid constructs containing AP-1 recognition sites is preferentially stimulated by l3-PKC in cardiac neonatal myocytes (Kariya et al. 1991). Phorbol-ester- induced hypertrophic growth of ne- onatal cardiomyocytes apparently in- volves an increase in ribosomal DNA transcription (Al10 et al. 1991), and

180 01992, Elsevier Science Publishing Co., 1050-1738/92/$5.00 TCM Vol. 2, No. 5, 1992

Page 6: Gene expression in cardiac hypertrophy

stretch-activated induction of the early response gene c-fos is suppressed by protein kinase C inhibitors and in- duced with phorbol esters (Komuro et al. 1991). From these results has de- veloped the hypothesis that activation of PKC may lead to phosphorylation of a transcription factor that can bind to the regulatory regions of accessible genes and the specific activation of responsive genes. Direct evidence in vivo is, however, lacking.

JM: 1991. Myocyte mitotic division in the

aging mammalian rat heart. Circ Res

69:1159-1164.

Black FM, Packer SE, Parker TG, et al.: 1991. The vascular smooth muscle a-actin gene is reactivated during cardiac hypertrophy pro- voked by load. J Clin Invest 88:1581-1588.

Boheler KR, Carrier L, de la Bastie D, et al.: 199 1. Skeletal actin mRNA increases in the human heart during ontogenic develop- ment and is the major isoform of control and failing adult hearts. J Clin Invest

88:323-330.

l Perspectives

As we have seen, cardiac hypertrophy due to chronic hemodynamic overload is a complicated process. At the molecular level, one of the primary goals of the future will be the complete analysis of the mechanisms regulating individual gene expressions in vivo. Several pos- sible mechanisms have been described, but direct proof and their roles in vivo are still lacking. With the availability of cardiomyocyte AT-l tumor cell lineage (Delcarpio et al. 1991), it should be possible to differentiate further the mech- anisms involved in gene regulation with cardiac hypertrophy, but the use of in vivo models and transgenic mice will be of paramount importance. They will provide the means to compare changes in gene expressions directly with physio- logic functions, leading to a more com- plete understanding of which pheno- typic changes have beneficial versus detrimental effects. This must be the ultimate goal if we are to elucidate the pathogenesis and treatment of heart failure.

Boheler KR, Chassagne C, Martin X, Wisnewsky C, Schwartz K: 1992. Cardiac expressions of a- and 8-myosin heavy chains and sarcomeric a-actins are regu- lated through transcriptional mechanisms. J Biol Chem 267:12,979-12,985.

Bristow MR, Hershberger RE, Port JD, et al.: 1990. P-Adrenergic pathways in nonfailing and failing human ventricular myocar- dium. Circulation 82(Suppl 1):112-125.

Carrier L, Boheler KR, Chassagne C, et al.: 1992. Expression of the sarcomeric actin isogenes in the rat heart with development and senescence. Circ Res 70:999-1005.

Chien KR, Knowlton KU, Zhu H, Chien S:

1991. Regulation of cardiac gene expres- sion during myocardial growth and hy- pertrophy: molecular studies of an adaptive physiologic response. FASEB J 5:3037- 3046.

Cooper G lV Kent RL, Mann DL: 1989. Load induction of cardiac hypertrophy. J Mol Cell Cardiol21 (Suppl5): 1 l-30.

Cummins P: 1982. Transitions in human atria1 and ventricular myosin light-chain

isoenzymes in response to cardiac pressure overload induced hypertrophy. Biochem J 205:195-204.

References

Allen PD, Schmidt TA, Marsh JD, Kjeldsen K: 1992. NaK-ATPase expression in normal and failing human left ventricle. Basic Res Cardiol (in press).

Al10 SN, McDermott PJ, Carl LL, Morgan HE: 1991. Phorbol ester stimulation of protein kinase C activity and ribosomal DNA tran- scription: role in hypertrophic growth of cultured cardiomyocytes. J Biol Chem 266:22,003-22,009.

De la Bastie D, Levitsky D, Rappaport L, et al.: 1990. Function of the sarcoplasmic reticu- lum and expression of its Ca*+-ATPase gene in pressure overload-induced cardiac hy- pertrophy in the rat. Circ Res 66:554-564.

Delcarpio JB, Nicholas A, Lanson NA, Field

LJ, Claycomb C: 199 1. Morphological char- acterization of cardiomyocytes isolated from a transplantable cardiac tumor derived from transgenic mouse atria (AT-~ cells). Circ Res 691591-1600.

Anderson PAW, Malouf NN, Oakeley AE, Pagani ED, Allen PD: 1991. Troponin T isoform expression in humans: a compari- son among normal and failing heart, fetal heart, and adult and fetal skeletal muscle. Circ Res 69:1226-1233.

Delcayre C, Swynghedauw B: 1991. Biological adaptation and dysadaptation of the heart to chronic arterial hypertension: a review. J Hypertens 9(Suppl2):S23-S29.

Delcayre C, Samuel JL, Marotte F, Best- Belpomme M, Mercadier JJ, Rappaport L: 1988. Synthesis of stress proteins in rat cardiac myocytes 24 days after imposition of hemodynamic overload. J Clin Invest 82r460-468.

Anversa P, Fitzpatrick D, Arganis S, Capasso Dunnmon PM, Iwaki K, Henderson SA, Sen A,

Chien f(R: 1990. Phorbol esters induce

immediate-early genes and activate cardiac gene transcription in neonatal rat myocar-

dial cells. J Mol Cell Cardiol 22:901-910.

Eschenhagen T, Mende U, Nose M, et al.:

1991. Isoprenaline-induced increase in mRNA levels of inhibitory G-protein a- subunits in rat heart. Naunyn Schmiede- bergs Arch Pharmacol343:609-615.

Feldman AM, Ray EP, Silan MC, et al.: 1991. Selective gene expression in failing human heart: quantification of steady-state levels of messenger RNA in endomyocardial biop-

sies using the polymerase chain reaction. Circulation 83:1866-1872.

Fontanet HL, Trask RV, Haas RC, et al.: 199 1. Regulation of expression of M, B, and mitochondrial creatine kinase mRNAs in the left ventricle after pressure overload in rats. Circ Res 68:1007-1012.

Gwathmey JK, Warren SE, Briggs GM, et al.: 199 1. Diastolic dysfunction in hypertrophic cardiomyopathy. J Clin Invest 87:1023- 1031.

Ingwall JS, Kramer MF, Fifer MA, et al.: 1985. The creatine kinase system in normal and diseased human myocardium. N Engl J Med 313:1050-1054.

Izumo S, Lompre AM, Matsuoaka R, et al.: 1987. Myosin heavy chain messenger RNA and protein isoform transitions during car- diac hypertrophy. J Clin Invest 79:970-977.

Izumo S, Nadal-Ginard B, Mahdavi V: 1988. Protooncogene induction and reprogram- ming of cardiac gene expression produced by pressure overload. Proc Natl Acad Sci USA 85:339-343.

Jackson T, Allard MF, Sreenan CM, et al.: 1990. The c-my proto-oncogene regulates cardiac development in transgenic mice. Mol Cell Biol 7:3709-37 16.

Kariya K, Karns LR, Simpson PC: 1991. Expression of a constitutively activated mutant of the 8-isozyme of protein kinase C in cardiac myocytes stimulates the pro- moter of the P-myosin heavy chain isogene. J Biol Chem 266:10,023-10,026.

Kerppola TK, Curran T: 1991. Fos Jun het- erodimers and Jun homodimers bend DNA in opposite orientations: implications for transcription factor cooperativity. Cell 66~317-326.

Komuro I, Kurabayashi M, Takaku F, Yazaki Y: 1988. Expression of cellular oncogenes in the myocardium during the developmen- tal stage and pressure-overloaded hypertro- phy of the rat heart. Circ Res 62:1075-1079.

Komuro I, Kurabayashi M, Shibazaki Y, Takaku F, Yazaki Y: 1989. Molecular clon- ing and characterization of a Ca*+ + Mgz+- dependent adenosine triphosphatase from rat cardiac sarcoplasmic reticulum: regula- tion of its expression by pressure overload

TCM Vol. 2, No. 5, I992 01992, Elsevier Science Publishing Co., 1050-1738/92/$5.00 181

Page 7: Gene expression in cardiac hypertrophy

and developmental stage. J Clin Invest 83:1102-l 108.

Komuro I, Katoh Y, Kaida T, et al.: 1991. Mechanical loading stimulates cell hypertro- phy and specific gene expression in cul- tured rat cardiac myocytes. J Biol Chem 266:1265-1268.

ATPase and phospholamban mRNA expres- sion in response to pressure overload and thyroid hormone. Proc Natl Acad Sci USA 86:2966-2970.

Leinwand LA, L&den JM: 1991. Gene transfer into cardiac myocytes in vivo. Trends Car- diovasc Med 1:271-276.

Lompre AM, Mercadier JJ, Schwartz K: 1991. Changes in gene expression during cardiac growth. Int Rev Cytol 124:137-186.

Mercadier JJ, Michel JB: 1990. Neurohormonal axis in cardiac hypertension and failure. In Swynghedauw B, ed. Research in Cardiac Hypertrophy and Failure. Paris, INSERMl John Libbey Eurotext, pp 401413.

Mercadier JJ, Lompr6 AM, Due P, et al.: 1990. Altered sarcoplasmic reticulum Ca2+- ATPase gene expression in the human ventricle during end-stage heart failure. J Clin Invest 85:305-309.

Neer EJ, Clapham DE: 1992. Signal transduc- tion through G protein in the cardiac myocyte. Trends Cardiovasc Med 2:6-l 1.

O’Neill L, Holbrook NJ, Fargnoli J, Lakatta EG: 1991. Progressive changes from young adult age to senescence in mRNA for rat cardiac myosin heavy chain genes. Cardio- science 2: l-5.

Rockman HA, Ross RS. Harris AN, et al.: 199 1. Segregation of atrial-specific and inducible expression of an atrial natriuretic factor transgene in an in vivo murine model of cardiac hypertrophy. Proc Natl Acad Sci USA 88:8277-8281.

Saggin L, Ausoni S, Gorza L, Sartore S, Schiaffino S: 1988. Troponin T switching in the developing rat heart. J Biol Chem 263:18,488-18,492.

Morgan HE, Baker KM: 1991. Cardiac hy- pet-trophy: mechanical, neural, and en- docrine dependence. Circulation 83: 13-25.

Nagai R, Pritzl N, Low RB, et al.: 1987. Myosin isozyme synthesis and mRNA levels in pressure-overloaded rabbit hearts. Circ Res 60:692-699.

Samuel JL, Barrieux A, Dufour S, et al.: 1991. Accumulation of fetal fibronectin mRNAs during the development of rat cardiac hypertrophy induced by pressure overload. J Clin Invest 88:1737-1746.

Schneider MD, Parker TG: 1991. Cardiac growth factors. Prog Growth Factors Res 3:1-26.

Nagai R, Zarain-Herzberg A, Brandl CJ, et al.: Schneider MD, Roberts R, Parker TG: 199 1. 1989. Regulation of myocardial Ca2+- Modulation of cardiac genes by mechanical

stress: the oncogene signalling hypothesis. Mel Biol Med 8: 167-183.

Schwartz K, de la Bastie D, Bouveret P, Oliviero P, Alonso S, Buckingham ME: 1986. a-Skeletal actin mRNAs accumulate in hypertrophied adult rat hearts. Circ Res 59:551-555.

Simpson PC, Long CS, Waspe LE, Henrich CJ. Ordahl CP: 1989. Transcription of early developmental isogenes in cardiac myocyte hypertrophy. J Mol Cell Cardiol 21(Suppl 5):79-89.

Snoeckx LHEH, Contard F, Samuel JL, Ma- rotte F, Rappaport L: 1991. Expression and cellular distribution of heat-shock and nu- clear oncogene proteins in rat hearts. Am J Physiol261:H1443-H1451.

Tapscott SJ, Weintraub H: 1991. MyoD and the regulation of myogenesis by helix-loop- helix proteins. J Clin Invest 87:1133-l 138.

Wankerl M, Bohm M, Moran0 I, Rtiegg JC, Eichhorn M, Erdmann E: 1990. Calcium sensitivity and myosin light chain pattern of atrial and ventricular skinned cardiac fibers from patients with various kinds of cardiac disease. J Mol Cell Cardio122: 1425- 1438.

Weber KT, Brilla CG: 1991. Pathological hypertrophy and cardiac interstitium: fi- brosis and renin-angiotensin-aldosterone system. Circulation 83:1849-1865.

TCM

Why wait for it to circulate? Read your own copy of

ardi6Kscular . MEDICINE.

If you share TCM with others at work, you know what goes around doesn’t always come around. And if it does, articles are often dated or sometimes clipped.

Why miss a single issue, or wait weeks to read one? By starting your personal subscription, for just $60 you’ll be guaranteed to read each and every issue.

TRENDS IN CARDIOVASCULAR MJXDICINE’s short reviews on ‘hot” topics can make a difference in your research and practice. Commissioned researchers and clinicians report on advances in cardiology and its related areas.

Use the business reply cards in this issue to order your personal subscription today-r call Elsevier Customer Service at:

(212)633-3950 or FAX (212)633-3990

182 01992, Elsevier Science Publishing Co., 1050-1738/92/$5.00 TCM Vol. 2, No. 5, 1992