• Non ci sono risultati.

1 7

N/A
N/A
Protected

Academic year: 2022

Condividi "1 7"

Copied!
17
0
0

Testo completo

(1)

17 Energy Metabolism

in the Normal and Diseased Heart

ARTHUR H. L. FROM, MD AND ROBERT J. BACHE, MD

CONTENTS

INTRODUCTION

MYOCARDIAL BLOOD FLOW

INTERMEDIARY METABOLISM AND BIOENERGETICS IN THE NORMAL HEART METABOLISM IN ABNORMAL MYOCARDIUM

SUMMARY

ACKNOWLEDGMENT REFERENCES

1. I N T R O D U C T I O N

Although this chapter is devoted to a discussion of myocar- dial metabolism, it should be understood that all energy-requir- ing processes in the biosphere are dependent on capture of solar energy and its entrapment in molecules that can be used to fuel biological processes such as cardiac contraction. The synthesis by plants of molecules such as glucose from CO2 and H20 is powered by the sun via the process of photosynthesis. Pho- tons radiated by the sun are trapped by chlorophyll and used to drive the synthesis of adenosine triphosphate (ATP). ATP has two phosphate bonds with high levels of stored chemical energy (Fig. 1).

The energy released by breakdown of these high-energy phosphate bonds is captured and used to drive the synthesis of glucose. Hence, a portion of the energy derived from the sun is ultimately stored in the form of chemical bonds resident in the chemically stable glucose molecule. The chemical energy stored in glucose can be released in a controlled fashion by enzyme-catalyzed reactions to drive the synthesis of other species of molecules, including fatty acids (another conve- nient storage molecule for chemical energy), as well as for resynthesis of ATP.

Because animals are unable to convert solar energy to a storable form of chemical energy, all animal life is ultimately

From: Handbook of Cardiac Anatomy, Physiology, and Devices Edited by: P. A. laizzo © Humana Press Inc., Totowa, NJ

powered by the photosynthetic process in plants. The ingestion of plants supplies animals with usable forms of stored chemi- cal energy. Even carnivores ultimately depend on energy stor- age compounds that have passed up the food chain from plants.

The general aim of this chapter is to discuss processes involved in the transfer of chemical bond energy contained in the ingested energy storage molecules that we call food into ATP, which is the common energy currency of the heart and all other biological tissues.

In the myocardium, as in other biological tissues, most energy-requiring processes use ATP as the immediate source of energy. Hydrolysis of the terminal phosphate bond of ATP releases energy that can be captured and used to drive energy- requiring processes, such as protein synthesis, muscle contrac- tion, and ion transport. The ability of the heart to pump blood to the pulmonary and systemic circulations is dependent on the availability of ATP. Because tissue stores of ATP are very modest, continuous synthesis is required to maintain ATP lev- els sufficient to drive energy-requiring processes. The energy required for ATP synthesis is derived from the controlled breakdown of the chemical bonds in carbohydrates (glucose) and fatty acids.

The purposes of this chapter are to: (1) describe mecha- nisms for delivery of carbon substrate and oxygen to the heart;

(2) describe the biochemical pathways that transfer chemical bond energy in carbon substrates to ATP in the normal heart;

and (3) discuss some of the alterations in these processes that 223

(2)

224 PART II1: PHYSIOLOGY A N D ASSESSMENT / FROM A N D BACHE

N H 2

H 0 - - "

o o

I i i

CH2-O--P-- O--P--O--P--O-

tl ~ it ~ It

l ~ / I 0 0 0

H O O H

Fig. 1. Adenosine triphosphate (ATP) is the major source of chemical energy used to power the reactions that support contractile and other processes in myocardium and all other tissues. The ~ symbol is used to designate a phosphate bond with a very high level of stored chemi- cal energy. Although ATP contains two ~ phosphate bonds, only hy- drolysis of the terminal phosphate bond releases the energy that directly powers cellular processes.

occur in the diseased heart. It should be understood that this relatively brief s u m m a r y of myocardial metabolism is of necessity a superficial overview of the individual topics dis- cussed. A list of references (primarily current topical reviews and research reports, plus several standard texts) is provided at the end of the chapter for more detailed review of the major topics presented here.

2. MYOCARDIAL BLOOD FLOW

Blood containing carbon substrate and oxygen is delivered to the heart by two main coronary arteries that originate from the proximal aorta and course over the surface of the heart (epi- cardium). These arteries arborize into progressively smaller branches that turn inward to penetrate the epicardium and sup- ply blood to the myocardium (Fig. 2). In the left ventricle, the heart muscle is typically subdivided into transmural layers;

these are termed the subepicardium (outer layer), the midmyo- cardium, and the subendocardium (innermost layer).

The coronary arterial tree terminates in muscular vessels 6 0 - 150 ~tm in diameter termed arterioles. The arterioles are the major locus of resistance to blood flow (MBF), and contraction or relaxation of the smooth muscle in the walls of the arterioles (vasomotion) provides the mechanism for control of the rate of blood flow into the myocardium.

Each arteriole supplies an array of capillaries, thin-walled tubes comprised of a single layer of endothelial cells, across which most of the exchange of nutrients, oxygen, and metabolic waste products occurs. At their terminal end, the capillaries coalesce into venules, the initial component of the cardiac venous system that conducts blood back into the venous circu- lation primarily through the coronary sinus, which drains into the right atrium.

2.1. Regulation of Myocardial Blood Flow

Coronary blood flow is closely regulated in response to changes in energy requirements of the myocardium. The prin- cipal work of the heart is muscle contraction, which generates the pressure that drives blood through the arterial system of the body. Elevation of cardiac ATP expenditure during exercise

or other stresses increases myocardial demand for oxygen and carbon substrate. Often, there is a need to assess the effects of changes of the rate of myocardial energy expenditure on cardiac performance (e.g., during exercise stress testing).

Because routine measurement of myocardial oxygen consump- tion

(MVO2)

is not practical, a rough estimate of the change in the energy demands of the heart can be obtained as the product of systolic blood pressure multiplied by the times per minute that pressure generation occurs (heart rate), termed the rate- pressure product. This measurement provides a simple esti- mate of changes in metabolic requirements of the heart in clinical situations.

The coronary circulation operates on the principle of "just- in-time" delivery of oxygen and carbon substrate. In other words, coronary blood flow is regulated to be only minimally greater than required to meet the ambient metabolic demands of the heart. As a result, the heart extracts 70-80% of the O2 from the blood as it flows through the coronary capillaries. Because of this high level of basal oxygen extraction, there is little ability to increase oxygen uptake by increased extraction of oxygen from the blood. As a result, increases in myocardial energy requirements during exercise or other stress must be satisfied by parallel increases of coronary blood flow. Because ATP and oxygen stores in the myocardium are very low, the response time for the increase in coronary flow during an increase in cardiac work must be rapid, on the order of a few seconds. From these considerations, it is clear that highly responsive signaling systems must exist between myocardial metabolic processes and vasomotor activity in the resistance vessels that control coronary blood flow. To date, the nature of these signaling systems is not totally clear despite intense study over the last 75 years.

2.2. Biological Signals

Regulating the Coronary Circulation

The regulatory signals can be classified as having feedback or feed-forward characteristics; the final common response to these signals is relaxation of the vascular smooth muscle cells that make up the resistance vessels that control coronary blood flow (Fig. 3). Several major feedback mechanisms resulting from increased cardiomyocyte metabolism (including adenos- ine, nitric oxide [NO], and other less-defined signals) cause opening of ATP-sensitive potassium channels (KATe) on the sarcolemma of smooth muscle cells of the coronary arterioles.

Opening of these channels allows potassium to escape from the cytosol of the smooth muscle cells, resulting in hyperpolariza- tion (increased negativity) of the cell membrane. The increased negativity of the m e m b r a n e causes sarcolemmal voltage- dependent calcium channels to close; as a result, calcium entry is reduced, the muscle relaxes (vasodilation), and coronary blood flow increases. In addition to effects on the KAT e chan- nels, adenosine (a product of ATP utilization in the cardio- myocyte) has potent, direct dilator effects on arteriolar smooth muscle.

Another feedback signal is NO generated by the vascular endothelium. Mechanotransduction of flow-induced shear forces exerted on the endothelial cells augments NO synthesis.

In addition to causing potassium channel opening, NO also

(3)

Intramural arteries Epicardial artery Class B Class A

Epicardium

LV wall

EndocardiL

Subendocardial plexus

Fig. 2. The transmural distribution of the coronary arterial system is depicted. The large conductance arteries traversing the epicardial surface supply shallow and deep branches to the subepicardium (outer-most myocardial layers) and subendocardium (inner-most myocardial layers), respectively. These perforating vessels arborize to create the arteriolar network that supplies the myocardial capillary bed. LV, left ventricle.

Reprinted from D.J. Duncker and R.J. Bache, Regulation of coronary vasomotor tone under normal conditions and during acute myocardial hypoperfusion, PharmacoI Ther., 86, 87-110, © 2000, with permission from Elsevier.

Major Mediators of Metabolic Feed-back Vasodilation

| . KAT I, channel 2. Adenosine 3. NO

Mediator of Metabolic Feed-forward Vasodilation 1. Increased Sympathetic Nerve NE Discharge

~adrenergic receptor stimulation of resistance vessels

Fig. 3. Major feedback and feed-forward mechanisms underlying metabolic vasodilation of resistance vessels are depicted. See text for discussion. NO, nitric oxide; KAT P channel, adenosine triphosphate-inhibited potassium channel; NE, norepinephrine.

directly relaxes vascular smooth muscle. Importantly, this brief discussion does not include all of the known feedback mecha- nisms involved in regulation of coronary blood flow.

Increases of cardiac sympathetic nerve activity, such as during exercise, activate a feed-forward mechanism for con- trol of coronary blood flow that augments the local metabolic vasodilator influences. The sympathetic neurotransmitter norepinephrine activates c~- and [3-adrenergic receptors in vas- cular smooth muscle. Activation of c~-adrenergic receptors causes modest constriction of the large coronary arteries; how- ever, because these arteries function as conduit vessels that offer little resistance to blood flow, this has little effect on coronary flow. However, activation of [3-adrenergic receptors on the coronary arterioles results in relaxation (vasodilation)

of these resistance vessels; the resultant decrease of coronary resistance causes an increase in blood flow that is not depen- dent on local metabolic mechanisms for regulation of coro- nary vasomotor tone.

Pharmacological studies have shown that simultaneous blockade of the coronary KAT P channels, adenosine, and NO pathways significantly decreases myocardial blood flow in the resting animal. Moreover, the increase of coronary flow that normally occurs during exercise is severely blunted, resulting in a perfusion-metabolism mismatch that is accompa- nied by evidence of ischemia in the normal heart. Hence, acti- vation of these three pathways appears to be the primary means by which metabolic vasodilation is achieved in the heart. How- ever, in the normal situation, blockade of any one of these

(4)

226 PART II1: PHYSIOLOGY A N D ASSESSMENT / FROM A N D BACHE

mechanisms for smooth muscle relaxation elicits compensa- tory activation of the other pathways to minimize changes in coronary blood flow.

2.3. Blood Flow in the Diseased Heart

Coronary blood flow in the diseased heart can be limited by:

(1) partial or complete obstruction of the large coronary arteries (e.g., as in atherosclerotic disease); (2) decreased responsive- ness of the signaling systems relating MBF to myocardial energy requirements; or (3) increases in extravascular forces acting to compress the small vessels in the wall of the left ven- tricle. In the case of obstructive coronary disease, a moderately narrowed vessel may functionally restrict blood flow only during periods of increased work (i.e., it reduces vasodilator reserve); a severely narrowed vessel may limit blood flow even when the subject is at rest. In the presence of a moderate coro- nary obstruction, the arteriolar bed can maintain adequate blood flow by metabolic signaling-based arteriolar vasodilation. That is, a decrease in small-vessel resistance can compensate for the resistance offered by the coronary artery stenosis. However, when the capacity for vasodilation of the arterioles has been exhausted, any further increase in cardiac work cannot result in an increase of blood flow, and the myocardium supplied by the narrowed vessel will become ischemic. There is also some evi- dence that malfunction of metabolic signaling pathways in the arteriolar resistance vessels can aggravate (e.g., in patients with nonobstructive atherosclerotic disease or with diabetes) or cause (syndrome X) myocardial ischemia.

In the normal heart, blood flow to the inner layers of the left ventricle occurs principally during diastole. This is because tissue pressures in the wall of the left ventricle during systole are so great that inner-layer arterioles are squeezed shut by the extravascular compressive forces. Diastolic left ventricular tis- sue pressures are also greatest in the subendocardium. As the heart fails and/or becomes hypertrophied, these extravascular compressive forces increase as left ventricular diastolic pres- sure (i.e., ventricular filling pressure) increases.

Slowing of myocyte relaxation also encroaches on the inter- val of diastole in the hypertrophie d or failing heart. In the normal heart, autoregulatory (i.e., metabolic vasodilation) processes cause arteriolar vasodilation in the subendocardium to compen- sate for systolic underperfusion. However, increases of left ven- tricular diastolic pressure that occur in the failing heart may compress the arteriolar bed in the inner myocardial layers suffi- ciently to overwhelm autoregulatory mechanisms, particularly those that normally maintain adequate subendocardial blood flow. Because subendocardial blood flow occurs predominantly during diastole, tachycardia also acts to impede blood flow in the subendocardium by shortening the interval of diastole.

Thus, even in the absence of obstructive coronary artery disease, functional abnormalities can limit blood flow to the inner myocardial layers of the diseased heart. These abnormali- ties are often then associated with a reduction of the ATP syn- thetic capacity in the subendocardium. Thus, the extravascular forces that act on the small coronary vessels embedded in the myocardium cause the subendocardium to be the region of the ventricular wall that is most vulnerable to hypoperfusion and ischemia.

3. INTERMEDIARY METABOLISM

AND BIOENERGETICS IN THE NORMAL HEART 3.1. Carbon Substrate Selection

The primary carbon substrates (Fig. 4) taken up and metabo- lized by the myocardium are free fatty acids and glucose. The heart readily takes up and metabolizes pyruvate, lactate and ketone bodies, but the generally low blood concentrations of these substrates limit their utilization. However, under certain conditions such as exercise, which can markedly elevate the blood lactate content, or during periods when blood levels of ketone bodies are elevated, utilization of these substrates may be dominant. A detailed discussion of the regulation of blood levels of these substrates exceeds the scope of this chapter, but a few orienting comments are appropriate regarding glucose and free fatty acid availability.

Blood glucose levels are maintained within a narrow range (N5 mM) in nondiabetic subjects. Its homeostasis is maintained by a balance among alimentary uptake of glucose, glucose release from the liver (which can synthesize glucose or release glucose from the glycogen storage pool), and the uptake and metabolism of glucose by the various organ systems. Glucose uptake is strongly influenced by the pancreatic secretion of insulin, a hormone that enhances glucose transport into the cells of many tissues.

The heart has often been labeled as an omnivore because of its capacity to consume virtually any available carbon substrate.

Glucose is transported into the cardiomyocyte by a family of sarcolemmal glucose transport proteins, some of which are insulin dependent. Fatty acids enter myocytes via a sarcolem- mal fatty acid transport protein, and pyruvate and lactate enter via a sarcolemmal monocarboxylic acid transporter. Utilization of glucose by the heart is largely regulated by the availability of fatty acids in the blood. Thus, in the fasted state, blood fatty acid levels are high, and fatty acids are the predominant cardiac substrate despite normal blood glucose levels.

In contrast, during vigorous exercise, blood lactate levels can rise markedly and compete strongly with fatty acids and glucose despite substantial blood levels of the latter substrates.

After a carbohydrate meal, however, blood glucose levels rise and elicit insulin secretion. Insulin stimulates glucose uptake by the heart and other tissues and causes blood fatty acid levels to decrease. As a result, myocardial glucose consumption in- creases, and fatty acid consumption decreases. Nevertheless, fatty acids and lactate are the favored cardiac substrates when they are available in sufficient concentrations in the blood.

3.2. Glucose Metabolism

Figure 5 shows a flowchart for glucose metabolism. Glu- cose enters the cardiomyocyte via the sarcolemmal glucose transport proteins, GLUT1 (which is insulin independent) and GLUT4 (which is insulin dependent). Once in the cell, glucose is phosphorylated to glucose-6-phosphate by the enzyme hex- okinase. Because glucose-6-phosphate is membrane imperme- able, it is effectively trapped within the cell, where it can enter the glycogen synthesis pathway (glycogen is a macromolecu- lar polymeric storage form of glucose) or undergo molecular rearrangement via the enzyme phosphohexose isomerase to

(5)

Food intake and Intermediary Metabolism Oxidative

Conversion to Phosphorylation

Metabolic Substrates

f ~ r ~ f

I Carbohydrate (mainly glucose)

Fat

(mainly free fatty acids)

\

Gly~t°lysis Acetyl CoA B g i d oxidation

/

[ ETC + NA~DH Oxygen+

Fig. 4. A general overview of carbon substrate metabolism in the heart. Ingested food is broken down into usable carbon substrates, primarily glucose and fatty acids. The pathways that convert amino acids and other molecules to glucose are not shown. Glucose and fatty acids are processed (via intermediary metabolic processes) to yield the reducing equivalents nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FADH2), which supply the electrons to the ETC that, in turn, powers oxidative phosphorylation. The last processes, which occur in mitochondria in the presence of oxygen, supply almost all of the adenosine triphosphate (ATP) synthesized in the heart. CoA, coenzyme A; ETC, electron transport chain; TCA, tricarboxylic acid.

Sarco~mm GI~ ogen

transporlms: nmlase

Glut I and ~ u t 4) Glyogem PbospboglKo-

• ATP ADP

Glucose G l u c o s e ~ Glucose-6-

llegel~m phosphate ~ g o .~

O ATP ADP

Fructose 6 N~4{ ~ Fructose 1 6 phosphate l~esi~e~eet~ bisphosphate m n u ~ e ~ye~ral~hyde ~

l~m~glyeerate i~hesl~egl~cerate d e h y d m ~

matase Idmse

2 P h o s p h o ~ 3 P h o s p h o ~ 1 3 b i s ~ Glycer ~d ~ Dihydro glycerate glycerate ATPY~ADP p h o s p h o NADHIIt ~NAD ÷ aldehyde 3r PboSpbotriose acetone

phosphate kemerue phosphate

Smlue ~ " glycerate ~

Phosphoenoi pyruvate l yravate

P.~vate ~ ~ deky~aase

ANA;HO

(tmmatteck~adr~l)

Pymvate I - - - I ~ A T P

Aerobic glucose metabolism produces ~38 molecules of

ATP/molecule glucose including 2 from glycolysis (O). Glycolysis actually produces 4 ATP/molecule of glucose but 2 ATP molecules are consumed in the initial steps of glycolysis (O).

Fig. 5. A flowchart depicting the cellular uptake of glucose and the pathways through which glucose metabolism proceeds. The red dots indicate reactions which consume ATP and the blue dots designate reactions which produce

ATP.See

text for discussion. ADP~ adenosine diphosphate;

ATP, adenosine-triphosphate; ETC, electron transport chain; NAD, NADH, nicotinamide adenine dinucleotide; TCA, tricarboxylic acid.

(6)

228 PART II1: PHYSIOLOGY A N D ASSESSMENT / FROM A N D BACHE

A Phosphoenol-

pyruvate

eyr vate

Idnnse

I. I /

~IL~ Pyruvate

~ , ~ ~ uxalo-acetate

Pyruvate Lactate

I

dehydrogenase I Pyruvate

~

I dehydrogenase ~k

A c e t y l C o A complex

Lactate + H +

Although under aerobic conditions most pyruvate produced is converted to acetyl CoA, which enters the TCA cycle, a certain amount is carboxylated to form oxaloacetate. The latter enters the TCA cycle intermediate pool.

This is an anaplerotic pathway.

B

3-KetoacyI-CoA

Thiolase

Acyl CoA Acetyl CoA (reenters oxidation

cycle minus two carbon atoms)

or AcetylCoA

o r

Propionyl CoA

If the long chain fatty acid being metabolized has an even number of carbon atoms, then the acyl CoA produced during final cycle through [3-lipid oxidation is acetyl CoA, which enters the TCA cycle to be metabolized.

If the long-chain fatty acid being metabolized has an odd number of carbon atoms, then the acyI-CoA produced during final cycle through [Gr beta]-lipid oxidation is propionyl CoA. Propionyi CoA is converted to succinyl CoA. The latter is a TCA cycle intermediate and enters the TCA cycle pool.

This is an anaplerotic pathway.

Fig. 6. Anaplerotic processes supply substrate used to maintain the tricarboxylic acid (TCA) cycle intermediate pool size. In contrast, cataplerotic processes remove substrates from the TCA cycle intermediate pool and decrease its size. The balance between these two processes determines the TCA cycle pool size. (A) Glucose (via its metabolic product pyruvate) contribution to anaplerosis. (B) The [3-oxidation of odd-numbered carbon chain fatty acids contribution to anaplerosis. Figure 8, which is a flowchart for [3-lipid oxidation, should be reviewed before examining Fig. 6B, which depicts the terminal reaction of [3-lipid oxidation. ADP, adenosine diphosphate; ATP, adenosine-triphosphate; CoA, coenzyme A.

form fructose-6-phosphate. A second phosphorylation of fruc- tose-6-phosphate via phosphofructo-kinase generates fructose- 1,6-bisphosphate. Each of these phosphorylations consumes one molecule of ATP.

Fructose-l-6-bisphosphate is next split into glyceralde- hyde-3-phosphate and dihydroacetone phosphate by the enzyme aldolase. These two products are in constant exchange with each other via the enzyme phosphotriose isomerase.

Glyceraldehyde-3-phosphate is phosphorylated to form 1,3- bisdiphos-phoglycerate by the enzyme glyceraldehyde-3- phosphate dehydrogenase. The phosphate bond in the one position has a high-energy content (signified by the ~P sym- bol); this reaction also simultaneously reduces oxidized nico- tinamide adenine dinucleotide (NAD ÷) to its reduced form (NADH). Cytosolic oxidation of NADH and transport of the two removed electrons and H ÷ into the mitochondrial matrix then occurs via the malate-aspartate shuttle (not shown in Fig. 5). In the mitochondrial matrix, NAD ÷ is reduced to

NADH which can be utilized by the mitochondria to generate ATP

(see

Section 3.8.).

In the next reaction, the high-energy phosphate bond in 1,3- bisdiphosphoglycerate is transferred to adenosine diphosphate (ADP) to form ATP and 3-phosphoglycerate via the enzyme phosphoglycerate kinase. The latter molecule is converted to 2- phosphoglycerate by the enzyme phosphoglycerate mutase.

Enolase, another cytosolic enzyme, then converts 2-phospho- glycerate to the -P-containing molecule phosphoenolpyruvate, which is converted to pyruvate by the enzyme pyruvate kinase.

During this reaction, the ~P in phosphoenolpyruvate is trans- ferred to ADP to form a second ATP molecule. Pyruvate can then enter the mitochondria to be metabolized further by the pyruvate dehydrogenase (PDH) complex of enzymes. Both the PDH complex, which converts pyruvate to acetyl CoA (coen- zyme A) (and NAD ÷ to NADH), and the tricarboxylic acid (TCA) cycle that metabolizes acetyl CoA are located within the mitochondria (Figs. 5, 6A, and 7). Within the mitochondria,

(7)

another metabolic pathway for pyruvate exists. The enzyme pyruvate carboxylase converts pyruvate to oxaloacetate, which is a TCA cycle intermediate ( Figs 6A and 9). The significance of the latter reaction is discussed in the legend of Fig. 6.

Within the cytosol, pyruvate can be converted to lactic acid by lactic acid dehydrogenase (LDH). Lactate is then exported from the cell via the monocarboxylic acid transporter (Figs. 5 and 13). This reaction results in the oxidation of NADH to NAD +, and as will be shown in Section 4.1.1., this reaction is critical when the availability of oxygen to the cardiomyocyte is limited. Conversely, under aerobic conditions, lactate can be transported into the cytosol by the monocarboxylic acid trans- porter to be converted to pyruvate by LDH. The LDH-catalyzed reaction also reduces NAD + to NADH, and the reducing equiva- lents from NADH can be transferred into the mitochondrial matrix by the malate-aspartate shuttle.

During glycolysis of one glucose molecule, two pyruvate molecules, four ATP molecules, and two NADH molecules are produced. However, because two ATP molecules are consumed early in the glycolytic pathway, the net production of ATP is two molecules/glucose molecule. As indicated, pyruvate and NADH are utilized in the mitochondria for oxidative generation of ATP. Therefore, complete metabolism of glucose, which includes oxidation of the products of glycolysis, results in for-

lntermemhrane Snace Outer

Mitochondrialis

Cytoplasm

Fig. 7. The major features of mitochondrial morphology. See text for discussion.

mation of many more ATP molecules (38) than does glycolysis alone (2). However, when mitochondrial function is severely oxygen limited, the oxidative contribution to ATP synthesis is lost, and there is a net production of two ATP molecules from each glucose molecule passing through the glycolytic pathway.

Triglyceride

Lipoprotein Hpase

~ Giycerol

Free Fatty Acids (FFA)

4 " " Sarcolemma

Acyl-CoA

synthetase

F F A ~

Acyl CoA

ATP" "ADP+ Pi I CPT 1

Mitochondrial ~ I CACT inner membrane ' ,~¢ CPT 2

Acyl CoA Aql CoA ~ FAD dehydrogmtase

ddtydrogenase3"Hydr°xyacyl CoA F~Hz

3-Ketoacyl CoA ~ 3-Hydroxyacyl CoA -~ A2-transenoyl-CoA

i " A2-trans-mtovl

T h i o l a s e ~

Acyl CoA 0 Acet (reenters oxidation cycle

minus 2 carbon atoms).

Also See Figure 6B. O = Products of [5-lipid oxidation

Fig. 8. Flowchart depicting the cellular uptake of free fatty acids and the pathways through which their metabolism proceeds. Blue dots indicate products of B-lipid oxidation. See text for discussion. ATP, adenosine triphosphate; CoA, coenzyme A; CPT, carnitine palmatoyl-transferase;

CACT, carnitine-acylcarnitine transporter: ETC, electron transport chain; FADH 2, flavin adenine dinucleotide; FFA, free fatty acids; NADH, nicotinamide adenine dinucleotide; TCA, tricarboxylic acid.

(8)

230 PART IIh PHYSIOLOGY AND ASSESSMENT / FROM AND BACHE

Acetyl.CoA

Malate ~ C

Dehydrogemuz _ Oxaloacetate Itrate Synthase NADH

L-Malate' NAD + Citrate

Fumarate Sllo~inll~ r FADEI2 • l)¢hydrogmase / ~ FAD

Succinate

SGOI~IIIII~e DP + Vi

"lltiokinase

¢onltam

Cis-aconitate AconlU lsocitrate

~~ Imcltrate

NADH D ~ y ~

Succinyl-CoA Oxalosuccinate

• eco2

D e h y d r o g c m ~ )" c t - K e t o g l u t a r a t e "

l)¢hydrogmam Compi~ co~2

FADH z NADH

E T C +

Oxygen • A T P

• =

Products of TCA cycle

Fig. 9. The tricarboxylic acid (TCA) cycle is depicted. The filled blue circles are the products resulting from one turn of the TCA cycle.

See text for discussion. ATP, adenosine triphosphate; CoA, coenzyme A; ETC, electron transport chain; FAD, oxidized flavin adenine dinucleotide; FADH2, reduced flavin adenine dinucleotide; GDP, guanosine 5'-diphosphate; GTP, guanosine 5'-triphosphate; NAD ÷, oxidized nicotinamide adenine dinucleotide; NADH, reduced nicotinamide adenine dinucleotide; Pi, inorganic phosphate.

3.3. The Mitochondrion

Mitochondria, which are the primary site of ATP synthesis in most mammalian cells, contain the [3-lipid oxidation enzymes, the TCA cycle enzymes, the electron transport chain (ETC), and the FIF0-H*-ATPase (also called F1F0-ATP syn- thase or ATP synthase). To understand better the location of these systems, mitochondrial structure is briefly reviewed (Figs. 7,10, and 11). The inner mitochondrial membrane con- tains the ETC, FIF0-H+-ATPase, adenine nucleotide trans- locase, and other transporters. The matrix contains the TCA cycle enzymes, 13-lipid oxidation enzymes, and other enzymes and reactants. The cristae are invaginations that markedly expand the surface area of the inner membrane and thereby the contents of enzymes associated with the inner membrane.

The membrane-bound components of the enzyme pathways are positioned to optimize the flow of substrates through their reaction sequences. The mitochondrial outer membrane forms a boundary between the cellular cytoplasm and the mitochondrial intermembrane space. The intermembrane space contains creat-

ine kinase, which is important to high-energy phosphate trans- port out of mitochondria (Fig. 12) and cytochrome-c, a compo- nent of the ETC. The importance of the intermembrane space to oxidative ATP synthesis is discussed in Sections 3.6. and 3.7.

3.4. Fatty Acid Metabolism

Figure 8 presents a flowchart for the [3-lipid oxidation path- way. Dietary long-chain, nonesterified, free fatty acids (NEFAs) are usually the predominant cardiac substrate. They are transported in the blood bound to plasma albumin or in the form of triacylglycerol, which is also bound to albumin.

The latter can be broken down to release NEFA by an enzyme present in the plasma and at the surface of the capillary and cardiomyocyte.

After dissociating from albumin, NEFAs are transported into the cardiomyocyte by sarcolemmal fatty acid transport proteins.

Within the cell, NEFAs are bound to fatty acid-binding pro- teins, which provide solubility. Once in the cell, NEFA mol-

(9)

Outer Membrane

lntermembrane Space

Matrix

Fig. 10. The electron transport chain (ETC) is depicted. See text for discussion. FADH2, flavin adenine dinucleotide; NAD, NADH, nicotina- mide adenine dinucleotide, FAD, oxidized flavin adenine dinucleotide; CoQ, coenzyme Q; Cyt c, cytochrome c.

ecules are either reesterified and stored as triglycerides or acti- vated by acyl CoA synthetase (in the presence of free CoA and ATP) to form a long-chain acyl CoA.

Because long-chain acyl CoA cannot readily diffuse through the mitochondrial inner membrane, the long-chain acyl CoA is converted to long-chain acyl carnitine by carnitine palmatoyl- transferase 1 (CPT1) at the outer surface of the inner mito- chondrial membrane and then transported across the inner membrane by carnitine-acylcarnitine translocase (which also transports free carnitine liberated by carnitine palmatoyl-trans- ferase 2 [CPT2] back into the intermembrane space).

The long-chain acylcarnitine is next converted back to acyl CoA at the inner surface of the mitochondrial inner membrane by CPT2. Long-chain acyl CoA is then processed by a sequence of enzyme-catalyzed reactions that comprise the [3-lipid oxida- tion pathway. If the long-chain acyl CoA has an even number of carbon atoms, then the final products of the last cycle of an acyl CoA molecule through the [3-oxidation sequence are two acetyl CoA molecules (Fig. 6B). However, if the long- chain acyl CoA has an odd number of carbon atoms, then

the

products of the last cycle through [5-oxidation are one acetyl CoA and one propionyl CoA. Unlike acetyl CoA, propionyl CoA cannot enter the TCA cycle.

Propionyl CoA can be converted to succinyl CoA, which is a TCA cycle intermediate (Fig. 6B). Hence, this pathway con- tributes to the maintenance of the TCA cycle intermediate pool size; both pyruvate molecules produced from glucose and odd- numbered fatty acid molecules that undergo complete [5-oxida- tion contribute to maintenance of

the

TCA intermediate pool.

Processes that add molecules to the TCA intermediate pool are termed

anaplerotic

(Fig. 6A,B), and those that remove interme-

diates from the pool are called

cataplerotic.

The importance of these processes to TCA cycle function is illustrated in a clinical example presented in Section 4.3.

The products of complete [5-oxidation of a fatty acid are acetyl CoA (and propionyl CoA if the acyl chain has an odd number of carbons), NADH, and reduced flavin adenine dinu- cleotide (FADH2). Yet, consumption of acetyl CoA by the TCA cycle and NADH and FADH 2 by the ETC cannot occur in the absence of oxygen. Thus, [5-oxidation also cannot occur under anoxic or ischemic conditions. Therefore, markedly ischemic myocardium does not support fatty acid-derived ATP synthe- sis, and the only remaining source of ATP synthesis is anaero- bic glycolysis.

3.5. The TCA Cycle

The carbon substrate consumed by

the

TCA cycle is acetyl CoA. Acetyl CoA is produced by glycolysis and [5-lipid oxi- dation as described in Sections 3.2. and 3.4., respectively.

Figure 9 shows the individual reactions of the TCA cycle. In the first step of the cycle, acetyl CoA contributes two carbon atoms to oxaloacetate to form citrate and free CoA. This reac- tion is catalyzed by the enzyme citrate synthase. Citrate then enters a sequence of reactions that ultimately generate two molecules of COz and four reducing equivalents in the form of NADH (3) and FADH2 (1). One high-energy phosphate mol- ecule (guanosine 5'-triphosphate) is also produced, which can be converted to ATP. During this process, citrate (a six-carbon molecule) is stepwise decarboxylated and ultimately con- verted back to oxaloacetate, a four-carbon molecule that, when condensed with a new acetyl CoA, reinitiates the sequence of reactions.

(10)

232 PART II1: PHYSIOLOGY A N D ASSESSMENT / FROM A N D BACHE

Intermembrane Space Inner H ÷

Meibrane l

ETC Complexes

H ÷ ADP ATP

ATP

ADP + Pi ATP

H ÷ Mitochondrial

Matrix

A D P ATP

Fig. 11. The mechanism of oxidative phosphorylation and the transport of adenosine triphosphate (ATP) from the mitochondrial matrix to the intermembrane space are depicted. See text for discussion. ADP, adenosine diphosphate; Pi, inorganic phosphate; ANT, adenine nucleotide transporter; ETC, electron transport chain.

The reducing equivalents generated, NADH and FADH2, deliver electrons to the ETC. Rate-limiting enzymes of the TCA cycle are the PDH complex (which precedes, but is not really a component of the TCA cycle), isocitrate dehydrogenase, and c~- ketoglutarate dehdrogenase; these enzymes are highly regulated, as discussed in Section 3.7. Interestingly, for elucidating the TCA cycle (also known as the Krebs cycle), Sir Hans Krebs was awarded the Nobel Prize in Medicine or Physiology in 1953.

3.6. The Electron Transport Chain and Oxidative Phosphorylation

The substrates of the ETC are NADH and FADH2. The ETC (Fig. 10) is composed of two freely diffusible compounds (ubiquinone, also known as coenzyme Q, which is confined to the mitochondrial inner membrane, and cytochrome-c, which is located in the intermembrane space) and four functional com- plexes (I, II, III, and IV), which are contained within the mito- chondrial inner membrane. The complexes are composed of catalytic and non-catalytic proteins (and their cofactors).

NADH interacts with the ETC by transferring electrons to (and thereby reducing) complex I, which in turn then reduces ubiquinone. FADH 2 interacts with the ETC by transferring elec- trons to complex II, which, like complex I, also transfers its electrons to ubiquinone. Reduced coenzyme Q then diffuses to complex III and transfers its electrons. Next, complex III re- duces cytochrome-c, which diffuses to complex IV and trans- fers electrons to this complex. The latter, when fully reduced by four electrons, reduces O 2 to two O -2 ions. Each 0 -2 combines with two H + to form H20 in an irreversible reaction.

During the process of electron transport, electrons release energy as they pass through the sequence of complexes. The purpose of the ETC is to capture this free energy and use it to pump H + from the mitochondrial matrix across the inner mem- brane into the intermembrane space. In this regard, complexes I, III, and IV function as proton pumps (Figs. 10 and 11). These pumps create an electrochemical gradient (A~tH +) composed of an electrical potential and a chemical potential (the latter reflected by ApH) across the inner mitochondrial membrane.

As a consequence of proton transport, the mitochondrial matrix is more negative than the intermembrane space. The FIF0-ATPase, which is also located in the inner mitochondrial membrane (Fig. 11), is the major pathway of H + return from the intermembrane space into the matrix. The energy released by the passage of H + down this electrochemical potential gradient is captured by the F~F0-ATPase and used to drive the reaction Pi + ADP --) ATP (Pi is inorganic phosphate). In this way, much of the energy released by the metabolism of glucose and fatty acids is transferred to the terminal high-energy phosphate bond of ATP.

The concept of proton pumping into the intermembrane space, and the use of the electrochemical gradient thus formed to synthesize ATP, is known as the chemiosmotic theory. In recognition of this (initially quite controversial) insight, Sir Peter Mitchell was awarded the Nobel Prize in Chemistry in 1978. The production of ATP by mitochondria is termed oxida- tive phosphorylation; in virtually every human tissue (heart, brain, kidney, etc.), oxidative phosphorylation is the primary source of ATP generation.

To date, the precise mechanisms by which the F1F0-H +- ATPase generates ATP from its substrates ADP and Pi remain under investigation. However, a number of characteristics of the process are well established. The F1F0-H+-ATPase has been shown to be a near-equilibrium enzyme. In other words, the reaction can occur in both directions. However, both in iso- lated mitochondria that are generating ATP under conditions of excess carbon substrate, ADP, and oxygen and also in per- fused and in vivo hearts, the F~F0-H+-ATPase operates far out of equilibrium, so virtually all fluxes are in the ADP + Pi -+

ATP direction. In the presence of abundant oxygen, the F~F0- H+-ATPase is kinetically controlled by the concentrations of its immediate substrates ADP and Pi, which are not usually limiting, and the magnitude of the proton electrochemical potential gradient across the mitochondrial inner membrane.

Hence, ADP levels alone can regulate the rate of ATP synthe- sis so long as Pi, oxygen and carbon substrates are not limiting and the resulting proton gradient is very large. However, in

(11)

vivo regulatory mechanisms are far more complicated than those described by this simple scheme.

3.7. Regulation of Oxidative Phosphorylation

As pointed out, a simple kinetic regulatory scheme for oxi- dative phosphorylation with A D P and Pi availabilities control- ling the A T P synthetic rate is valid when both oxygen and reducing equivalents are in excess. Under these conditions, the high proton electrochemical gradient will drive A T P synthesis until A D P falls to levels that are rate limiting. From this point, the rate of A D P production determines the steady-state rate of A T P synthesis. Because A D P is a product of A T P hydrolysis, the steady-state rate of A T P synthesis is determined by the steady-state rate of A T P hydrolysis (i.e., the rate of A T P utili- zation) and the steady-state A D P level required to support that rate of A T P synthesis.

This simple feedback regulation of A T P synthesis can be demonstrated in the perfused rat heart if the perfusate contains unphysiologically high concentrations of carbon substrates such as pyruvate or octanoate. Under these experimental con- ditions, myocardial A D P concentrations become low enough to regulate oxygen consumption kinetically; an increase in the rate of A T P utilization will cause A D P levels to rise, which in turn stimulates an increase in the rate of oxygen consumption. In this construct, the increased rate of delivery of A D P to the F~F0- ATPase increases the rate of A T P synthesis as long as A D P levels rise to values capable of supporting the new higher rate of A T P synthesis. In other words, A D P levels must increase concordant with the rate of A T P synthesis. This type of regula- tion is well described by the single-substrate Michalis-Menten model.

H o w e v e r , in large animal hearts (e.g., dog, swine, and human) in vivo A D P levels are higher than in the isolated perfused rat heart and are well above the usual kinetic regula- tory range described in studies of isolated mitochondria or perfused rat hearts. Further, a s M V O 2 increases, A D P levels actually remain fairly constant in the in vivo large animal heart. This does not mean that A D P availability is not crucial to the rate of A T P synthesis. W h e n the rate of A T P expendi- ture is increased, there must always be an identical increase in the rate of A D P delivery to mitochondria (once a new steady state is reached). This increase in the rate of A D P delivery to the A T P synthase is obligatory, but a change in steady-state A D P levels is not.

The observation that A D P levels did not increase in the in vivo m y o c a r d i u m as the rate o f A T P synthesis increased led to the recognition that additional regulatory systems were required to explain the data in in vivo models. Further, the over- all regulatory processes must have very rapid response times regarding facilitation o f A T P synthesis. The last is required because myocardial contents of A T P and creatine phosphate (PCr; another high-energy phosphate bond storage molecule that serves to buffer A T P levels) are only sufficient to support a few seconds of A T P expenditure in the absence of A T P syn- thesis. The reaction converting cytosolic A T P to PCr is rapid and near equilibrium via the e n z y m e creatine kinase. Figure 12 provides a brief overview of the Np and A D P transport function of PCr.

Contrm.-tile processes

ATP

Cytosol

ATP ADP

PCr Cr

Membrane °-t

PCr

Inter-membrane ~

space

ADP A' 'P

Inner

Membrane

Matrix

ANT

ADP ATP

Fig. 12. The creatine kinase shuttle (CKS) hypothesis is depicted. The diffusion rate of adenosine diphosphate (ADP) across the outer mito- chondrial membrane and through the cytosol is relatively slow. The CKS (and other shuttle systems not described; see source at end of this chapter) is thought to facilitate the transfer of ADP from sites of adenosine triphosphate (ATP) utilizing reactions to the mitochondrial intermembrane space and to facilitate the transfer of ATP to the sites of utilization. The way the shuttle is proposed to function is as follows:

The adenine nucleotide transporter (ANT), which is not rate limiting, transports ADP from the mitochondrial intermembrane space to the mitochondrial matrix, where it is rephosphorylated back to ATP.

The ANT simultaneously transports newly synthesized ATP from the matrix to the intermembrane space. Within the intermembrane space, the mitochondrial isozyme of creatine kinase (MCK) transfers the terminal high-energy phosphate of ATP to creatine (Cr) to form phos- phocreatine (PCr). PCr then diffuses into the cytosol, and the ADP produced by this reaction in the intermembrane space is returned to the mitochondrial matrix by ANT, where it is rephosphorylated. The PCr that has diffused into the cytosol is then used as a high-energy phos- phate donor for the rephosphorylation of cytosolic ADP by the cyto- solic isozyme of creatine kinase (CCK). Because CCK has an extremely rapid turnover rate, it can facilitate transport of ATP from the mitochondria to the site of its hydrolysis and of ADP from the hydrolysis site back to the mitochondrial outer membrane. There, PCr leaving the mitochondria is used to rephosphorylate ADP, and the Cr liberated from PCr diffuses into the mitochondrial inner membrane space to be rephosphorylated. As a result, cytosolic ADP levels (including those in proximity to the points of ATP utilization) are kept at low levels even if the rate of ATP utilization increases. The stability of ADP levels in the face of increasing ATP utilization permits ATP to maintain a high level of free energy that can be transferred to energy requiring cellular processes.

(12)

234 PART Ill: P H Y S I O L O G Y A N D ASSESSMENT / FROM A N D BACHE

A discussion of additional regulatory mechanisms for oxida- tive phosphorylation is presented next, and a review of the dis- cussion of cardiomyocyte contraction mechanisms presented in Chapter 8 will refresh understanding of cardiomyocyte Ca >

dynamics relevant to the following discussion.

During exercise, increased norepinephrine release by the sympathetic nervous system activates [3-adrenergic receptors in cardiomyocytes. Their activation causes an increase of heart rate (by acting on the sinoatrial node) and an increase of Ca -,+

entry into the cytosol of the cardiomyocytes via the sarcolem- mal voltage-dependent Ca 2+ channel. The increased Ca 2÷ enter- ing the cytosol, together with [3-adrenergic receptor-stimulated activation of sarcoplasmic reticulum Ca 2+ pumping, increases sarcoplasmic reticulum Ca 2÷ stores. The more fully loaded sar- coplasmic reticulum can then release more Ca 2+ per beat into the cytosol, resulting in a larger systolic cytosolic Ca > transient.

Both the increased frequency (heart rate) and the larger Ca 2÷ transient size serve to increase the "average" Ca 2+ level in the cytosol. Mitochondria normally transport Ca 2+ in and out of the matrix and maintain a "steady-state" matrix Ca 2+

level that is related to average cytosolic Ca -'+. In response to higher average cytosolic Ca > levels, mitochondrial Ca 2+

uptake increases, resulting in an increase in the steady-state level of matrix Ca -,+ .

Increased mitochondrial matrix Ca 2+ has at least two impor- tant effects on oxidative phosphorylation. First, as mentioned, the TCA cycle has three rate-limiting enzymes (PDH, isocitrate dehydrogenase, and et-ketoglutarate dehydrogenase), and the activities of these enzymes are partly regulated by mitochon- drial matrix Ca 2÷ levels. When these enzymes interact with Ca 2+, their sensitivities to their respective substrates are increased.

This accelerates their reaction rates and increases the rates of acetyl CoA entry into and flux through the TCA cycle without requiring pyruvate, acetyl CoA levels, or TCA cycle pool inter- mediate levels to rise, provided the rate of delivery of acetyl CoA to the TCA cycle is increased sufficiently to support the increased TCA cycle flux rate. Therefore, glucose uptake and glycolysis-mediated delivery of pyruvate to, and its rate of metabolism by, PDH must increase or the rates of fatty acid uptake by the cell and [3-oxidative delivery of acetyl CoA to the TCA cycle must increase to support the increased rate of TCA cycle flux. In practice, the rates of cellular substrate uptake, glycolytic flux, and [3-0xidation all respond to metabolic sig- naling, and the required increases in the rates of acetyl CoA production occur.

A major consequence of an increased TCA cycle flux is more rapid generation of mitochondrial NADH and FADH 2, which are the major substrates for the ETC. The subsequent stimulation of ETC flux results from an increased rate of NADH generation and a primary stimulation of ETC complex activi- ties by other metabolic signals present during periods of increased ATP demand. Increased rates of ETC flux result in increased rates of H ÷ pumping by the ETC, which serve to maintain the electrochemical gradient across the inner mito- chondrial membrane at a level adequate to support the increased rate of ATP synthesis. In addition, the FtFo-ATPase is regu- lated such that the fraction of this enzyme that is in the active state is increased by elevation of mitochondrial matrix Ca :+.

An increased fraction of the enzyme in the active state is ben- eficial during periods of increased ATP synthesis because an increased amount of active enzyme increases the rate of ATP synthesis without requiring ADP levels to rise as long as the rate of ADP delivery to mitochondria increases appropriately (which it will when the rate of ATP utilization increases).

These observations support the concept that increased aver- age cardiomyocyte cytosolic Ca 2+ levels (e.g., that occur dur- ing exercise) act as a f e e d - f o r w a r d signal for oxidative phosphorylation. This signal stimulates, in a parallel manner, ATP utilization by the contractile processes and ATP synthesis by the FiF0-ATPase. The net effect of these regulatory mecha- nisms (and additional feedback regulatory mechanisms not dis- cussed) is to facilitate flux of basic food-derived carbon substrates through the metabolic sequences without requiring the cytosolic concentrations of the initial and intermediate substrates (including ADP) to increase.

The rapid response time of the entire system also allows ATP levels to remain relatively constant despite wide and rapid fluctuations in the rate of ATP utilization. Because these mechanisms allow both ADP and ATP levels and the ATP/

ADP ratio to remain stable at increased ATP synthetic rates, the free energy that can be released by ATP hydrolysis and transferred to other reactions is maintained constant despite the increased rate of ATP expenditure. Control theory as applied to the regulation of oxidative phosphorylation, glyco- lysis, and [3-lipid oxidation, and the TCA cycle has been the subject of detailed study. A more detailed discussion of this topic is presented in the sources listed at the end of this chapter.

4. METABOLISM IN ABNORMAL M Y O C A R D I U M 4.1. Metabolism in Ischemic Myocardium

The most common cause of inadequate myocardial ATP synthesis is ischemia, and the most frequent cause of myocar- dial ischemia is occlusive coronary artery disease. If narrowed coronary arteries can conduct adequate blood flow to sustain aerobic metabolism in the dependent myocardium during basal levels of activity but do not allow sufficient increases in flow to meet increased demands for ATP production during states of increased cardiac work (as would be caused by exercise), then it is classified as demand ischemia. This state can often be observed in patients with stable, inducible angina pectoris; the patient is asymptomatic at rest or during mild exertion, but when the work of the heart exceeds the ability of a narrowed coronary artery to meet the increased blood flow demand, then ischemia ensues, and the patient experiences chest pain. In response to the symptom of chest pain, the patient will com- monly discontinue exercise. During the ensuing rest period, the cardiac work falls into the range at which the narrowed coronary artery can meet the blood flow demand, and the ischemia is relieved.

In contrast, in supply ischemia the coronary artery nar- rowings are so severe that they limit blood flow in the absence of increased ATP demand; that is, ischemia occurs when the patient is at rest. Supply ischemia is generally caused by an acute narrowing or abrupt complete closure of coronary ves- sels; the latter often causes myocardial infarction. Less com-

(13)

Glucose

! / c e t y i Co-A ----p

Glycolysis PDH~_..._~ NAD H b

: 2 A - - ~ T p + 2 P y r u v a ~ t e + 2 N A D H

~

N A D H

L D H NAD+

Lactate + H ÷

/

Sarcolemma --~] t [ Lactate + H +

TCA Cyde ETC

• 2ATP per glucose molecule from anaerobic glycolysis as compared to 38 ATP/glucose molecule from aerobic glycolysis

Fig. 13. A flowchart for anaerobic glucose metabolism is depicted. A large red X indicates metabolic pathways blocked during ischemia. See text for discussion. ATP, adenosine-triphosphate; CoA, coenzyme A; ETC, electron transport chain; LDH, lactic acid dehydrogenase; NAD, oxidized nicotinamide adenine dinucleotide ; NADH, reduced nicotinamide adenine dinucleotide; PDH, pyruvate dehydrogenase; TCA, tricarboxylic acid.

monly, supply ischemia can occur as the result of coronary spasm, in which inappropriate (pathological) vasoconstriction of an epicardial coronary artery causes transient total or near total reduction of flow.

Although the nature of the myocardial ischemic process is qualitatively similar in supply and demand ischemias, supply ischemia is generally considered more clinically threatening.

Because of the greater vulnerability of the inner myocardial layers of the left ventricle to blood flow reduction, mild ischemia is usually confined to the subendocardium; as the coronary obstruction becomes more severe, ischemia proceeds as a wavefront from subendocardium to subepicardium until the full wall is involved.

When myocardial blood flow is limited, both oxygen and carbon substrate availabilities in turn can limit the ATP syn- thetic rate, but it is usually the oxygen deficit that is most criti- cal. This is because the myocyte has substantial carbon substrate reserves (i.e., glycogen and triglyceride stores), but is unable to store a significant quantity of oxygen. Consequently, when coronary blood flow is insufficient to meet myocardial needs, the severity of the oxygen deficit determines the extent of limi- tation of oxidative phosphorylation. Nevertheless, the effect of oxygen limitation is confined to oxidative phosphorylation because glycolytic ATP synthesis does not require oxygen (Figs. 5 and 13).

During hypoxia (in which blood oxygen content is markedly reduced, but blood flow is not limited) or in ischemic conditions (in which blood flow is restricted, but blood oxygen content is normal), electron transport by the ETC slows or stops depend- ing on the severity of the oxygen deficit because electrons can- not be transferred from cytochrome oxidase to molecular

oxygen (the terminal electron acceptor). As a consequence, the pumping of H + from the mitochondrial matrix into the inter- membrane space stops because it depends on energy liberated as electrons pass through the ETC. As a result, the protonmotive force across the inner mitochondrial membrane begins to dissi- pate, and mitochondrial ATP synthesis fails.

If the dissipation of the H + gradient is marked, then the F~F0- H+-ATPase operates in the reverse direction; that is, it converts ATP to ADP and uses energy liberated by ATP hydrolysis to pump H + back into the intermembrane space. Hence, mitochon- dria can either synthesize ATP under physiological conditions or degrade ATP when oxygen is severely limited. Because the H + electrochemical potential is used to energize other mito- chondrial transport functions in addition to supporting ATP synthesis, mitochondrial degradation of ATP during periods of oxygen limitation may serve to preserve nonoxidative mito- chondrial function for a time at the expense of hastening the fall in cytoplasmic ATP levels.

Glycolysis, which in principle can proceed normally in the absence of oxygen, is also limited when blood flow is reduced during ischemia. Oxygen limitation causes cessation of ETC function and thereby stops the mitochondrial consumption of reducing equivalents transferred from cytosolic NADH. This occurs because mitochondrial NADH oxidation also stops, mitochondrial NAD ÷ levels fall, and there is no recipient for reducing equivalents transported from the cytosolic NADH pool. Cytosolic NAD + is a substrate of glyceraldehyde phos- phate dehydrogenase (GAPDH), and if the NADH generated by that reaction is not oxidized back to NAD + by the transfer of reducing equivalents into the mitochondria by the malate- aspartate shuttle or the conversion of glycolytically produced

(14)

236 PART II1: PHYSIOLOGY A N D ASSESSMENT / FROM A N D BACHE

pyruvate into lactate by LDH and the subsequent efflux of lactate and accompanying H+ions from the myocyte, then the lack of NAD ÷ inhibits GAPDH.

Therefore, factors that limit the rates of glycolytic ATP syn- thesis during ischemia are accumulations of cytosolic NADH, lactate and H ÷ ions. Under conditions of unrestricted coronary flow, the myocardium can rapidly export both of these ions even when the glycolytic rate is markedly increased; that is, a low oxygen content of coronary arterial blood which limits oxygen delivery to the myocyte does n o t p e r se limit lactate and H ÷ export. In contrast, when the myocardium is ischemic, the capacity to export lactate and protons decreases in proportion to the severity of the blood flow reduction. Because the conver- sion of pyruvate to lactate results in the oxidation of cytosolic NADH, the availability of NAD ÷ to GAPDH is further reduced when rising levels of lactate slow this reaction and glycolysis ultimately stops for this reason or because glycogen stores are exhausted (Figs. 5 and 9). In contrast, during hypoxia, coronary flow is not restricted, and glycolysis proceeds at a rapid rate because metabolic signaling activates both the rate-limiting glycolytic reactions and the transport of glucose into the myo- cyte. During moderate ischemia, when both glucose and oxy- gen delivery are impaired, myocyte glycogen stores are the main source of glucose for glycolysis. When glycogen stores are exhausted, the rates of glucose delivery to, and lactate clear- ance from, the ischemic myocyte (which are limited) control the rate of glycolysis.

These concepts have specific importance in patients with coronary artery disease because they define the boundaries for viability of the ischemic myocardium. In the moderately ischemic myocardium, contractile function rapidly decreases in response to: (1) decreased ATP availability, (2) accumula- tion of H ÷ ions, and (3) other less-well-defined factors. How- ever, if residual coronary blood flow during ischemia is sufficient to permit some glycolysis (by allowing the export of some lactate and H ÷ and the import of glucose), then modest ATP production and, consequently, viability may be preserved for some time. In contrast, during acute total coronary occlu- sion, glycolysis is rapidly inhibited, and myocardial death will occur. Hence, a level of coronary blood flow that is insufficient to maintain oxidative phosphorylation and contractile function may support sufficient glycolytic ATP production to sustain myocardial viability until appropriate medical or surgical treat- ment can restore blood flow to the ischemic myocardium.

4.2. Metabolism in Hypertrophied and Failing Hearts

During fetal life, the myocardium primarily metabolizes glucose rather than fatty acids. This pattern of metabolism is well suited to the lower oxygen levels found in fetal arterial blood; glucose metabolism requires less oxygen than fatty acid metabolism. In addition, the lower cardiac work levels (lower blood pressure) occurring in the fetus are adequately supported by ATP production through glycolysis and pyruvate oxidation.

After birth, substrate preference of the heart (now in a higher oxygen environment and operating at a higher workload) rap- idly shifts to the adult pattern of predominant fatty acid oxida-

tion. This change of substrate preference is the result of upreg- ulated expression of genes controlling the transport of fatty acids into the myocyte and their subsequent metabolism in the cytosol and mitochondria. This upregulation in the expression of genes involved in fatty acid metabolism, and oxidative phos- phorylation is commonly referred to as the shift from a fetal to an adult gene expression pattern.

A chronic mechanical overload produced by hypertension, heart valve abnormalities, or destruction of a portion of the left ventricle by infarct initially results in some degree of left ven- tricular dilation. Left ventricular wall tension increases as a con- sequence of the La Place relationship between wall tension and chamber radius, The increased wall tension activates changes in the expression patterns of a number of genes involved in both myocyte growth (hypertrophy) and metabolism. The resultant increase of left ventricular wall thickness returns systolic wall stresses more toward normal, often resulting in a prolonged period of compensated myocardial hypertrophy.

However, myocardial hypertrophy is associated with impor- tant changes in energy metabolism. These include reductions of high-energy phosphate compounds (HEP; i.e., ATP and CP) levels and alterations in carbon substrate preference with decreased utilization of free fatty acids and increased depen- dence on glucose. This shift in the myocardial gene expression to increased glucose and decreased fatty acid utilization is referred to as a reversion to the fetal gene expression pattern.

Subsequently, reductions in the levels of enzymes and trans- porters crucial to oxidative phosphorylation also occur.

Although the increased glucose metabolism and decreased fatty acid metabolism associated with this reversion (to a fetal gene expression pattern) may have energetic benefits, the overall ATP synthetic capacity of severely hypertrophied and failing myocardium is significantly reduced. Whatever the cause, the period of compensated hypertrophy in an over- loaded heart is often followed by contractile dysfunctions of the myocytes with ultimate progressive left ventricular dys- function, then to be diagnosed as the clinical syndrome of heart failure.

It has been unclear whether decreased ATP synthetic capacity in hypertrophied or failing cardiomyocytes actually constrains contractile performance and contributes to the pro- gression to a decompensated state. The issue has been whether contractile dysfunction in the failing heart is caused primarily by a limited ability of the contractile apparatus to consume ATP or whether the reduced ability to produce ATP is also contribu- tory. Experimental data have shown that the upregulation of glucose metabolism that occurs in hypertrophied or failing myocardium may not adequately compensate (from the stand- point of ATP synthetic capacity) for the downregulation of fatty acid metabolism that occurs in these hearts.

Interestingly, transgenic mice that overexpressed a cardio- myocyte sarcolemmal glucose transporter (which caused an increase of glucose uptake and metabolism) were far less likely than nontransgenic mice to develop cardiac decompensation or die when subjected to chronic pressure overload of the left ventricle (induced by surgical narrowing of the ascending aorta). This finding implies that augmentation of glucose uptake is beneficial to the chronically overloaded heart, but that

Riferimenti

Documenti correlati

BEHA, DMEA and TOA were used for the solvent extraction of silver(I) ions from synthetic chloride solutions, and tested from chloride pregnant leach

We performed a mass modelling of the rotation curve of a spiral galaxy, representing it as a function of two parameters that govern its shape and one that governs its amplitude,

Without any doubt, critical events serve as triggering events as the authors demonstrate throughout the chapters, but people – ethnic minority youth – also engage

The aim of the current study was to investigate the effect of using a new CTC (vs. an STC) among well-trained professional dancers on pain perception and toe deviation angle (TDA)..

L’individuazione delle varianti genetiche associate con il rischio di sviluppare tali tumori permetterebbe l’identificazione, nella popolazione generale, degli individui ad

WISC = Wechsler Intelligence Scale for Children; SLD = specific learning disorder; VCI = Verbal Comprehension Index; PRI = Perceptual Reasoning Index; WMI = Working Memory Index; PSI

In questa fase è molto importante il ruolo delle misure di protezione attiva, poiché è maggiore la probabilità di spegnimento dell’incendio: la rilevazione automatica di

no statutory paternity leave. The maximum period of leave in Croatia is 12 or 14 months for the first and the second child, depending on whether and how a mother and father share