Download as pdf or txt
Download as pdf or txt
You are on page 1of 40

PA R T

II
BIOENERGETICS AND METABOLISM
13 Principles of Bioenergetics 480 Although metabolism embraces hundreds of differ-
14 Glycolysis, Gluconeogenesis, and the Pentose ent enzyme-catalyzed reactions, our major concern in
Phosphate Pathway 521 Part II is the central metabolic pathways, which are few
in number and remarkably similar in all forms of life.
15 Principles of Metabolic Regulation, Illustrated with Living organisms can be divided into two large groups
the Metabolism of Glucose and Glycogen 560 according to the chemical form in which they obtain
16 The Citric Acid Cycle 601 carbon from the environment. Autotrophs (such as
17 Fatty Acid Catabolism 631 photosynthetic bacteria and vascular plants) can use
carbon dioxide from the atmosphere as their sole source
18 Amino Acid Oxidation and the Production of carbon, from which they construct all their carbon-
of Urea 666 containing biomolecules (see Fig. 1–5). Some auto-
19 Oxidative Phosphorylation and trophic organisms, such as cyanobacteria, can also use
Photophosphorylation 700 atmospheric nitrogen to generate all their nitrogenous
20 Carbohydrate Biosynthesis in Plants components. Heterotrophs cannot use atmospheric
and Bacteria 761 carbon dioxide and must obtain carbon from their en-
vironment in the form of relatively complex organic mol-
21 Lipid Biosynthesis 797 ecules such as glucose. Multicellular animals and most
22 Biosynthesis of Amino Acids, Nucleotides, and microorganisms are heterotrophic. Autotrophic cells
Related Molecules 843 and organisms are relatively self-sufficient, whereas het-
23 Integration and Hormonal Regulation of Mammalian erotrophic cells and organisms, with their requirements
Metabolism 891 for carbon in more complex forms, must subsist on the
products of other organisms.
Many autotrophic organisms are photosynthetic
and obtain their energy from sunlight, whereas het-
erotrophic organisms obtain their energy from the
etabolism is a highly coordinated cellular activity
M in which many multienzyme systems (metabolic
pathways) cooperate to (1) obtain chemical energy by
degradation of organic nutrients produced by auto-
trophs. In our biosphere, autotrophs and heterotrophs
live together in a vast, interdependent cycle in which
capturing solar energy or degrading energy-rich nutrients autotrophic organisms use atmospheric carbon dioxide
from the environment; (2) convert nutrient molecules to build their organic biomolecules, some of them gen-
into the cell’s own characteristic molecules, including erating oxygen from water in the process. Heterotrophs
precursors of macromolecules; (3) polymerize mono- in turn use the organic products of autotrophs as nu-
meric precursors into macromolecules: proteins, nucleic trients and return carbon dioxide to the atmosphere.
acids, and polysaccharides; and (4) synthesize and Some of the oxidation reactions that produce carbon
degrade biomolecules required for specialized cellular dioxide also consume oxygen, converting it to water.
functions, such as membrane lipids, intracellular mes- Thus carbon, oxygen, and water are constantly cycled
sengers, and pigments. between the heterotrophic and autotrophic worlds, with

481
482 Part II Bioenergetics and Metabolism

solar energy as the driving force for this global process Atmospheric
N2
(Fig. 1).
All living organisms also require a source of nitro-
gen, which is necessary for the synthesis of amino acids,
nucleotides, and other compounds. Plants can generally
use either ammonia or nitrate as their sole source of ni- Nitrogen-
Denitrifying
fixing
trogen, but vertebrates must obtain nitrogen in the form bacteria
bacteria
of amino acids or other organic compounds. Only a few
organisms—the cyanobacteria and many species of soil
bacteria that live symbiotically on the roots of some
plants—are capable of converting (“fixing”) atmos- Ammonia
pheric nitrogen (N2) into ammonia. Other bacteria (the
nitrifying bacteria) oxidize ammonia to nitrites and ni-
trates; yet others convert nitrate to N2. Thus, in addi- Nitrifying
Animals
tion to the global carbon and oxygen cycle, a nitrogen bacteria
cycle operates in the biosphere, turning over huge
amounts of nitrogen (Fig. 2). The cycling of carbon, oxy-
gen, and nitrogen, which ultimately involves all species,
depends on a proper balance between the activities of Amino Nitrates,
the producers (autotrophs) and consumers (het- acids nitrites
erotrophs) in our biosphere.
These cycles of matter are driven by an enormous
flow of energy into and through the biosphere, begin- Plants
ning with the capture of solar energy by photosynthetic
organisms and use of this energy to generate energy-
FIGURE 2 Cycling of nitrogen in the biosphere. Gaseous nitrogen
rich carbohydrates and other organic nutrients; these (N2) makes up 80% of the earth’s atmosphere.
nutrients are then used as energy sources by het-
erotrophic organisms. In metabolic processes, and in all
energy transformations, there is a loss of useful energy
(free energy) and an inevitable increase in the amount through the biosphere; organisms cannot regenerate
of unusable energy (heat and entropy). In contrast useful energy from energy dissipated as heat and
to the cycling of matter, therefore, energy flows one way entropy. Carbon, oxygen, and nitrogen recycle continu-
ously, but energy is constantly transformed into unus-
able forms such as heat.
Metabolism, the sum of all the chemical transfor-
mations taking place in a cell or organism, occurs
through a series of enzyme-catalyzed reactions that con-
stitute metabolic pathways. Each of the consecutive
steps in a metabolic pathway brings about a specific,
small chemical change, usually the removal, transfer, or
addition of a particular atom or functional group. The
precursor is converted into a product through a series
O2 of metabolic intermediates called metabolites. The
term intermediary metabolism is often applied to the
nic produc
ga combined activities of all the metabolic pathways that
Or

ts

interconvert precursors, metabolites, and products of


Photosynthetic Heterotrophs low molecular weight (generally, Mr 1,000).
autotrophs
Catabolism is the degradative phase of metabolism
in which organic nutrient molecules (carbohydrates,
C O2 fats, and proteins) are converted into smaller, simpler
H 2O end products (such as lactic acid, CO2, NH3). Catabolic
pathways release energy, some of which is conserved in
FIGURE 1 Cycling of carbon dioxide and oxygen between the auto- the formation of ATP and reduced electron carriers
trophic (photosynthetic) and heterotrophic domains in the biosphere. (NADH, NADPH, and FADH2); the rest is lost as heat.
The flow of mass through this cycle is enormous; about 4  1011 met- In anabolism, also called biosynthesis, small, simple
ric tons of carbon are turned over in the biosphere annually. precursors are built up into larger and more complex
Part II Bioenergetics and Metabolism 483

molecules, including lipids, polysaccharides, proteins, simultaneous synthesis and degradation of fatty acids
and nucleic acids. Anabolic reactions require an input would be wasteful, however, and this is prevented by
of energy, generally in the form of the phosphoryl group reciprocally regulating the anabolic and catabolic reac-
transfer potential of ATP and the reducing power of tion sequences: when one sequence is active, the other
NADH, NADPH, and FADH2 (Fig. 3). is suppressed. Such regulation could not occur if ana-
Some metabolic pathways are linear, and some are bolic and catabolic pathways were catalyzed by exactly
branched, yielding multiple useful end products from a the same set of enzymes, operating in one direction for
single precursor or converting several starting materi- anabolism, the opposite direction for catabolism: inhi-
als into a single product. In general, catabolic pathways bition of an enzyme involved in catabolism would also
are convergent and anabolic pathways divergent (Fig. inhibit the reaction sequence in the anabolic direction.
4). Some pathways are cyclic: one starting component Catabolic and anabolic pathways that connect the same
of the pathway is regenerated in a series of reactions two end points (glucose n n pyruvate and pyruvate
that converts another starting component into a prod- n n glucose, for example) may employ many of the
uct. We shall see examples of each type of pathway in same enzymes, but invariably at least one of the steps
the following chapters. is catalyzed by different enzymes in the catabolic and
Most cells have the enzymes to carry out both the anabolic directions, and these enzymes are the sites of
degradation and the synthesis of the important cate- separate regulation. Moreover, for both anabolic and
gories of biomolecules—fatty acids, for example. The catabolic pathways to be essentially irreversible, the re-
actions unique to each direction must include at least
one that is thermodynamically very favorable—in other
words, a reaction for which the reverse reaction is very
Cell Energy- unfavorable. As a further contribution to the separate
macromolecules containing
nutrients
regulation of catabolic and anabolic reaction sequences,
Proteins paired catabolic and anabolic pathways commonly take
Polysaccharides Carbohydrates
Lipids Fats place in different cellular compartments: for example,
Nucleic acids Proteins fatty acid catabolism in mitochondria, fatty acid syn-
thesis in the cytosol. The concentrations of intermedi-
ates, enzymes, and regulators can be maintained at
different levels in these different compartments. Be-
cause metabolic pathways are subject to kinetic con-
ADP  HPO2
4
trol by substrate concentration, separate pools of
NAD anabolic and catabolic intermediates also contribute to
NADP
FAD the control of metabolic rates. Devices that separate
anabolic and catabolic processes will be of particular
interest in our discussions of metabolism.
Anabolism ATP Catabolism
Metabolic pathways are regulated at several levels,
NADH
NADPH from within the cell and from outside. The most imme-
FADH2 diate regulation is by the availability of substrate; when
the intracellular concentration of an enzyme’s substrate
is near or below Km (as is commonly the case), the rate
Chemical of the reaction depends strongly upon substrate con-
energy centration (see Fig. 6–11). A second type of rapid con-
trol from within is allosteric regulation (p. 225) by a
metabolic intermediate or coenzyme—an amino acid or
Precursor Energy- ATP, for example—that signals the cell’s internal meta-
molecules depleted bolic state. When the cell contains an amount of, say,
end products
Amino acids aspartate sufficient for its immediate needs, or when the
Sugars CO2 cellular level of ATP indicates that further fuel con-
Fatty acids H2O
Nitrogenous bases NH3 sumption is unnecessary at the moment, these signals
allosterically inhibit the activity of one or more enzymes
in the relevant pathway. In multicellular organisms the
FIGURE 3 Energy relationships between catabolic and anabolic metabolic activities of different tissues are regulated and
pathways. Catabolic pathways deliver chemical energy in the form integrated by growth factors and hormones that act from
of ATP, NADH, NADPH, and FADH2. These energy carriers are used outside the cell. In some cases this regulation occurs
in anabolic pathways to convert small precursor molecules into cell virtually instantaneously (sometimes in less than a mil-
macromolecules. lisecond) through changes in the levels of intracellular
484 Part II Bioenergetics and Metabolism

Rubber Carotenoid Steroid


pigments hormones

Phospholipids Isopentenyl- Bile


pyrophosphate Cholesterol
acids
Triacylglycerols Fatty acids

Mevalonate Vitamin K Cholesteryl


Phenyl- esters
Starch Alanine
alanine
Acetate
Glycogen Glucose Pyruvate (acetyl-CoA) Acetoacetyl-CoA Eicosanoids

Sucrose Serine Leucine


Isoleucine Fatty acids Triacylglycerols

(a) Converging catabolism


Citrate CDP-diacylglycerol Phospholipids
Oxaloacetate (b) Diverging anabolism

CO2

CO2
(c) Cyclic pathway

FIGURE 4 Three types of nonlinear metabolic pathways. (a) Con- the breakdown product of a variety of fuels (a), serves as the precur-
verging, catabolic; (b) diverging, anabolic; and (c) cyclic, in which sor for an array of products (b), and is consumed in the catabolic path-
one of the starting materials (oxaloacetate in this case) is regenerated way known as the citric acid cycle (c).
and reenters the pathway. Acetate, a key metabolic intermediate, is

messengers that modify the activity of existing enzyme Before reviewing the five main reaction classes of
molecules by allosteric mechanisms or by covalent mod- biochemistry, let’s consider two basic chemical princi-
ification such as phosphorylation. In other cases, the ex- ples. First, a covalent bond consists of a shared pair of
tracellular signal changes the cellular concentration of electrons, and the bond can be broken in two general
an enzyme by altering the rate of its synthesis or degra- ways (Fig. 5). In homolytic cleavage, each atom leaves
dation, so the effect is seen only after minutes or hours. the bond as a radical, carrying one of the two electrons
The number of metabolic transformations taking (now unpaired) that held the bonded atoms together.
place in a typical cell can seem overwhelming to a be- In the more common, heterolytic cleavage, one atom re-
ginning student. Most cells have the capacity to carry tains both bonding electrons. The species generated
out thousands of specific, enzyme-catalyzed reactions: when COC and COH bonds are cleaved are illustrated
for example, transformation of a simple nutrient such in Figure 5. Carbanions, carbocations, and hydride ions
as glucose into amino acids, nucleotides, or lipids; ex- are highly unstable; this instability shapes the chemistry
traction of energy from fuels by oxidation; or polymer- of these ions, as described further below.
ization of monomeric subunits into macromolecules. The second chemical principle of interest here is that
Fortunately for the student of biochemistry, there are many biochemical reactions involve interactions between
patterns within this multitude of reactions; you do not nucleophiles (functional groups rich in electrons and
need to learn all these reactions to comprehend the capable of donating them) and electrophiles (electron-
molecular logic of biochemistry. Most of the reactions deficient functional groups that seek electrons). Nucle-
in living cells fall into one of five general categories: ophiles combine with, and give up electrons to, elec-
(1) oxidation-reductions; (2) reactions that make or trophiles. Common nucleophiles and electrophiles are
break carbon–carbon bonds; (3) internal rearrangements, listed in Figure 6–21. Note that a carbon atom can act
isomerizations, and eliminations; (4) group transfers; as either a nucleophile or an electrophile, depending on
and (5) free radical reactions. Reactions within each which bonds and functional groups surround it.
general category usually proceed by a limited set of We now consider the five main reaction classes you
mechanisms and often employ characteristic cofactors. will encounter in upcoming chapters.
Part II Bioenergetics and Metabolism 485

Homolytic catalyze these oxidations are generally called oxidases


C H C  H
cleavage or, if the oxygen atom is derived directly from molecu-
Carbon
lar oxygen (O2), oxygenases.
H atom
radical Every oxidation must be accompanied by a reduc-
tion, in which an electron acceptor acquires the electrons
C C C  C
removed by oxidation. Oxidation reactions generally
release energy (think of camp fires: the compounds in
Carbon radicals wood are oxidized by oxygen molecules in the air). Most
living cells obtain the energy needed for cellular work by
oxidizing metabolic fuels such as carbohydrates or fat;
Heterolytic C H C   H photosynthetic organisms can also trap and use the en-
cleavage
ergy of sunlight. The catabolic (energy-yielding) path-
Carbanion Proton ways described in Chapters 14 through 19 are oxidative
reaction sequences that result in the transfer of electrons
from fuel molecules, through a series of electron carri-
C H C  H 
ers, to oxygen. The high affinity of O2 for electrons makes
the overall electron-transfer process highly exergonic,
Carbocation Hydride providing the energy that drives ATP synthesis—the
central goal of catabolism.
C C C   C
2. Reactions that make or break carbon–carbon bonds Het-
erolytic cleavage of a COC bond yields a carbanion and
Carbanion Carbocation
a carbocation (Fig. 5). Conversely, the formation of a
FIGURE 5 Two mechanisms for cleavage of a COC or COH bond. COC bond involves the combination of a nucleophilic
In homolytic cleavages, each atom keeps one of the bonding elec- carbanion and an electrophilic carbocation. Groups with
trons, resulting in the formation of carbon radicals (carbons having electronegative atoms play key roles in these reactions.
unpaired electrons) or uncharged hydrogen atoms. In heterolytic cleav- Carbonyl groups are particularly important in the chem-
ages, one of the atoms retains both bonding electrons. This can result ical transformations of metabolic pathways. As noted
in the formation of carbanions, carbocations, protons, or hydride ions. above, the carbon of a carbonyl group has a partial pos-
itive charge due to the electron-withdrawing nature of
1. Oxidation-reduction reactions Carbon atoms encoun- the adjacent bonded oxygen, and thus is an electrophilic
tered in biochemistry can exist in five oxidation states, carbon. The presence of a carbonyl group can also
depending on the elements with which carbon shares facilitate the formation of a carbanion on an adjoining
electrons (Fig. 6). In many biological oxidations, a com- carbon, because the carbonyl group can delocalize elec-
pound loses two electrons and two hydrogen ions (that trons through resonance (Fig. 8a, b). The importance
is, two hydrogen atoms); these reactions are commonly of a carbonyl group is evident in three major classes of
called dehydrogenations and the enzymes that catalyze reactions in which COC bonds are formed or broken
them are called dehydrogenases (Fig. 7). In some, but (Fig 8c): aldol condensations (such as the aldolase
not all, biological oxidations, a carbon atom becomes co- reaction; see Fig. 14–5), Claisen condensations (as
valently bonded to an oxygen atom. The enzymes that in the citrate synthase reaction; see Fig. 16–9), and

OH 2H  2e O
CH2 CH3 Alkane O O
CH3 CH C CH3 C C
CH2 CH2OH Alcohol
O O
O 2H  2e
CH2 C Aldehyde (ketone) Lactate lactate Pyruvate
dehydrogenase
H(R)
O FIGURE 7 An oxidation-reduction reaction. Shown here is the oxi-
CH2 C Carboxylic acid dation of lactate to pyruvate. In this dehydrogenation, two electrons
OH and two hydrogen ions (the equivalent of two hydrogen atoms) are re-
O C O Carbon dioxide moved from C-2 of lactate, an alcohol, to form pyruvate, a ketone. In
cells the reaction is catalyzed by lactate dehydrogenase and the elec-
FIGURE 6 The oxidation states of carbon in biomolecules. Each com- trons are transferred to a cofactor called nicotinamide adenine dinu-
pound is formed by oxidation of the red carbon in the compound cleotide. This reaction is fully reversible; pyruvate can be reduced by
listed above it. Carbon dioxide is the most highly oxidized form of electrons from the cofactor. In Chapter 13 we discuss the factors that
carbon found in living systems. determine the direction of a reaction.
486 Part II Bioenergetics and Metabolism

decarboxylations (as in the acetoacetate decarboxylase electrons results in isomerization, transposition of dou-
reaction; see Fig. 17–18). Entire metabolic pathways are ble bonds, or cis-trans rearrangements of double bonds.
organized around the introduction of a carbonyl group An example of isomerization is the formation of fruc-
in a particular location so that a nearby carbon–carbon tose 6-phosphate from glucose 6-phosphate during
bond can be formed or cleaved. In some reactions, this sugar metabolism (Fig 9a; this reaction is discussed in
role is played by an imine group or a specialized cofac- detail in Chapter 14). Carbon-1 is reduced (from alde-
tor such as pyridoxal phosphate, rather than by a car- hyde to alcohol) and C-2 is oxidized (from alcohol to
bonyl group. ketone). Figure 9b shows the details of the electron
movements that result in isomerization.
3. Internal rearrangements, isomerizations, and eliminations A simple transposition of a CUC bond occurs dur-
Another common type of cellular reaction is an in- ing metabolism of the common fatty acid oleic acid (see
tramolecular rearrangement, in which redistribution of Fig. 17–9), and you will encounter some spectacular ex-
amples of double-bond repositioning in the synthesis of
cholesterol (see Fig. 21–35).
Elimination of water introduces a CUC bond be-
O
tween two carbons that previously were saturated (as
(a) C in the enolase reaction; see Fig. 6–23). Similar reactions
can result in the elimination of alcohols and amines.
O O
H H
(b) C C C C H2O R H
R C C R1 C C
H2O H R1
H OH
O R2 R3 O R2 R3
H

(c) R1 C C C O R1 C C C OH
4. Group transfer reactions The transfer of acyl, glycosyl,
H R4 H R4 and phosphoryl groups from one nucleophile to another
Aldol condensation is common in living cells. Acyl group transfer generally
involves the addition of a nucleophile to the carbonyl
O H R1 O H R1
H carbon of an acyl group to form a tetrahedral interme-

CoA-S C C C O CoA-S C C C OH diate.
H R2 H R2
O O O
Claisen ester condensation
C R C X C
O H O H R X Y R Y
O H
Y Tetrahedral X
R C C C R C C H  CO2
intermediate
H O H
Decarboxylation of a -keto acid The chymotrypsin reaction is one example of acyl group
transfer (see Fig. 6–21). Glycosyl group transfers in-
FIGURE 8 Carbon–carbon bond formation reactions. (a) The carbon volve nucleophilic substitution at C-1 of a sugar ring,
atom of a carbonyl group is an electrophile by virtue of the electron- which is the central atom of an acetal. In principle, the
withdrawing capacity of the electronegative oxygen atom, which results substitution could proceed by an SN1 or SN2 path, as
in a resonance hybrid structure in which the carbon has a partial pos-
described for the enzyme lysozyme (see Fig. 6–25).
itive charge. (b) Within a molecule, delocalization of electrons into a
Phosphoryl group transfers play a special role in
carbonyl group facilitates the transient formation of a carbanion on an
metabolic pathways. A general theme in metabolism is
adjacent carbon. (c) Some of the major reactions involved in the for-
the attachment of a good leaving group to a metabolic
mation and breakage of COC bonds in biological systems. For both the
intermediate to “activate” the intermediate for subse-
aldol condensation and the Claisen condensation, a carbanion serves
as nucleophile and the carbon of a carbonyl group serves as elec-
quent reaction. Among the better leaving groups in
trophile. The carbanion is stabilized in each case by another carbonyl
nucleophilic substitution reactions are inorganic or-
at the carbon adjoining the carbanion carbon. In the decarboxylation thophosphate (the ionized form of H3PO4 at neutral pH,
reaction, a carbanion is formed on the carbon shaded blue as the CO2 a mixture of H2PO 2
4 and HPO4 , commonly abbreviated
leaves. The reaction would not occur at an appreciable rate but for Pi) and inorganic pyrophosphate (P2O74, abbreviated
the stabilizing effect of the carbonyl adjacent to the carbanion car- PPi); esters and anhydrides of phosphoric acid are
bon. Wherever a carbanion is shown, a stabilizing resonance with the effectively activated for reaction. Nucleophilic substi-
adjacent carbonyl, as shown in (a), is assumed. The formation of the tution is made more favorable by the attachment of a
carbanion is highly disfavored unless the stabilizing carbonyl group, phosphoryl group to an otherwise poor leaving group
or a group of similar function such as an imine, is present. such as OOH. Nucleophilic substitutions in which the
Part II Bioenergetics and Metabolism 487

(a) H OH H H H O H OH H H H O
H C 1 2
C C C C C O P O H C 1 2
C C C C C O P O
phosphohexose
O OH H OH OH H O isomerase OH O H OH OH H O
Glucose 6-phosphate Fructose 6-phosphate

(b)
B1 1 B1 abstracts a B1 B1
proton.
H 5 An electron leaves
the C C bond to form H
H 2 This allows the a C H bond with
formation of a C C C C the proton donated C C
C C double bond. by B1.
O O OH O
O OH H H
3 Electrons from 4 B2 abstracts a
H carbonyl form an proton, allowing
B2 H
O H bond with the formation of
B2 the hydrogen ion aC O bond.
B2
donated by B2.
Enediol intermediate

FIGURE 9 Isomerization and elimination reactions. (a) The conver- rows represent the movement of bonding electrons from nucleophile
sion of glucose 6-phosphate to fructose 6-phosphate, a reaction of (pink) to electrophile (blue). B1 and B2 are basic groups on the
sugar metabolism catalyzed by phosphohexose isomerase. (b) This re- enzyme; they are capable of donating and accepting hydrogen ions
action proceeds through an enediol intermediate. The curved blue ar- (protons) as the reaction progresses.

phosphoryl group (OPO32) serves as a leaving group charge and can therefore act as an electrophile. In a very
occur in hundreds of metabolic reactions. large number of metabolic reactions, a phosphoryl group
Phosphorus can form five covalent bonds. The con- (OPO32) is transferred from ATP to an alcohol (form-
ventional representation of Pi (Fig. 10a), with three ing a phosphate ester) (Fig. 10c) or to a carboxylic acid
POO bonds and one PUO bond, is not an accurate pic- (forming a mixed anhydride). When a nucleophile at-
ture. In Pi, four equivalent phosphorus–oxygen bonds tacks the electrophilic phosphorus atom in ATP, a rela-
share some double-bond character, and the anion has a tively stable pentacovalent structure is formed as a re-
tetrahedral structure (Fig. 10b). As oxygen is more elec- action intermediate (Fig. 10d). With departure of the
tronegative than phosphorus, the sharing of electrons is leaving group (ADP), the transfer of a phosphoryl group
unequal: the central phosphorus bears a partial positive is complete. The large family of enzymes that catalyze

(a)
O O (c) O O O
 O
O P O P O Adenine Ribose O P O P O P O HO R

O O O 
O 
O  Glucose
ATP
O O
O
O O O
P O O P O
Adenine Ribose O P O P O  
O P O R
O O
  
O O O
(b)
ADP Glucose 6-phosphate,
3 a phosphate ester
O O
O P O (d)

O O O
O P
Z P W ZR OH
O O
W  ADP
O

FIGURE 10 Alternative ways of showing the structure of inorganic all four phosphorus–oxygen bonds with some double-bond character;
orthophosphate. (a) In one (inadequate) representation, three oxygens the hybrid orbitals so represented are arranged in a tetrahedron with
are single-bonded to phosphorus, and the fourth is double-bonded, P at its center. (c) When a nucleophile Z (in this case, the OOH on
allowing the four different resonance structures shown. (b) The four C-6 of glucose) attacks ATP, it displaces ADP (W). In this SN2 reac-
resonance structures can be represented more accurately by showing tion, a pentacovalent intermediate (d) forms transiently.
488 Part II Bioenergetics and Metabolism

phosphoryl group transfers with ATP as donor are called chemiosmotic energy coupling, a universal mechanism
kinases (Greek kinein, “to move”). Hexokinase, for ex- in which a transmembrane electrochemical potential,
ample, “moves” a phosphoryl group from ATP to glucose. produced either by substrate oxidation or by light ab-
Phosphoryl groups are not the only activators of this sorption, drives the synthesis of ATP.
type. Thioalcohols (thiols), in which the oxygen atom Chapters 20 through 22 describe the major anabolic
of an alcohol is replaced with a sulfur atom, are also pathways by which cells use the energy in ATP to pro-
good leaving groups. Thiols activate carboxylic acids by duce carbohydrates, lipids, amino acids, and nucleotides
forming thioesters (thiol esters) with them. We will dis- from simpler precursors. In Chapter 23 we step back
cuss a number of cases, including the reactions cat- from our detailed look at the metabolic pathways—as
alyzed by the fatty acyl transferases in lipid synthesis they occur in all organisms, from Escherichia coli to
(see Fig. 21–2), in which nucleophilic substitution at the humans—and consider how they are regulated and in-
carbonyl carbon of a thioester results in transfer of the tegrated in mammals by hormonal mechanisms.
acyl group to another moiety. As we undertake our study of intermediary metab-
olism, a final word. Keep in mind that the myriad re-
5. Free radical reactions Once thought to be rare, the actions described in these pages take place in, and play
homolytic cleavage of covalent bonds to generate free crucial roles in, living organisms. As you encounter each
radicals has now been found in a range of biochemical reaction and each pathway ask, What does this chemi-
processes. Some examples are the reactions of methyl- cal transformation do for the organism? How does this
malonyl-CoA mutase (see Box 17–2), ribonucleotide pathway interconnect with the other pathways operat-
reductase (see Fig. 22–41), and DNA photolyase (see ing simultaneously in the same cell to produce the en-
Fig. 25–25). ergy and products required for cell maintenance and
growth? How do the multilayered regulatory mecha-
We begin Part II with a discussion of the basic en- nisms cooperate to balance metabolic and energy in-
ergetic principles that govern all metabolism (Chapter puts and outputs, achieving the dynamic steady state
13). We then consider the major catabolic pathways by of life? Studied with this perspective, metabolism pro-
which cells obtain energy from the oxidation of various vides fascinating and revealing insights into life, with
fuels (Chapters 14 through 19). Chapter 19 is the piv- countless applications in medicine, agriculture, and
otal point of our discussion of metabolism; it concerns biotechnology.
chapter
13
PRINCIPLES OF BIOENERGETICS
13.1 Bioenergetics and Thermodynamics 490 heat and that this process of
13.2 Phosphoryl Group Transfers and ATP 496 respiration is essential to life.
He observed that
13.3 Biological Oxidation-Reduction Reactions 507
. . . in general, respiration
is nothing but a slow com-
The total energy of the universe is constant; the total bustion of carbon and hy-
entropy is continually increasing. drogen, which is entirely
similar to that which oc-
—Rudolf Clausius, The Mechanical Theory of Heat with Its
curs in a lighted lamp or
Applications to the Steam-Engine and to the Physical
candle, and that, from this
Properties of Bodies, 1865 (trans. 1867)
point of view, animals that Antoine Lavoisier,
respire are true com- 1743–1794
The isomorphism of entropy and information establishes a bustible bodies that burn
link between the two forms of power: the power to do and and consume themselves . . . One may say that this
the power to direct what is done. analogy between combustion and respiration has
—François Jacob, La logique du vivant: une histoire de l’hérédité not escaped the notice of the poets, or rather the
(The Logic of Life: A History of Heredity), 1970 philosophers of antiquity, and which they had ex-
pounded and interpreted. This fire stolen from
heaven, this torch of Prometheus, does not only rep-
iving cells and organisms must perform work to stay resent an ingenious and poetic idea, it is a faithful
L alive, to grow, and to reproduce. The ability to har-
ness energy and to channel it into biological work is a
picture of the operations of nature, at least for an-
imals that breathe; one may therefore say, with the
fundamental property of all living organisms; it must ancients, that the torch of life lights itself at the mo-
have been acquired very early in cellular evolution. Mod- ment the infant breathes for the first time, and it
ern organisms carry out a remarkable variety of energy does not extinguish itself except at death.*
transductions, conversions of one form of energy to an-
In this century, biochemical studies have revealed
other. They use the chemical energy in fuels to bring
much of the chemistry underlying that “torch of life.”
about the synthesis of complex, highly ordered macro-
Biological energy transductions obey the same physical
molecules from simple precursors. They also convert the
laws that govern all other natural processes. It is there-
chemical energy of fuels into concentration gradients
fore essential for a student of biochemistry to under-
and electrical gradients, into motion and heat, and, in a
stand these laws and how they apply to the flow of
few organisms such as fireflies and some deep-sea fish,
energy in the biosphere. In this chapter we first review
into light. Photosynthetic organisms transduce light en-
the laws of thermodynamics and the quantitative rela-
ergy into all these other forms of energy.
tionships among free energy, enthalpy, and entropy. We
The chemical mechanisms that underlie biological
then describe the special role of ATP in biological
energy transductions have fascinated and challenged
biologists for centuries. Antoine Lavoisier, before he lost *From a memoir by Armand Seguin and Antoine Lavoisier, dated 1789,
his head in the French Revolution, recognized that an- quoted in Lavoisier, A. (1862) Oeuvres de Lavoisier, Imprimerie
imals somehow transform chemical fuels (foods) into Impériale, Paris.

489
490 Chapter 13 Principles of Bioenergetics

energy exchanges. Finally, we consider the importance not violate the second law; they operate strictly within
of oxidation-reduction reactions in living cells, the en- it. To discuss the application of the second law to bio-
ergetics of electron-transfer reactions, and the electron logical systems, we must first define those systems and
carriers commonly employed as cofactors of the en- their surroundings.
zymes that catalyze these reactions. The reacting system is the collection of matter that
is undergoing a particular chemical or physical process;
it may be an organism, a cell, or two reacting com-
13.1 Bioenergetics and Thermodynamics pounds. The reacting system and its surroundings to-
gether constitute the universe. In the laboratory, some
Bioenergetics is the quantitative study of the energy chemical or physical processes can be carried out in iso-
transductions that occur in living cells and of the nature lated or closed systems, in which no material or energy
and function of the chemical processes underlying these is exchanged with the surroundings. Living cells and or-
transductions. Although many of the principles of ther- ganisms, however, are open systems, exchanging both
modynamics have been introduced in earlier chapters material and energy with their surroundings; living sys-
and may be familiar to you, a review of the quantitative tems are never at equilibrium with their surroundings,
aspects of these principles is useful here. and the constant transactions between system and sur-
roundings explain how organisms can create order
Biological Energy Transformations Obey the Laws within themselves while operating within the second law
of Thermodynamics of thermodynamics.
Many quantitative observations made by physicists and In Chapter 1 (p. 23) we defined three thermody-
chemists on the interconversion of different forms of namic quantities that describe the energy changes oc-
energy led, in the nineteenth century, to the formula- curring in a chemical reaction:
tion of two fundamental laws of thermodynamics. The
first law is the principle of the conservation of energy: Gibbs free energy, G, expresses the amount of
for any physical or chemical change, the total energy capable of doing work during a reaction
amount of energy in the universe remains constant; at constant temperature and pressure. When a
energy may change form or it may be transported reaction proceeds with the release of free energy
from one region to another, but it cannot be created (that is, when the system changes so as to
or destroyed. The second law of thermodynamics, which possess less free energy), the free-energy change,
can be stated in several forms, says that the universe G, has a negative value and the reaction is said
always tends toward increasing disorder: in all natu- to be exergonic. In endergonic reactions, the
ral processes, the entropy of the universe increases. system gains free energy and G is positive.
Living organisms consist of collections of molecules Enthalpy, H, is the heat content of the reacting
much more highly organized than the surrounding ma- system. It reflects the number and kinds of
terials from which they are constructed, and organisms chemical bonds in the reactants and products.
maintain and produce order, seemingly oblivious to the When a chemical reaction releases heat, it is
second law of thermodynamics. But living organisms do said to be exothermic; the heat content of the
products is less than that of the reactants and
H has, by convention, a negative value. Reacting
systems that take up heat from their surroundings
are endothermic and have positive values of H.
Entropy, S, is a quantitative expression for the
randomness or disorder in a system (see Box 1–3).
When the products of a reaction are less complex
and more disordered than the reactants, the
reaction is said to proceed with a gain in entropy.

The units of G and H are joules/mole or calories/mole


(recall that 1 cal  4.184 J); units of entropy are
joules/mole  Kelvin (J/mol  K) (Table 13–1).
Under the conditions existing in biological systems
(including constant temperature and pressure),
changes in free energy, enthalpy, and entropy are re-
lated to each other quantitatively by the equation
G   H  T S (13–1)
13.1 Bioenergetics and Thermodynamics 491

free energy into ATP and other energy-rich compounds


TABLE 13–1 Some Physical Constants and capable of providing energy for biological work at con-
Units Used in Thermodynamics stant temperature.
Boltzmann constant, k  1.381  1023 J/K
Avogadro’s number, N  6.022  1023 mol1 The Standard Free-Energy Change Is Directly Related
Faraday constant,  96,480 J/V  mol to the Equilibrium Constant
Gas constant, R  8.315 J/mol  K
( 1.987 cal/mol  K) The composition of a reacting system (a mixture of
chemical reactants and products) tends to continue
Units of G and H are J/mol (or cal/mol) changing until equilibrium is reached. At the equilibrium
Units of S are J/mol  K (or cal/mol  K) concentration of reactants and products, the rates of the
1 cal  4.184 J forward and reverse reactions are exactly equal and no
Units of absolute temperature, T, are Kelvin, K further net change occurs in the system. The concen-
25 C  298 K trations of reactants and products at equilibrium
At 25 C, RT  2.479 kJ/mol define the equilibrium constant, Keq (p. 26). In the
( 0.592 kcal/mol) general reaction aA  bB y z cC  dD, where a, b, c, and
d are the number of molecules of A, B, C, and D par-
ticipating, the equilibrium constant is given by

in which G is the change in Gibbs free energy of the [C]c[D]d


Keq  a (13–2)
reacting system, H is the change in enthalpy of the [A] [B]b
system, T is the absolute temperature, and S is the
where [A], [B], [C], and [D] are the molar concentrations
change in entropy of the system. By convention, S has
of the reaction components at the point of equilibrium.
a positive sign when entropy increases and H, as noted
When a reacting system is not at equilibrium, the
above, has a negative sign when heat is released by the
tendency to move toward equilibrium represents a driv-
system to its surroundings. Either of these conditions,
ing force, the magnitude of which can be expressed as
which are typical of favorable processes, tend to make
the free-energy change for the reaction, G. Under stan-
G negative. In fact, G of a spontaneously reacting sys-
dard conditions (298 K  25 C), when reactants and
tem is always negative.
products are initially present at 1 M concentrations or,
The second law of thermodynamics states that the
for gases, at partial pressures of 101.3 kilopascals (kPa),
entropy of the universe increases during all chemical
or 1 atm, the force driving the system toward equilib-
and physical processes, but it does not require that the
rium is defined as the standard free-energy change, G .
entropy increase take place in the reacting system it-
By this definition, the standard state for reactions that
self. The order produced within cells as they grow and
involve hydrogen ions is [H]  1 M, or pH 0. Most bio-
divide is more than compensated for by the disorder
chemical reactions, however, occur in well-buffered
they create in their surroundings in the course of growth
aqueous solutions near pH 7; both the pH and the con-
and division (see Box 1–3, case 2). In short, living or-
centration of water (55.5 M) are essentially constant.
ganisms preserve their internal order by taking from the
For convenience of calculations, biochemists therefore
surroundings free energy in the form of nutrients or sun-
define a different standard state, in which the concen-
light, and returning to their surroundings an equal
tration of H is 107 M (pH 7) and that of water is
amount of energy as heat and entropy.
55.5 M; for reactions that involve Mg2 (including most
in which ATP is a reactant), its concentration in solu-
Cells Require Sources of Free Energy
tion is commonly taken to be constant at 1 mM. Physi-
Cells are isothermal systems—they function at essen- cal constants based on this biochemical standard state
tially constant temperature (they also function at con- are called standard transformed constants and are
stant pressure). Heat flow is not a source of energy for written with a prime (such as G
and K
eq) to distin-
cells, because heat can do work only as it passes to a guish them from the untransformed constants used by
zone or object at a lower temperature. The energy that chemists and physicists. (Notice that most other text-
cells can and must use is free energy, described by the books use the symbol G
rather than G
. Our use of
Gibbs free-energy function G, which allows prediction G
, recommended by an international committee of
of the direction of chemical reactions, their exact equi- chemists and biochemists, is intended to emphasize that
librium position, and the amount of work they can in the transformed free energy G
is the criterion for equi-
theory perform at constant temperature and pressure. librium.) By convention, when H2O, H, and/or Mg2
Heterotrophic cells acquire free energy from nutrient are reactants or products, their concentrations are not
molecules, and photosynthetic cells acquire it from ab- included in equations such as Equation 13–2 but are in-
sorbed solar radiation. Both kinds of cells transform this stead incorporated into the constants K
eq and G
.
492 Chapter 13 Principles of Bioenergetics

Just as K
eq is a physical constant characteristic for
each reaction, so too is G
a constant. As we noted in TABLE 13–3 Relationships among K
eq, G,
Chapter 6, there is a simple relationship between K
eq and the Direction of Chemical Reactions under
and G
: Standard Conditions
G
 RT ln K
eq Starting with all
components at 1 M,
The standard free-energy change of a chemical re-
When K
eq is . . . G
is . . . the reaction . . .
action is simply an alternative mathematical way of
expressing its equilibrium constant. Table 13–2 1.0 negative proceeds forward
shows the relationship between G
and K
eq. If the 1.0 zero is at equilibrium
equilibrium constant for a given chemical reaction is 1.0, 1.0 positive proceeds in reverse
the standard free-energy change of that reaction is 0.0
(the natural logarithm of 1.0 is zero). If K
eq of a reac-
tion is greater than 1.0, its G
is negative. If K
eq is less
than 1.0, G
is positive. Because the relationship be- As an example, let’s make a simple calculation of
tween G
and K
eq is exponential, relatively small the standard free-energy change of the reaction cat-
changes in G
correspond to large changes in K
eq. alyzed by the enzyme phosphoglucomutase:
It may be helpful to think of the standard free- Glucose 1-phosphate 34 glucose 6-phosphate
energy change in another way. G
is the difference be-
tween the free-energy content of the products and the Chemical analysis shows that whether we start with, say,
free-energy content of the reactants, under standard 20 mM glucose 1-phosphate (but no glucose 6-phosphate)
conditions. When G
is negative, the products contain or with 20 mM glucose 6-phosphate (but no glucose
less free energy than the reactants and the reaction will 1-phosphate), the final equilibrium mixture at 25 C and
proceed spontaneously under standard conditions; all pH 7.0 will be the same: 1 mM glucose 1-phosphate and
chemical reactions tend to go in the direction that re- 19 mM glucose 6-phosphate. (Remember that enzymes do
sults in a decrease in the free energy of the system. A not affect the point of equilibrium of a reaction; they
positive value of G
means that the products of the merely hasten its attainment.) From these data we can
reaction contain more free energy than the reactants, calculate the equilibrium constant:
and this reaction will tend to go in the reverse direction [glucose 6-phosphate] 19 mM
if we start with 1.0 M concentrations of all components K
eq      19
[glucose 1-phosphate] 1 mM
(standard conditions). Table 13–3 summarizes these
points. From this value of K
eq we can calculate the standard
free-energy change:

TABLE 13–2 Relationship between the G


 RT ln K
eq
 (8.315 J/mol  K)(298 K)(ln 19)
Equilibrium Constants and Standard Free-Energy
 7.3 kJ/mol
Changes of Chemical Reactions
Because the standard free-energy change is negative,
G
when the reaction starts with 1.0 M glucose 1-phosphate
K
eq (kJ/mol) (kcal/mol)* and 1.0 M glucose 6-phosphate, the conversion of glu-
cose 1-phosphate to glucose 6-phosphate proceeds with
103 17.1 4.1 a loss (release) of free energy. For the reverse reaction
102 11.4 2.7 (the conversion of glucose 6-phosphate to glucose
101 5.7 1.4 1-phosphate), G
has the same magnitude but the op-
1 0.0 0.0 posite sign.
101 5.7 1.4 Table 13–4 gives the standard free-energy changes
102 11.4 2.7 for some representative chemical reactions. Note that
103 17.1 4.1 hydrolysis of simple esters, amides, peptides, and gly-
104 22.8 5.5 cosides, as well as rearrangements and eliminations,
105 28.5 6.8 proceed with relatively small standard free-energy
106 34.2 8.2 changes, whereas hydrolysis of acid anhydrides is ac-
companied by relatively large decreases in standard free
*Although joules and kilojoules are the standard units of energy and are used energy. The complete oxidation of organic compounds
throughout this text, biochemists sometimes express G
values in kilocalories per such as glucose or palmitate to CO2 and H2O, which in
mole. We have therefore included values in both kilojoules and kilocalories in this table
and in Tables 13–4 and 13–6. To convert kilojoules to kilocalories, divide the number
cells requires many steps, results in very large decreases
of kilojoules by 4.184. in standard free energy. However, standard free-energy
13.1 Bioenergetics and Thermodynamics 493

TABLE 13–4 Standard Free-Energy Changes of Some Chemical Reactions


at pH 7.0 and 25 C (298 K)
G

Reaction type (kJ/mol) (kcal/mol)
Hydrolysis reactions
Acid anhydrides
Acetic anhydride  H2O On 2 acetate 91.1 21.8
ATP  H2O 88n ADP  Pi 30.5 7.3
ATP  H2O 88n AMP  PPi 45.6 10.9
PPi  H2O 88n 2Pi 19.2 4.6
UDP-glucose  H2O 88n UMP  glucose 1-phosphate 43.0 10.3
Esters
Ethyl acetate  H2O 88n ethanol  acetate 19.6 4.7
Glucose 6-phosphate  H2O 88n glucose  Pi 13.8 3.3
Amides and peptides
Glutamine  H2O 88n glutamate  NH
4 14.2 3.4
Glycylglycine  H2O 88n 2 glycine 9.2 2.2
Glycosides
Maltose  H2O 88n 2 glucose 15.5 3.7
Lactose  H2O 88n glucose  galactose 15.9 3.8
Rearrangements
Glucose 1-phosphate 88n glucose 6-phosphate 7.3 1.7
Fructose 6-phosphate 88n glucose 6-phosphate 1.7 0.4
Elimination of water
Malate 88n fumarate  H2O 3.1 0.8
Oxidations with molecular oxygen
Glucose  6O2 88n 6CO2  6H2O 2,840 686
Palmitate  23O2 88n 16CO2  16H2O 9,770 2,338

changes such as those in Table 13–4 indicate how much energy change tells us in which direction and how far a
free energy is available from a reaction under standard given reaction must go to reach equilibrium when the
conditions. To describe the energy released under the initial concentration of each component is 1.0 M, the
conditions existing in cells, an expression for the actual pH is 7.0, the temperature is 25 C, and the pressure is
free-energy change is essential. 101.3 kPa. Thus G
is a constant: it has a character-
istic, unchanging value for a given reaction. But the ac-
Actual Free-Energy Changes Depend on Reactant tual free-energy change, G, is a function of reactant
and product concentrations and of the temperature pre-
and Product Concentrations
vailing during the reaction, which will not necessarily
We must be careful to distinguish between two differ- match the standard conditions as defined above. More-
ent quantities: the free-energy change, G, and the stan- over, the G of any reaction proceeding spontaneously
dard free-energy change, G
. Each chemical reaction toward its equilibrium is always negative, becomes less
has a characteristic standard free-energy change, which negative as the reaction proceeds, and is zero at the
may be positive, negative, or zero, depending on the point of equilibrium, indicating that no more work can
equilibrium constant of the reaction. The standard free- be done by the reaction.
494 Chapter 13 Principles of Bioenergetics

DG and DG
for any reaction A  B 3
4 C  D are urable rates. For example, combustion of firewood to
related by the equation CO2 and H2O is very favorable thermodynamically, but
firewood remains stable for years because the activation
[C][D] energy (see Figs 6–2 and 6–3) for the combustion re-
G  G
 RT ln  (13–3)
[A][B]
action is higher than the energy available at room tem-
in which the terms in red are those actually prevail- perature. If the necessary activation energy is provided
ing in the system under observation. The concentration (with a lighted match, for example), combustion will be-
terms in this equation express the effects commonly gin, converting the wood to the more stable products
called mass action, and the term [C][D]/[A][B] is called CO2 and H2O and releasing energy as heat and light. The
the mass-action ratio, Q. As an example, let us sup- heat released by this exothermic reaction provides the
pose that the reaction A  B 3 4 C  D is taking place activation energy for combustion of neighboring regions
at the standard conditions of temperature (25 C) and of the firewood; the process is self-perpetuating.
pressure (101.3 kPa) but that the concentrations of A, In living cells, reactions that would be extremely
B, C, and D are not equal and none of the components slow if uncatalyzed are caused to proceed, not by sup-
is present at the standard concentration of 1.0 M. To de- plying additional heat but by lowering the activation en-
termine the actual free-energy change, G, under these ergy with an enzyme. An enzyme provides an alternative
nonstandard conditions of concentration as the reaction reaction pathway with a lower activation energy than the
proceeds from left to right, we simply enter the actual uncatalyzed reaction, so that at room temperature a large
concentrations of A, B, C, and D in Equation 13–3; the fraction of the substrate molecules have enough thermal
values of R, T, and G
are the standard values. G is energy to overcome the activation barrier, and the re-
negative and approaches zero as the reaction proceeds action rate increases dramatically. The free-energy
because the actual concentrations of A and B decrease change for a reaction is independent of the pathway
and the concentrations of C and D increase. Notice that by which the reaction occurs; it depends only on the
when a reaction is at equilibrium—when there is no nature and concentration of the initial reactants and the
force driving the reaction in either direction and G is final products. Enzymes cannot, therefore, change equi-
zero—Equation 13–3 reduces to librium constants; but they can and do increase the rate
at which a reaction proceeds in the direction dictated by
[C]eq[D]eq thermodynamics.
0  G  G
 RT ln 
[A]eq[B]eq

or Standard Free-Energy Changes Are Additive


G
 RT ln K
eq In the case of two sequential chemical reactions, A 3 4
B and B 3 4 C, each reaction has its own equilibrium
which is the equation relating the standard free-energy
constant and each has its characteristic standard free-
change and equilibrium constant given earlier.
energy change, G1
and G2
. As the two reactions are
The criterion for spontaneity of a reaction is the
sequential, B cancels out to give the overall reaction
value of G, not G
. A reaction with a positive G

A3 4 C, which has its own equilibrium constant and thus
can go in the forward direction if G is negative. This
its own standard free-energy change, G
total. The G

is possible if the term RT ln ([products]/[reactants]) in
values of sequential chemical reactions are additive.
Equation 13–3 is negative and has a larger absolute
For the overall reaction A 3 4 C, G
total is the sum of
value than G
. For example, the immediate removal
the individual standard free-energy changes, G1
and
of the products of a reaction can keep the ratio [prod-
G2
, of the two reactions: G

total  G1
 G2
.
ucts]/[reactants] well below 1, such that the term RT ln
([products]/[reactants]) has a large, negative value. (1) A88nB G1

G
and G are expressions of the maximum (2) B 88nC G2

amount of free energy that a given reaction can theo- Sum: A88nC G1
 G2

retically deliver—an amount of energy that could be
This principle of bioenergetics explains how a ther-
realized only if a perfectly efficient device were avail-
modynamically unfavorable (endergonic) reaction can
able to trap or harness it. Given that no such device is
be driven in the forward direction by coupling it to
possible (some free energy is always lost to entropy dur-
a highly exergonic reaction through a common inter-
ing any process), the amount of work done by the re-
mediate. For example, the synthesis of glucose 6-
action at constant temperature and pressure is always
phosphate is the first step in the utilization of glucose
less than the theoretical amount.
by many organisms:
Another important point is that some thermody-
namically favorable reactions (that is, reactions for Glucose  Pi 88n glucose 6-phosphate  H2O
which G
is large and negative) do not occur at meas- G
 13.8 kJ/mol
13.1 Bioenergetics and Thermodynamics 495

The positive value of G


predicts that under standard cose 6-phosphate synthesis, the K
eq for formation of
conditions the reaction will tend not to proceed spon- glucose 6-phosphate has been raised by a factor of about
taneously in the direction written. Another cellular re- 2  105.
action, the hydrolysis of ATP to ADP and Pi, is very This common-intermediate strategy is employed by
exergonic: all living cells in the synthesis of metabolic intermediates
and cellular components. Obviously, the strategy works
ATP  H2O 88n ADP  Pi G
 30.5 kJ/mol
only if compounds such as ATP are continuously avail-
These two reactions share the common intermediates able. In the following chapters we consider several of the
Pi and H2O and may be expressed as sequential reac- most important cellular pathways for producing ATP.
tions:
(1) Glucose  Pi 88n glucose 6-phosphate  H2O SUMMARY 13.1 Bioenergetics and Thermodynamics
(2) ATP  H2O 88n ADP  Pi
Sum: ATP  glucose 88n ADP  glucose 6-phosphate ■ Living cells constantly perform work. They
require energy for maintaining their highly
The overall standard free-energy change is obtained by organized structures, synthesizing cellular
adding the G
values for individual reactions: components, generating electric currents, and
G
 13.8 kJ/mol  (30.5 kJ/mol)  16.7 kJ/mol many other processes.

The overall reaction is exergonic. In this case, energy ■ Bioenergetics is the quantitative study of
stored in ATP is used to drive the synthesis of glucose energy relationships and energy conversions in
6-phosphate, even though its formation from glucose biological systems. Biological energy
and inorganic phosphate (Pi) is endergonic. The path- transformations obey the laws of
way of glucose 6-phosphate formation by phosphoryl thermodynamics.
transfer from ATP is different from reactions (1) and ■ All chemical reactions are influenced by two
(2) above, but the net result is the same as the sum of forces: the tendency to achieve the most stable
the two reactions. In thermodynamic calculations, all bonding state (for which enthalpy, H, is a
that matters is the state of the system at the beginning useful expression) and the tendency to achieve
of the process and its state at the end; the route be- the highest degree of randomness, expressed
tween the initial and final states is immaterial. as entropy, S. The net driving force in a
We have said that G
is a way of expressing the reaction is G, the free-energy change, which
equilibrium constant for a reaction. For reaction (1) represents the net effect of these two factors:
above, G   H  T S.
[glucose 6-phosphate] ■ The standard transformed free-energy change,
K
eq1    3.9  103 M1 G
, is a physical constant that is
[glucose][Pi]
characteristic for a given reaction and can be
Notice that H2O is not included in this expression, as its calculated from the equilibrium constant for
concentration (55.5 M) is assumed to remain unchanged the reaction: G
 RT ln K
eq.
by the reaction. The equilibrium constant for the hy-
drolysis of ATP is ■ The actual free-energy change, G, is a
variable that depends on G
and on the
[ADP][Pi] concentrations of reactants and products:
K
eq2    2.0  105 M
[ATP] G  G
 RT ln ([products]/[reactants]).
The equilibrium constant for the two coupled reactions ■ When G is large and negative, the reaction
is tends to go in the forward direction; when G
is large and positive, the reaction tends to go in
[glucose 6-phosphate][ADP][Pi]
K
eq3   the reverse direction; and when G  0, the
[glucose][Pi][ATP]
system is at equilibrium.
 (K
eq1)(K
eq2)  (3.9  103 M1) (2.0  105 M)
■ The free-energy change for a reaction is
 7.8  102 independent of the pathway by which the
This calculation illustrates an important point about reaction occurs. Free-energy changes are
equilibrium constants: although the G
values for two additive; the net chemical reaction that results
reactions that sum to a third are additive, the K
eq for from successive reactions sharing a common
a reaction that is the sum of two reactions is the prod- intermediate has an overall free-energy change
uct of their individual K
eq values. Equilibrium constants that is the sum of the G values for the
are multiplicative. By coupling ATP hydrolysis to glu- individual reactions.
496 Chapter 13 Principles of Bioenergetics

13.2 Phosphoryl Group Transfers and ATP O O O


B B B

OO PO O OP OO OP OO O Rib O Adenine
Having developed some fundamental principles of en- A A A
ergy changes in chemical systems, we can now exam- H O O O O ATP 4

ine the energy cycle in cells and the special role of ATP H
as the energy currency that links catabolism and an-
O 1
abolism (see Fig. 1–28). Heterotrophic cells obtain free B hydrolysis,
energy in a chemical form by the catabolism of nutrient  with relief
O OP OOH of charge
molecules, and they use that energy to make ATP from A
repulsion
Pi O
ADP and Pi. ATP then donates some of its chemical en-
resonance
ergy to endergonic processes such as the synthesis of 2 stabilization
metabolic intermediates and macromolecules from
smaller precursors, the transport of substances across  3
O
membranes against concentration gradients, and me- A
chanical motion. This donation of energy from ATP gen- OOP OO  H


A
erally involves the covalent participation of ATP in the O

reaction that is to be driven, with the eventual result
that ATP is converted to ADP and Pi or, in some reac- O O
B B
tions, to AMP and 2 Pi. We discuss here the chemical HO OP OO OP OO O Rib O Adenine
basis for the large free-energy changes that accompany A A
O O ADP2
hydrolysis of ATP and other high-energy phosphate
compounds, and we show that most cases of energy
3 ionization
donation by ATP involve group transfer, not simple hy-
drolysis of ATP. To illustrate the range of energy trans-
O O
ductions in which ATP provides the energy, we consider B B
 
the synthesis of information-rich macromolecules, the H  O OP OO OP O OO Rib O Adenine
A A
transport of solutes across membranes, and motion pro- O O ADP3
duced by muscle contraction.

ATP4  H2O ADP 3  P 2
i  H

The Free-Energy Change for ATP Hydrolysis G


 30.5 kJ/mol
Is Large and Negative
FIGURE 13–1 Chemical basis for the large free-energy change asso-
Figure 13–1 summarizes the chemical basis for the rel- ciated with ATP hydrolysis. 1 The charge separation that results from
atively large, negative, standard free energy of hydrol- hydrolysis relieves electrostatic repulsion among the four negative
ysis of ATP. The hydrolytic cleavage of the terminal charges on ATP. 2 The product inorganic phosphate (Pi) is stabilized
phosphoric acid anhydride (phosphoanhydride) bond in by formation of a resonance hybrid, in which each of the four phos-
ATP separates one of the three negatively charged phorus–oxygen bonds has the same degree of double-bond character
phosphates and thus relieves some of the electrostatic and the hydrogen ion is not permanently associated with any one of
repulsion in ATP; the Pi (HPO42) released is stabilized the oxygens. (Some degree of resonance stabilization also occurs in
by the formation of several resonance forms not possi- phosphates involved in ester or anhydride linkages, but fewer reso-
ble in ATP; and ADP2, the other direct product of nance forms are possible than for Pi.) 3 The product ADP2 imme-
hydrolysis, immediately ionizes, releasing H into a diately ionizes, releasing a proton into a medium of very low [H]
medium of very low [H] (~107 M). Because the con- (pH 7). A fourth factor (not shown) that favors ATP hydrolysis is the
greater degree of solvation (hydration) of the products Pi and ADP rel-
centrations of the direct products of ATP hydrolysis are,
ative to ATP, which further stabilizes the products relative to the re-
in the cell, far below the concentrations at equilibrium
actants.
(Table 13–5), mass action favors the hydrolysis reaction
in the cell.
Although the hydrolysis of ATP is highly exergonic and Pi are not identical and are much lower than the
(G
 30.5 kJ/mol), the molecule is kinetically sta- 1.0 M of standard conditions (Table 13–5). Furthermore,
ble at pH 7 because the activation energy for ATP Mg2 in the cytosol binds to ATP and ADP (Fig. 13–2),
hydrolysis is relatively high. Rapid cleavage of the phos- and for most enzymatic reactions that involve ATP as
phoanhydride bonds occurs only when catalyzed by an phosphoryl group donor, the true substrate is MgATP2.
enzyme. The relevant G
is therefore that for MgATP2 hy-
The free-energy change for ATP hydrolysis is drolysis. Box 13–1 shows how G for ATP hydrolysis in
30.5 kJ/mol under standard conditions, but the actual the intact erythrocyte can be calculated from the data
free energy of hydrolysis (G) of ATP in living cells is in Table 13–5. In intact cells, G for ATP hydrolysis,
very different: the cellular concentrations of ATP, ADP, usually designated Gp, is much more negative than
13.2 Phosphoryl Group Transfers and ATP 497

TABLE 13–5 Adenine Nucleotide, Inorganic Phosphate, and


Phosphocreatine Concentrations in Some Cells
Concentration (mM)*
ATP ADP† AMP Pi PCr
Rat hepatocyte 3.38 1.32 0.29 4.8 0
Rat myocyte 8.05 0.93 0.04 8.05 28
Rat neuron 2.59 0.73 0.06 2.72 4.7
Human erythrocyte 2.25 0.25 0.02 1.65 0
E. coli cell 7.90 1.04 0.82 7.9 0

*For erythrocytes the concentrations are those of the cytosol (human erythrocytes lack a nucleus and mitochondria). In the
other types of cells the data are for the entire cell contents, although the cytosol and the mitochondria have very different
concentrations of ADP. PCr is phosphocreatine, discussed on p. 489.

This value reflects total concentration; the true value for free ADP may be much lower (see Box 13–1).

G
, ranging from 50 to 65 kJ/mol. Gp is often
called the phosphorylation potential. In the follow- O O O
B B B

ing discussions we use the standard free-energy change O OP OO OP O OO PO OO Rib O Adenine
A A A
for ATP hydrolysis, because this allows comparison, on O O O MgATP 2
the same basis, with the energetics of other cellular ø ø2
Mg
reactions. Remember, however, that in living cells G is
the relevant quantity—for ATP hydrolysis and all other O O
B B
reactions—and may be quite different from G
. 
OO PO OOPO OO Rib O Adenine
A A
O O MgADP
Other Phosphorylated Compounds and Thioesters ø ø2
Mg
Also Have Large Free Energies of Hydrolysis
Phosphoenolpyruvate (Fig. 13–3) contains a phosphate FIGURE 13–2 Mg2 and ATP. Formation of Mg2 complexes partially
ester bond that undergoes hydrolysis to yield the enol shields the negative charges and influences the conformation of the
phosphate groups in nucleotides such as ATP and ADP.
form of pyruvate, and this direct product can immedi-
ately tautomerize to the more stable keto form of pyru-
vate. Because the reactant (phosphoenolpyruvate) has
only one form (enol) and the product (pyruvate) has two
possible forms, the product is stabilized relative to the 49.3 kJ/mol), which can, again, be explained in terms
reactant. This is the greatest contributing factor to of the structure of reactant and products. When H2O is
the high standard free energy of hydrolysis of phospho- added across the anhydride bond of 1,3-bisphospho-
enolpyruvate: G
 61.9 kJ/mol. glycerate, one of the direct products, 3-phosphoglyceric
Another three-carbon compound, 1,3-bisphospho- acid, can immediately lose a proton to give the car-
glycerate (Fig. 13–4), contains an anhydride bond be- boxylate ion, 3-phosphoglycerate, which has two equally
tween the carboxyl group at C-1 and phosphoric acid. probable resonance forms (Fig. 13–4). Removal of the
Hydrolysis of this acyl phosphate is accompanied by a direct product (3-phosphoglyceric acid) and formation of
large, negative, standard free-energy change (G
 the resonance-stabilized ion favor the forward reaction.


O O
G J
O P H2 O O O

J D G  
J 
J FIGURE 13–3 Hydrolysis of phosphoenol-
O OC O O O OC OH tautomerization O OC O
G D hydrolysis G D G J pyruvate (PEP). Catalyzed by pyruvate kinase,
C Pi C C
B B A this reaction is followed by spontaneous
CH2 CH2 CH3 tautomerization of the product, pyruvate.
PEP Pyruvate Pyruvate Tautomerization is not possible in PEP, and
(enol form) (keto form) thus the products of hydrolysis are stabilized
relative to the reactants. Resonance
PEP3  H2O pyruvate  P 2
i stabilization of Pi also occurs, as shown
G
 61.9 kJ/mol in Figure 13–1.
498 Chapter 13 Principles of Bioenergetics

 FIGURE 13–4 Hydrolysis of 1,3-


O O
G J bisphosphoglycerate. The direct
P  
D G  product of hydrolysis is 3-phospho-
O O O O OH O O resonance
M D M D G D stabilization glyceric acid, with an undissociated
1C C C
A A A carboxylic acid group, but
Pi H
2 CHOH CHOH CHOH dissociation occurs immediately.
A A A
3 CH2 CH2 CH2 This ionization and the resonance
A H2O A A structures it makes possible stabilize
O O O the product relative to the reactants.
A hydrolysis A ionization A
  
OO PP O O O PP O OO PP O Resonance stabilization of Pi further
A A A
contributes to the negative free-
O O O
energy change.
1,3-Bisphosphoglycerate 3-Phosphoglyceric acid 3-Phosphoglycerate


1,3-Bisphosphoglycerate4  H2O 3-phosphoglycerate3  P 2
i H
G
 49.3 kJ/mol

BOX 13–1 WORKING IN BIOCHEMISTRY

The Free Energy of Hydrolysis of ATP within Cells: much larger than the standard free-energy change
The Real Cost of Doing Metabolic Business (30.5 kJ/mol). By the same token, the free energy
required to synthesize ATP from ADP and Pi under
The standard free energy of hydrolysis of ATP is
the conditions prevailing in the erythrocyte would be
30.5 kJ/mol. In the cell, however, the concentrations
52 kJ/mol.
of ATP, ADP, and Pi are not only unequal but much
Because the concentrations of ATP, ADP, and Pi
lower than the standard 1 M concentrations (see Table
differ from one cell type to another (see Table 13–5),
13–5). Moreover, the cellular pH may differ somewhat
Gp for ATP hydrolysis likewise differs among cells.
from the standard pH of 7.0. Thus the actual free
Moreover, in any given cell, Gp can vary from time
energy of hydrolysis of ATP under intracellular con-
to time, depending on the metabolic conditions in the
ditions (Gp) differs from the standard free-energy
cell and how they influence the concentrations of ATP,
change, G
. We can easily calculate Gp.
ADP, Pi, and H (pH). We can calculate the actual
In human erythrocytes, for example, the concentra-
free-energy change for any given metabolic reaction
tions of ATP, ADP, and Pi are 2.25, 0.25, and 1.65 mM,
as it occurs in the cell, providing we know the con-
respectively. Let us assume for simplicity that the pH
centrations of all the reactants and products of the re-
is 7.0 and the temperature is 25 C, the standard pH
action and know about the other factors (such as pH,
and temperature. The actual free energy of hydrolysis
temperature, and concentration of Mg2) that may af-
of ATP in the erythrocyte under these conditions is
fect the G
and thus the calculated free-energy
given by the relationship
change, Gp.
[ADP][Pi] To further complicate the issue, the total concen-
Gp  G
 RT ln 
[ATP] trations of ATP, ADP, Pi, and H may be substantially
higher than the free concentrations, which are the
Substituting the appropriate values we obtain thermodynamically relevant values. The difference is
Gp  30.5 kJ/mol  due to tight binding of ATP, ADP, and Pi to cellular
proteins. For example, the concentration of free ADP
(0.25  103)(1.65  103)
(8.315 J/mol  K)(298 K) ln 
2.25  103  in resting muscle has been variously estimated at be-
tween 1 and 37 M. Using the value 25 M in the cal-
culation outlined above, we get a Gp of 58 kJ/mol.
 30.5 kJ/mol  (2.48 kJ/mol) ln 1.8  104
Calculation of the exact value of Gp is perhaps
 30.5 kJ/mol  21 kJ/mol
less instructive than the generalization we can make
 52 kJ/mol
about actual free-energy changes: in vivo, the energy
Thus Gp, the actual free-energy change for ATP hy- released by ATP hydrolysis is greater than the stan-
drolysis in the intact erythrocyte (52 kJ/mol), is dard free-energy change, G
.
13.2 Phosphoryl Group Transfers and ATP 499

COO COO COO FIGURE 13–5 Hydrolysis of phospho-


A A A creatine. Breakage of the PON bond

O CH2 H2O CH2 CH2
B H A A H2N A in phosphocreatine produces creatine,

O OP ONOCONOCH3 H2NOCONOCH3 C NOCH3 which is stabilized by formation of a
A B hydrolysis B resonance 
H2N
O 
NH2 Pi

NH2 stabilization

resonance hybrid. The other product,
Phosphocreatine Creatine Pi, is also resonance stabilized.

Phosphocreatine  H2O 2
creatine  P 2
i
G
 43.0 kJ/mol

In phosphocreatine (Fig. 13–5), the PON bond can these compounds is activated for transacylation, con-
be hydrolyzed to generate free creatine and Pi. The re- densation, or oxidation-reduction reactions. Thioesters
lease of Pi and the resonance stabilization of creatine undergo much less resonance stabilization than do oxy-
favor the forward reaction. The standard free-energy gen esters; consequently, the difference in free energy
change of phosphocreatine hydrolysis is again large, between the reactant and its hydrolysis products, which
43.0 kJ/mol. are resonance-stabilized, is greater for thioesters than
In all these phosphate-releasing reactions, the sev- for comparable oxygen esters (Fig. 13–7). In both cases,
eral resonance forms available to Pi (Fig. 13–1) stabi- hydrolysis of the ester generates a carboxylic acid,
lize this product relative to the reactant, contributing to which can ionize and assume several resonance forms.
an already negative free-energy change. Table 13–6 lists Together, these factors result in the large, negative G

the standard free energies of hydrolysis for a number of (31 kJ/mol) for acetyl-CoA hydrolysis.
phosphorylated compounds. To summarize, for hydrolysis reactions with large,
Thioesters, in which a sulfur atom replaces the negative, standard free-energy changes, the products
usual oxygen in the ester bond, also have large, nega- are more stable than the reactants for one or more of
tive, standard free energies of hydrolysis. Acetyl-coen- the following reasons: (1) the bond strain in reactants
zyme A, or acetyl-CoA (Fig. 13–6), is one of many due to electrostatic repulsion is relieved by charge sep-
thioesters important in metabolism. The acyl group in aration, as for ATP; (2) the products are stabilized by

TABLE 13–6 Standard Free Energies of


Hydrolysis of Some Phosphorylated Compounds O
and Acetyl-CoA (a Thioester) CH3 OC
J
Acetyl-CoA
G
S-CoA
G

H2O hydrolysis
(kJ/mol) (kcal/mol)
CoASH
Phosphoenolpyruvate 61.9 14.8 O
1,3-bisphosphoglycerate J
CH3 OC Acetic acid
(n 3-phosphoglycerate  Pi) 49.3 11.8 G
OH
Phosphocreatine 43.0 10.3
ADP (n AMP  Pi) 32.8 7.8 ionization

ATP (n ADP  Pi) 30.5 7.3 H


ATP (n AMP  PPi) 45.6 10.9 O


AMP (n adenosine  Pi) 14.2 3.4 CH3 OC


D
Acetate
PPi (n 2Pi) 19.2 4.0 G 
O
Glucose 1-phosphate 20.9 5.0 resonance
Fructose 6-phosphate 15.9 3.8 stabilization
Glucose 6-phosphate 13.8 3.3
Acetyl-CoA  H2O acetate  CoA  H
Glycerol 1-phosphate 9.2 2.2
G
 31.4 kJ/mol
Acetyl-CoA 31.4 7.5
FIGURE 13–6 Hydrolysis of acetyl-coenzyme A. Acetyl-CoA is a
Source: Data mostly from Jencks, W.P. (1976) in Handbook of Biochemistry and Molecular thioester with a large, negative, standard free energy of hydrolysis.
Biology, 3rd edn (Fasman, G.D., ed.), Physical and Chemical Data, Vol. I, pp. 296–304,
Thioesters contain a sulfur atom in the position occupied by an oxy-
CRC Press, Boca Raton, FL. The value for the free energy of hydrolysis of PPi is from Frey,
P.A. & Arabshahi, A. (1995) Standard free-energy change for the hydrolysis of the –- gen atom in oxygen esters. The complete structure of coenzyme A
phosphoanhydride bridge in ATP. Biochemistry 34, 11,307–11,310. (CoA, or CoASH) is shown in Figure 8–41.
500 Chapter 13 Principles of Bioenergetics

Thioester Extra stabilization of


oxygen ester by resonance
O
J
CH3 OC
G Oxygen
S OR FIGURE 13–7 Free energy of hydrolysis
ester 
Free energy, G

O resonance for thioesters and oxygen esters. The


O
J stabilization products of both types of hydrolysis
CH3 OC CH3 C
G for G reaction have about the same free-energy
thioester O OR OO R
hydrolysis  content (G), but the thioester has a higher
G for oxygen free-energy content than the oxygen ester.
ester hydrolysis Orbital overlap between the O and C
atoms allows resonance stabilization
O O
J J in oxygen esters; orbital overlap between
CH3 OC  ROSH CH3 OC  RO OH
G G S and C atoms is poorer and provides
OH OH
little resonance stabilization.

ionization, as for ATP, acyl phosphates, and thioesters; chanical motion. This occurs in muscle contraction and
(3) the products are stabilized by isomerization (tau- in the movement of enzymes along DNA or of ribosomes
tomerization), as for phosphoenolpyruvate; and/or (4) along messenger RNA. The energy-dependent reactions
the products are stabilized by resonance, as for creatine catalyzed by helicases, RecA protein, and some topo-
released from phosphocreatine, carboxylate ion re- isomerases (Chapter 25) also involve direct hydrolysis
leased from acyl phosphates and thioesters, and phos- of phosphoanhydride bonds. GTP-binding proteins that
phate (Pi) released from anhydride or ester linkages. act in signaling pathways directly hydrolyze GTP to
drive conformational changes that terminate signals
ATP Provides Energy by Group Transfers,
Not by Simple Hydrolysis (a) Written as a one-step reaction

Throughout this book you will encounter reactions or COO COO


processes for which ATP supplies energy, and the con- A
 A
H3NOCH ATP ADP  Pi H3NOCH
tribution of ATP to these reactions is commonly indi- A A
cated as in Figure 13–8a, with a single arrow showing CH2 CH2
A  NH3 A
the conversion of ATP to ADP and Pi (or, in some cases, CH2 CH2
of ATP to AMP and pyrophosphate, PPi). When written A A
C C
this way, these reactions of ATP appear to be simple hy- J G  J G
drolysis reactions in which water displaces Pi (or PPi), O O O NH2
Glutamate Glutamine
and one is tempted to say that an ATP-dependent re-
action is “driven by the hydrolysis of ATP.” This is not ATP NH3
the case. ATP hydrolysis per se usually accomplishes ADP COO
nothing but the liberation of heat, which cannot drive a A
H3NOCH
chemical process in an isothermal system. A single re- 1 A 2
CH2 Pi
action arrow such as that in Figure 13–8a almost in- A
variably represents a two-step process (Fig. 13–8b) in CH2
which part of the ATP molecule, a phosphoryl or py- A
C
rophosphoryl group or the adenylate moiety (AMP), is J G
O O O
first transferred to a substrate molecule or to an amino G J
P
acid residue in an enzyme, becoming covalently at- 
D G 
O O
tached to the substrate or the enzyme and raising its
Enzyme-bound
free-energy content. Then, in a second step, the phos- glutamyl phosphate
phate-containing moiety transferred in the first step is
displaced, generating Pi, PPi, or AMP. Thus ATP partic- (b) Actual two-step reaction
ipates covalently in the enzyme-catalyzed reaction to
which it contributes free energy. FIGURE 13–8 ATP hydrolysis in two steps. (a) The contribution of
Some processes do involve direct hydrolysis of ATP ATP to a reaction is often shown as a single step, but is almost always
(or GTP), however. For example, noncovalent binding a two-step process. (b) Shown here is the reaction catalyzed by ATP-
of ATP (or of GTP), followed by its hydrolysis to ADP dependent glutamine synthetase. 1 A phosphoryl group is transferred
(or GDP) and Pi, can provide the energy to cycle some from ATP to glutamate, then 2 the phosphoryl group is displaced by
proteins between two conformations, producing me- NH3 and released as Pi.
13.2 Phosphoryl Group Transfers and ATP 501

triggered by hormones or by other extracellular factors direct donation of a phosphoryl group from PEP to ADP
(Chapter 12). is thermodynamically feasible:
The phosphate compounds found in living organisms G
(kJ/mol)
can be divided somewhat arbitrarily into two groups,
(1) PEP  H2O 8n pyruvate  Pi 61.9
based on their standard free energies of hydrolysis
(2) ADP  Pi 8n ATP  H2O 30.5
(Fig. 13–9). “High-energy” compounds have a G
of
Sum: PEP  ADP 8n pyruvate  ATP 31.4
hydrolysis more negative than 25 kJ/mol; “low-energy”
compounds have a less negative G
. Based on this cri- Notice that while the overall reaction above is repre-
terion, ATP, with a G
of hydrolysis of 30.5 kJ/mol sented as the algebraic sum of the first two reactions,
(7.3 kcal/mol), is a high-energy compound; glucose the overall reaction is actually a third, distinct reaction
6-phosphate, with a G
of hydrolysis of 13.8 kJ/mol that does not involve Pi; PEP donates a phosphoryl
(3.3 kcal/mol), is a low-energy compound. group directly to ADP. We can describe phosphorylated
The term “high-energy phosphate bond,” long used compounds as having a high or low phosphoryl group
by biochemists to describe the POO bond broken in hy- transfer potential, on the basis of their standard free en-
drolysis reactions, is incorrect and misleading as it ergies of hydrolysis (as listed in Table 13–6). The phos-
wrongly suggests that the bond itself contains the en- phoryl group transfer potential of phosphoenolpyruvate
ergy. In fact, the breaking of all chemical bonds requires is very high, that of ATP is high, and that of glucose 6-
an input of energy. The free energy released by hy- phosphate is low (Fig. 13–9).
drolysis of phosphate compounds does not come from Much of catabolism is directed toward the synthesis
the specific bond that is broken; it results from the prod- of high-energy phosphate compounds, but their forma-
ucts of the reaction having a lower free-energy content tion is not an end in itself; they are the means of acti-
than the reactants. For simplicity, we will sometimes use vating a very wide variety of compounds for further
the term “high-energy phosphate compound” when re- chemical transformation. The transfer of a phosphoryl
ferring to ATP or other phosphate compounds with a group to a compound effectively puts free energy into
large, negative, standard free energy of hydrolysis. that compound, so that it has more free energy to give
As is evident from the additivity of free-energy up during subsequent metabolic transformations. We de-
changes of sequential reactions, any phosphorylated scribed above how the synthesis of glucose 6-phosphate
compound can be synthesized by coupling the synthe- is accomplished by phosphoryl group transfer from ATP.
sis to the breakdown of another phosphorylated com- In the next chapter we see how this phosphorylation of
pound with a more negative free energy of hydrolysis. glucose activates, or “primes,” the glucose for catabolic
For example, because cleavage of Pi from phospho- reactions that occur in nearly every living cell. Because
enolpyruvate (PEP) releases more energy than is of its intermediate position on the scale of group trans-
needed to drive the condensation of Pi with ADP, the fer potential, ATP can carry energy from high-energy

70
COO
A
CO OO P Phosphoenolpyruvate
B
60
CH2
O OO P
M D
C
A
50 CHOH P O Creatine
G
of hydrolysis (kJ/mol)

A
CH2OOO P Phosphocreatine
40 1,3-Bisphosphoglycerate
FIGURE 13–9 Ranking of biological phosphate
Adenine O Rib O P O P O P compounds by standard free energies of hydrol-
30 High-energy ysis. This shows the flow of phosphoryl groups,
ATP
compounds represented by P , from high-energy phosphoryl
Low-energy donors via ATP to acceptor molecules (such as
20 compounds glucose and glycerol) to form their low-energy
phosphate derivatives. This flow of phosphoryl
Glucose 6- P Glycerol- P
groups, catalyzed by enzymes called kinases,
10 proceeds with an overall loss of free energy
under intracellular conditions. Hydrolysis of low-
energy phosphate compounds releases Pi, which
0 Pi has an even lower phosphoryl group transfer
potential (as defined in the text).
502 Chapter 13 Principles of Bioenergetics

phosphate compounds produced by catabolism to com- (p. 218) involves attack at the  position of the ATP
pounds such as glucose, converting them into more re- molecule.
active species. ATP thus serves as the universal energy Attack at the  phosphate of ATP displaces AMP and
currency in all living cells. transfers a pyrophosphoryl (not pyrophosphate) group
One more chemical feature of ATP is crucial to its to the attacking nucleophile (Fig. 13–10b). For exam-
role in metabolism: although in aqueous solution ATP is ple, the formation of 5
-phosphoribosyl-1-pyrophosphate
thermodynamically unstable and is therefore a good (p. XXX), a key intermediate in nucleotide synthesis,
phosphoryl group donor, it is kinetically stable. Because results from attack of an OOH of the ribose on the 
of the huge activation energies (200 to 400 kJ/mol) re- phosphate.
quired for uncatalyzed cleavage of its phosphoanhydride Nucleophilic attack at the  position of ATP displaces
bonds, ATP does not spontaneously donate phosphoryl PPi and transfers adenylate (5
-AMP) as an adenylyl
groups to water or to the hundreds of other potential group (Fig. 13–10c); the reaction is an adenylylation
acceptors in the cell. Only when specific enzymes are (a-den
-i-li-la-
-shun, probably the most ungainly word
present to lower the energy of activation does phos- in the biochemical language). Notice that hydrolysis of
phoryl group transfer from ATP proceed. The cell is the – phosphoanhydride bond releases considerably
therefore able to regulate the disposition of the energy more energy (~46 kJ/mol) than hydrolysis of the –
carried by ATP by regulating the various enzymes that bond (~31 kJ/mol) (Table 13–6). Furthermore, the PPi
act on it. formed as a byproduct of the adenylylation is hydrolyzed
to two Pi by the ubiquitous enzyme inorganic pyro-
ATP Donates Phosphoryl, Pyrophosphoryl, phosphatase, releasing 19 kJ/mol and thereby provid-
ing a further energy “push” for the adenylylation reac-
and Adenylyl Groups
tion. In effect, both phosphoanhydride bonds of ATP are
The reactions of ATP are generally SN2 nucleophilic dis- split in the overall reaction. Adenylylation reactions are
placements (p. II.8), in which the nucleophile may be, therefore thermodynamically very favorable. When the
for example, the oxygen of an alcohol or carboxylate, or energy of ATP is used to drive a particularly unfavor-
a nitrogen of creatine or of the side chain of arginine or able metabolic reaction, adenylylation is often the mech-
histidine. Each of the three phosphates of ATP is sus- anism of energy coupling. Fatty acid activation is a good
ceptible to nucleophilic attack (Fig. 13–10), and each example of this energy-coupling strategy.
position of attack yields a different type of product. The first step in the activation of a fatty acid—
Nucleophilic attack by an alcohol on the  phos- either for energy-yielding oxidation or for use in the syn-
phate (Fig. 13–10a) displaces ADP and produces a new thesis of more complex lipids—is the formation of its
phosphate ester. Studies with 18O-labeled reactants thiol ester (see Fig. 17–5). The direct condensation of
have shown that the bridge oxygen in the new com- a fatty acid with coenzyme A is endergonic, but the for-
pound is derived from the alcohol, not from ATP; the mation of fatty acyl–CoA is made exergonic by stepwise
group transferred from ATP is a phosphoryl (OPO32), removal of two phosphoryl groups from ATP. First,
not a phosphate (OOPO32). Phosphoryl group transfer adenylate (AMP) is transferred from ATP to the car-
from ATP to glutamate (Fig. 13–8) or to glucose boxyl group of the fatty acid, forming a mixed anhydride

Three positions on ATP for attack by the nucleophile R18O


 
O O O FIGURE 13–10 Nucleophilic displacement reac-
tions of ATP. Any of the three P atoms (, , or )
O P O P O P O Rib Adenine may serve as the electrophilic target for
O O O nucleophilic attack—in this case, by the labeled
nucleophile RO18O:. The nucleophile may be an
R18O R18O R18O
alcohol (ROH), a carboxyl group (RCOO), or a
phosphoanhydride (a nucleoside mono- or
diphosphate, for example). (a) When the oxygen
O O O O of the nucleophile attacks the  position, the bridge
oxygen of the product is labeled, indicating that
R18O P O R18O P O P O R18O P O Rib Adenine the group transferred from ATP is a phosphoryl
O O O O (OPO32), not a phosphate (OOPO32). (b) Attack
   on the  position displaces AMP and leads to the
ADP AMP PPi transfer of a pyrophosphoryl (not pyrophosphate)
Phosphoryl Pyrophosphoryl Adenylyl group to the nucleophile. (c) Attack on the 
transfer transfer transfer position displaces PPi and transfers the adenylyl
(a) (b) (c) group to the nucleophile.
13.2 Phosphoryl Group Transfers and ATP 503

(fatty acyl adenylate) and liberating PPi. The thiol group for the breakage of these bonds, or 45.6 kJ/mol 
of coenzyme A then displaces the adenylate group and (19.2) kJ/mol:
forms a thioester with the fatty acid. The sum of these
ATP  2H2O 88n AMP  2Pi G
 64.8 kJ/mol
two reactions is energetically equivalent to the exer-
gonic hydrolysis of ATP to AMP and PPi (G
 45.6 The activation of amino acids before their polymer-
kJ/mol) and the endergonic formation of fatty acyl–CoA ization into proteins (see Fig. 27–14) is accomplished
(G
 31.4 kJ/mol). The formation of fatty acyl–CoA by an analogous set of reactions in which a transfer RNA
is made energetically favorable by hydrolysis of the PPi molecule takes the place of coenzyme A. An interesting
by inorganic pyrophosphatase. Thus, in the activation use of the cleavage of ATP to AMP and PPi occurs in
of a fatty acid, both phosphoanhydride bonds of ATP are the firefly, which uses ATP as an energy source to pro-
broken. The resulting G
is the sum of the G
values duce light flashes (Box 13–2).

BOX 13–2 THE WORLD OF BIOCHEMISTRY

Firefly Flashes: Glowing Reports of ATP pyrophosphate cleavage of ATP to form luciferyl
Bioluminescence requires considerable amounts of adenylate. In the presence of molecular oxygen and
energy. In the firefly, ATP is used in a set of reactions luciferase, the luciferin undergoes a multistep oxida-
that converts chemical energy into light energy. In the tive decarboxylation to oxyluciferin. This process is
1950s, from many thousands of fireflies collected by accompanied by emission of light. The color of the
children in and around Baltimore, William McElroy light flash differs with the firefly species and appears
and his colleagues at The Johns Hopkins University to be determined by differences in the structure of the
isolated the principal biochemical components: lu- luciferase. Luciferin is regenerated from oxyluciferin
ciferin, a complex carboxylic acid, and luciferase, an in a subsequent series of reactions.
enzyme. The generation of a light flash requires acti- In the laboratory, pure firefly luciferin and lu-
vation of luciferin by an enzymatic reaction involving ciferase are used to measure minute quantities of ATP
by the intensity of the light flash produced. As little
as a few picomoles (1012 mol) of ATP can be meas-
ured in this way. An enlightening extension of the
studies in luciferase was the cloning of the luciferase
gene into tobacco plants. When watered with a solu-
tion containing luciferin, the plants glowed in the dark
(see Fig. 9–29).

H O
N N A
C O P O Rib Adenine
H
HO S S O O
H AMP
Luciferyl adenylate

PPi
O2

luciferase
ATP light
The firefly, a beetle of the Lampyridae family.
H
N N
COO
H CO2  AMP
HO S S
H
Luciferin N N O

HO S S
Important components in the regenerating Oxyluciferin
firefly bioluminescence cycle. reactions
504 Chapter 13 Principles of Bioenergetics

Assembly of Informational Macromolecules relax into a second conformation until another molecule
Requires Energy of ATP binds. The binding and subsequent hydrolysis of
ATP (by myosin ATPase) provide the energy that forces
When simple precursors are assembled into high mo- cyclic changes in the conformation of the myosin head.
lecular weight polymers with defined sequences (DNA, The change in conformation of many individual myosin
RNA, proteins), as described in detail in Part III, energy
is required both for the condensation of monomeric
units and for the creation of ordered sequences. The
precursors for DNA and RNA synthesis are nucleoside O
RNA A
triphosphates, and polymerization is accompanied by OPP OO
chain
cleavage of the phosphoanhydride linkage between the A
O
 and  phosphates, with the release of PPi (Fig. 13–11). A
The moieties transferred to the growing polymer in CH2 O Base
these reactions are adenylate (AMP), guanylate (GMP),
H H
cytidylate (CMP), or uridylate (UMP) for RNA synthe- H H
sis, and their deoxy analogs (with TMP in place of UMP)
for DNA synthesis. As noted above, the activation of SOH OH

amino acids for protein synthesis involves the donation O O O
B B A
of adenylate groups from ATP, and we shall see in Chap- 
OO PO O OP OO OP PO
ter 27 that several steps of protein synthesis on the ri- A A A
O O O
bosome are also accompanied by GTP hydrolysis. In all A
these cases, the exergonic breakdown of a nucleoside CH2 O Guanine GTP
first anhydride
triphosphate is coupled to the endergonic process of bond broken
H H
synthesizing a polymer of a specific sequence. H H
O O OH OH
ATP Energizes Active Transport 
B B
OOP OOOP OO
and Muscle Contraction A A
 
O O
ATP can supply the energy for transporting an ion or a PPi
molecule across a membrane into another aqueous com- second
partment where its concentration is higher (see Fig. anhydride
bond
11–36). Transport processes are major consumers of en- broken
ergy; in human kidney and brain, for example, as much
as two-thirds of the energy consumed at rest is used to 2 Pi O
pump Na and K across plasma membranes via the A

O P OO
P
NaK ATPase. The transport of Na and K is driven A
by cyclic phosphorylation and dephosphorylation of the O
A
transporter protein, with ATP as the phosphoryl group CH2 O Base
donor (see Fig. 11–37). Na-dependent phosphorylation
of the NaK ATPase forces a change in the protein’s H H
H H
conformation, and K-dependent dephosphorylation
favors return to the original conformation. Each cycle in O OH
A
the transport process results in the conversion of ATP 
OO PP O
to ADP and Pi, and it is the free-energy change of ATP A
RNA chain O
hydrolysis that drives the cyclic changes in protein con- A
lengthened
formation that result in the electrogenic pumping of Na by one
CH2 O Guanine
and K. Note that in this case ATP interacts covalently nucleotide
H H
by phosphoryl group transfer to the enzyme, not the H H
substrate.
OH OH
In the contractile system of skeletal muscle cells,
myosin and actin are specialized to transduce the chem- FIGURE 13–11 Nucleoside triphosphates in RNA synthesis. With
ical energy of ATP into motion (see Fig. 5–33). ATP each nucleoside monophosphate added to the growing chain, one PPi
binds tightly but noncovalently to one conformation of is released and hydrolyzed to two Pi. The hydrolysis of two phospho-
myosin, holding the protein in that conformation. When anhydride bonds for each nucleotide added provides the energy for
myosin catalyzes the hydrolysis of its bound ATP, the forming the bonds in the RNA polymer and for assembling a specific
ADP and Pi dissociate from the protein, allowing it to sequence of nucleotides.
13.2 Phosphoryl Group Transfers and ATP 505

molecules results in the sliding of myosin fibrils along intermediate; then the phosphoryl group is transferred
actin filaments (see Fig. 5–32), which translates into from the P –His residue to an NDP acceptor. Because
macroscopic contraction of the muscle fiber. the enzyme is nonspecific for the base in the NDP and
As we noted earlier, this production of mechanical works equally well on dNDPs and NDPs, it can synthe-
motion at the expense of ATP is one of the few cases in size all NTPs and dNTPs, given the corresponding NDPs
which ATP hydrolysis per se, rather than group trans- and a supply of ATP.
fer from ATP, is the source of the chemical energy in a Phosphoryl group transfers from ATP result in an
coupled process. accumulation of ADP; for example, when muscle is con-
tracting vigorously, ADP accumulates and interferes
Transphosphorylations between Nucleotides with ATP-dependent contraction. During periods of in-
Occur in All Cell Types tense demand for ATP, the cell lowers the ADP con-
centration, and at the same time acquires ATP, by the
Although we have focused on ATP as the cell’s energy action of adenylate kinase:
currency and donor of phosphoryl groups, all other nu-
cleoside triphosphates (GTP, UTP, and CTP) and all the Mg2
2ADP 3:::4 ATP  AMP DG
 0
deoxynucleoside triphosphates (dATP, dGTP, dTTP, and
dCTP) are energetically equivalent to ATP. The free- This reaction is fully reversible, so after the intense de-
energy changes associated with hydrolysis of their mand for ATP ends, the enzyme can recycle AMP by
phosphoanhydride linkages are very nearly identical converting it to ADP, which can then be phosphorylated
with those shown in Table 13–6 for ATP. In preparation to ATP in mitochondria. A similar enzyme, guanylate ki-
for their various biological roles, these other nucleotides nase, converts GMP to GDP at the expense of ATP. By
are generated and maintained as the nucleoside triphos- pathways such as these, energy conserved in the cata-
phate (NTP) forms by phosphoryl group transfer to the bolic production of ATP is used to supply the cell with
corresponding nucleoside diphosphates (NDPs) and all required NTPs and dNTPs.
monophosphates (NMPs). Phosphocreatine (Fig. 13–5), also called creatine
ATP is the primary high-energy phosphate com- phosphate, serves as a ready source of phosphoryl
pound produced by catabolism, in the processes of gly- groups for the quick synthesis of ATP from ADP. The
colysis, oxidative phosphorylation, and, in photosyn- phosphocreatine (PCr) concentration in skeletal mus-
thetic cells, photophosphorylation. Several enzymes cle is approximately 30 mM, nearly ten times the con-
then carry phosphoryl groups from ATP to the other nu- centration of ATP, and in other tissues such as smooth
cleotides. Nucleoside diphosphate kinase, found in muscle, brain, and kidney [PCr] is 5 to 10 mM. The en-
all cells, catalyzes the reaction zyme creatine kinase catalyzes the reversible reaction
Mg2 Mg2
ATP  NDP (or dNDP) 3:::4 ADP  NTP (or dNTP) ADP  PCr 3:::4 ATP  Cr DG
 12.5 kJ/mol
DG
 0
When a sudden demand for energy depletes ATP, the
Although this reaction is fully reversible, the relatively PCr reservoir is used to replenish ATP at a rate consid-
high [ATP]/[ADP] ratio in cells normally drives the re- erably faster than ATP can be synthesized by catabolic
action to the right, with the net formation of NTPs and pathways. When the demand for energy slackens, ATP
dNTPs. The enzyme actually catalyzes a two-step phos- produced by catabolism is used to replenish the PCr
phoryl transfer, which is a classic case of a double-dis- reservoir by reversal of the creatine kinase reaction. Or-
placement (Ping-Pong) mechanism (Fig. 13–12; see also ganisms in the lower phyla employ other PCr-like mole-
Fig. 6–13b). First, phosphoryl group transfer from ATP cules (collectively called phosphagens) as phosphoryl
to an active-site His residue produces a phosphoenzyme reservoirs.

Adenosine P P P Enz His Nucleoside P P P


(ATP) (any NTP or dNTP)
Ping Pong

Adenosine P P Enz His P Nucleoside P P


(ADP) (any NDP or dNDP)

FIGURE 13–12 Ping-Pong mechanism of nucleoside diphosphate phate replaces it, and this is converted to the corresponding triphos-
kinase. The enzyme binds its first substrate (ATP in our example), and phate by transfer of the phosphoryl group from the phosphohistidine
a phosphoryl group is transferred to the side chain of a His residue. residue.
ADP departs, and another nucleoside (or deoxynucleoside) diphos-
506 Chapter 13 Principles of Bioenergetics

Inorganic Polyphosphate Is a Potential fers a survival advantage. Deletion of the genes for
Phosphoryl Group Donor polyphosphate kinases diminishes the ability of certain
pathogenic bacteria to invade animal tissues. The en-
Inorganic polyphosphate (polyP) is a linear polymer zymes may therefore prove to be vulnerable targets in
composed of many tens or hundreds of Pi residues the development of new antimicrobial drugs.
linked through phosphoanhydride bonds. This polymer, No gene in yeast encodes a PPK-like protein, but
present in all organisms, may accumulate to high levels four genes—unrelated to bacterial PPK genes—are nec-
in some cells. In yeast, for example, the amount of polyP essary for the synthesis of polyphosphate. The mecha-
that accumulates in the vacuoles would represent, if dis- nism for polyphosphate synthesis in eukaryotes seems
tributed uniformly throughout the cell, a concentration to be quite different from that in prokaryotes.
of 200 mM! (Compare this with the concentrations of
other phosphoryl donors listed in Table 13–5.) Biochemical and Chemical Equations
Are Not Identical
O O O O O
O P O P O P O P O P O Biochemists write metabolic equations in a simplified
way, and this is particularly evident for reactions in-
O O O O O volving ATP. Phosphorylated compounds can exist in
Inorganic polyphosphate (polyP)
several ionization states and, as we have noted, the dif-
ferent species can bind Mg2. For example, at pH 7 and
One potential role for polyP is to serve as a phos-
2 mM Mg2, ATP exists in the forms ATP4, HATP3,
phagen, a reservoir of phosphoryl groups that can be
H2ATP2, MgHATP, and Mg2ATP. In thinking about the
used to generate ATP, as creatine phosphate is used in
biological role of ATP, however, we are not always in-
muscle. PolyP has about the same phosphoryl group
terested in all this detail, and so we consider ATP as an
transfer potential as PPi. The shortest polyphosphate,
entity made up of a sum of species, and we write its hy-
PPi (n  2), can serve as the energy source for active
drolysis as the biochemical equation
transport of H in plant vacuoles. For at least one form
of the enzyme phosphofructokinase in plants, PPi is the ATP  H2O 8n ADP  Pi
phosphoryl group donor, a role played by ATP in ani-
where ATP, ADP, and Pi are sums of species. The
mals and microbes (p. XXX). The finding of high con-
corresponding apparent equilibrium constant, K
eq 
centrations of polyP in volcanic condensates and steam
[ADP][Pi]/[ATP], depends on the pH and the concentra-
vents suggests that it could have served as an energy
tion of free Mg2. Note that H and Mg2 do not ap-
source in prebiotic and early cellular evolution.
pear in the biochemical equation because they are held
In prokaryotes, the enzyme polyphosphate ki-
constant. Thus a biochemical equation does not balance
nase-1 (PPK-1) catalyzes the reversible reaction
H, Mg, or charge, although it does balance all other el-
Mg2 ements involved in the reaction (C, N, O, and P in the
ATP  polyPn 3:::4 ADP  polyPn1 equation above).
DG
 20 kJ/mol We can write a chemical equation that does balance
by a mechanism involving an enzyme-bound phospho- for all elements and for charge. For example, when ATP
histidine intermediate (recall the mechanism of nucle- is hydrolyzed at a pH above 8.5 in the absence of Mg2,
oside diphosphate kinase, described above). A second the chemical reaction is represented by
enzyme, polyphosphate kinase-2 (PPK-2), catalyzes ATP4  H2O 8n ADP3  HPO42  H
the reversible synthesis of GTP (or ATP) from poly-
phosphate and GDP (or ADP): The corresponding equilibrium constant, K
eq 
[ADP3][HPO42][H]/[ATP4], depends only on tem-
Mn2 perature, pressure, and ionic strength.
GDP  polyPn1 3:::4 GTP  polyPn
Both ways of writing a metabolic reaction have value
PPK-2 is believed to act primarily in the direction of in biochemistry. Chemical equations are needed when
GTP and ATP synthesis, and PPK-1 in the direction of we want to account for all atoms and charges in a re-
polyphosphate synthesis. PPK-1 and PPK-2 are present action, as when we are considering the mechanism of a
in a wide variety of prokaryotes, including many patho- chemical reaction. Biochemical equations are used to
genic bacteria. determine in which direction a reaction will proceed
In prokaryotes, elevated levels of polyP have been spontaneously, given a specified pH and [Mg2], or to
shown to promote expression of a number of genes in- calculate the equilibrium constant of such a reaction.
volved in adaptation of the organism to conditions of Throughout this book we use biochemical equa-
starvation or other threats to survival. In Escherichia tions, unless the focus is on chemical mechanism, and
coli, for example, polyP accumulates when cells are we use values of G
and K
eq as determined at pH 7
starved for amino acids or Pi, and this accumulation con- and 1 mM Mg2.
13.3 Biological Oxidation-Reduction Reactions 507

SUMMARY 13.2 Phosphoryl Group Transfers The carriers in turn donate electrons to acceptors with
and ATP higher electron affinities, with the release of energy.
Cells contain a variety of molecular energy transducers,
■ ATP is the chemical link between catabolism which convert the energy of electron flow into useful
and anabolism. It is the energy currency of the work.
living cell. The exergonic conversion of ATP to We begin our discussion with a description of the
ADP and Pi, or to AMP and PPi, is coupled to general types of metabolic reactions in which electrons
many endergonic reactions and processes. are transferred. After considering the theoretical and
experimental basis for measuring the energy changes in
■ Direct hydrolysis of ATP is the source of oxidation reactions in terms of electromotive force, we
energy in the conformational changes that discuss the relationship between this force, expressed
produce muscle contraction but, in general, it in volts, and the free-energy change, expressed in joules.
is not ATP hydrolysis but the transfer of a We conclude by describing the structures and oxidation-
phosphoryl, pyrophosphoryl, or adenylyl group reduction chemistry of the most common of the spe-
from ATP to a substrate or enzyme molecule cialized electron carriers, which you will encounter
that couples the energy of ATP breakdown to repeatedly in later chapters.
endergonic transformations of substrates.
■ Through these group transfer reactions, ATP The Flow of Electrons Can Do Biological Work
provides the energy for anabolic reactions,
including the synthesis of informational Every time we use a motor, an electric light or heater,
molecules, and for the transport of molecules or a spark to ignite gasoline in a car engine, we use the
and ions across membranes against flow of electrons to accomplish work. In the circuit that
concentration gradients and electrical potential powers a motor, the source of electrons can be a bat-
gradients. tery containing two chemical species that differ in affin-
ity for electrons. Electrical wires provide a pathway for
■ Cells contain other metabolites with large, electron flow from the chemical species at one pole of
negative, free energies of hydrolysis, including the battery, through the motor, to the chemical species
phosphoenolpyruvate, 1,3-bisphosphoglycerate, at the other pole of the battery. Because the two chem-
and phosphocreatine. These high-energy ical species differ in their affinity for electrons, electrons
compounds, like ATP, have a high phosphoryl flow spontaneously through the circuit, driven by a force
group transfer potential; they are good donors proportional to the difference in electron affinity, the
of the phosphoryl group. Thioesters also have electromotive force (emf). The electromotive force
high free energies of hydrolysis. (typically a few volts) can accomplish work if an ap-
■ Inorganic polyphosphate, present in all cells, propriate energy transducer—in this case a motor—is
may serve as a reservoir of phosphoryl groups placed in the circuit. The motor can be coupled to a va-
with high group transfer potential. riety of mechanical devices to accomplish useful work.
Living cells have an analogous biological “circuit,”
with a relatively reduced compound such as glucose as
the source of electrons. As glucose is enzymatically ox-
13.3 Biological Oxidation-Reduction idized, the released electrons flow spontaneously
Reactions through a series of electron-carrier intermediates to an-
other chemical species, such as O2. This electron flow
The transfer of phosphoryl groups is a central feature is exergonic, because O2 has a higher affinity for elec-
of metabolism. Equally important is another kind of trons than do the electron-carrier intermediates. The
transfer, electron transfer in oxidation-reduction reac- resulting electromotive force provides energy to a vari-
tions. These reactions involve the loss of electrons by ety of molecular energy transducers (enzymes and other
one chemical species, which is thereby oxidized, and the proteins) that do biological work. In the mitochondrion,
gain of electrons by another, which is reduced. The flow for example, membrane-bound enzymes couple electron
of electrons in oxidation-reduction reactions is respon- flow to the production of a transmembrane pH differ-
sible, directly or indirectly, for all work done by living ence, accomplishing osmotic and electrical work. The
organisms. In nonphotosynthetic organisms, the sources proton gradient thus formed has potential energy, some-
of electrons are reduced compounds (foods); in photo- times called the proton-motive force by analogy with
synthetic organisms, the initial electron donor is a chem- electromotive force. Another enzyme, ATP synthase in
ical species excited by the absorption of light. The path the inner mitochondrial membrane, uses the proton-
of electron flow in metabolism is complex. Electrons motive force to do chemical work: synthesis of ATP from
move from various metabolic intermediates to special- ADP and Pi as protons flow spontaneously across the
ized electron carriers in enzyme-catalyzed reactions. membrane. Similarly, membrane-localized enzymes in
508 Chapter 13 Principles of Bioenergetics

E. coli convert electromotive force to proton-motive Biological Oxidations Often Involve Dehydrogenation
force, which is then used to power flagellar motion.
The carbon in living cells exists in a range of oxidation
The principles of electrochemistry that govern en-
states (Fig. 13–13). When a carbon atom shares an elec-
ergy changes in the macroscopic circuit with a motor
tron pair with another atom (typically H, C, S, N, or O),
and battery apply with equal validity to the molecular
the sharing is unequal in favor of the more electroneg-
processes accompanying electron flow in living cells. We
ative atom. The order of increasing electronegativity is
turn now to a discussion of those principles.
H  C  S  N  O. In oversimplified but useful terms,
the more electronegative atom “owns” the bonding elec-
Oxidation-Reductions Can Be Described
trons it shares with another atom. For example, in
as Half-Reactions methane (CH4), carbon is more electronegative than the
Although oxidation and reduction must occur together, four hydrogens bonded to it, and the C atom therefore
it is convenient when describing electron transfers to “owns” all eight bonding electrons (Fig. 13–13). In
consider the two halves of an oxidation-reduction reac- ethane, the electrons in the COC bond are shared
tion separately. For example, the oxidation of ferrous equally, so each C atom owns only seven of its eight
ion by cupric ion, bonding electrons. In ethanol, C-1 is less electronega-
tive than the oxygen to which it is bonded, and the O
Fe2  Cu2 34 Fe3  Cu
atom therefore “owns” both electrons of the COO bond,
can be described in terms of two half-reactions: leaving C-1 with only five bonding electrons. With each
formal loss of electrons, the carbon atom has undergone
(1) Fe2 34 Fe3  e
oxidation—even when no oxygen is involved, as in the
(2) Cu2  e 34 Cu
conversion of an alkane (OCH2OCH2O) to an alkene
The electron-donating molecule in an oxidation- (OCHUCHO). In this case, oxidation (loss of elec-
reduction reaction is called the reducing agent or reduc- trons) is coincident with the loss of hydrogen. In bio-
tant; the electron-accepting molecule is the oxidizing logical systems, oxidation is often synonymous with de-
agent or oxidant. A given agent, such as an iron cation hydrogenation, and many enzymes that catalyze
existing in the ferrous (Fe2) or ferric (Fe3) state, func- oxidation reactions are dehydrogenases. Notice that
tions as a conjugate reductant-oxidant pair (redox pair), the more reduced compounds in Figure 13–13 (top) are
just as an acid and corresponding base function as a con- richer in hydrogen than in oxygen, whereas the more
jugate acid-base pair. Recall from Chapter 2 that in acid- oxidized compounds (bottom) have more oxygen and
base reactions we can write a general equation: proton less hydrogen.
donor 3 4 H  proton acceptor. In redox reactions we Not all biological oxidation-reduction reactions in-
can write a similar general equation: electron donor 3 4 volve carbon. For example, in the conversion of molec-
e  electron acceptor. In the reversible half-reaction (1) ular nitrogen to ammonia, 6H  6e  N2 n 2NH3,
above, Fe2 is the electron donor and Fe3 is the elec- the nitrogen atoms are reduced.
tron acceptor; together, Fe2 and Fe3 constitute a con- Electrons are transferred from one molecule (elec-
jugate redox pair. tron donor) to another (electron acceptor) in one of
The electron transfers in the oxidation-reduction four different ways:
reactions of organic compounds are not fundamentally
different from those of inorganic species. In Chapter 7 1. Directly as electrons. For example, the Fe2/Fe3
we considered the oxidation of a reducing sugar (an redox pair can transfer an electron to the
aldehyde or ketone) by cupric ion (see Fig. 7–10a): Cu/Cu2 redox pair:
Fe2  Cu2 34 Fe3  Cu
O O
R C  4OH  2Cu2 R C  Cu2O  2H2O 2. As hydrogen atoms. Recall that a hydrogen atom
H OH consists of a proton (H) and a single electron (e).
In this case we can write the general equation
This overall reaction can be expressed as two half-
reactions: AH2 34 A  2e  2H
O O where AH2 is the hydrogen/electron donor.
(1) R C  2OH R C  2e  H2O (Do not mistake the above reaction for an acid
H OH dissociation; the H arises from the removal of a
(2) 2Cu 2 
 2e  2OH 
34 Cu2O  H2O hydrogen atom, H  e.) AH2 and A together
constitute a conjugate redox pair (A/AH2), which
Because two electrons are removed from the aldehyde
can reduce another compound B (or redox pair,
carbon, the second half-reaction (the one-electron re-
B/BH2) by transfer of hydrogen atoms:
duction of cupric to cuprous ion) must be doubled to
balance the overall equation. AH2  B 34 A  BH2
13.3 Biological Oxidation-Reduction Reactions 509

3. As a hydride ion (:H), which has two electrons. potential of 0.00 V. When this hydrogen electrode is con-
This occurs in the case of NAD-linked dehydroge- nected through an external circuit to another half-cell
nases, described below. in which an oxidized species and its corresponding re-
4. Through direct combination with oxygen. In this duced species are present at standard concentrations
case, oxygen combines with an organic reductant (each solute at 1 M, each gas at 101.3 kPa), electrons tend
and is covalently incorporated in the product, as to flow through the external circuit from the half-cell of
in the oxidation of a hydrocarbon to an alcohol:
1
RXCH3  2O2 88n RXCH2XOH

The hydrocarbon is the electron donor and the H


oxygen atom is the electron acceptor.
Methane H C H 8
H
All four types of electron transfer occur in cells. The
neutral term reducing equivalent is commonly used to
designate a single electron equivalent participating in an H H
oxidation-reduction reaction, no matter whether this Ethane H C C H 7
(alkane)
equivalent is an electron per se, a hydrogen atom, or a hy- H H
dride ion, or whether the electron transfer takes place in
a reaction with oxygen to yield an oxygenated product. H H
Because biological fuel molecules are usually enzymati- Ethene C C 6
(alkene) H H
cally dehydrogenated to lose two reducing equivalents at
a time, and because each oxygen atom can accept two re-
ducing equivalents, biochemists by convention regard the H H
unit of biological oxidations as two reducing equivalents Ethanol H C C O H 5
passing from substrate to oxygen. (alcohol)
H H

Reduction Potentials Measure Affinity for Electrons Acetylene


(alkyne) H C C H 5
When two conjugate redox pairs are together in solu-
tion, electron transfer from the electron donor of one
pair to the electron acceptor of the other may proceed H
Formaldehyde C O 4
spontaneously. The tendency for such a reaction de- H
pends on the relative affinity of the electron acceptor
of each redox pair for electrons. The standard reduc-
H H
tion potential, E, a measure (in volts) of this affin- Acetaldehyde
ity, can be determined in an experiment such as that H C C 3
(aldehyde)
described in Figure 13–14. Electrochemists have cho- H O
sen as a standard of reference the half-reaction
H O H
1
H  e 88n 2 H2
Acetone H C C C H 2
The electrode at which this half-reaction occurs (called (ketone)
a half-cell) is arbitrarily assigned a standard reduction H H

O
Formic acid
(carboxylic H C 2
FIGURE 13–13 Oxidation states of carbon in the biosphere. The O
acid)
oxidation states are illustrated with some representative compounds. H
Focus on the red carbon atom and its bonding electrons. When this
carbon is bonded to the less electronegative H atom, both bonding Carbon C O 2
electrons (red) are assigned to the carbon. When carbon is bonded to monoxide
another carbon, bonding electrons are shared equally, so one of the
two electrons is assigned to the red carbon. When the red carbon is H
Acetic acid O
bonded to the more electronegative O atom, the bonding electrons
(carboxylic H C C 1
are assigned to the oxygen. The number to the right of each compound acid) O
H H
is the number of electrons “owned” by the red carbon, a rough ex-
pression of the oxidation state of that carbon. When the red carbon
undergoes oxidation (loses electrons), the number gets smaller. Thus Carbon O C O 0
the oxidation state increases from top to bottom of the list. dioxide
510 Chapter 13 Principles of Bioenergetics

Device for where R and T have their usual meanings, n is the num-
measuring emf ber of electrons transferred per molecule, and is the
Faraday constant (Table 13–1). At 298 K (25 C), this
expression reduces to
0.026 V [electron acceptor]
E  E   ln  (13–5)
n [electron donor]
H2 gas
Many half-reactions of interest to biochemists in-
(standard
pressure)
Salt bridge volve protons. As in the definition of G
, biochemists
(KCl solution)
define the standard state for oxidation-reduction reac-
tions as pH 7 and express reduction potential as E
, the
standard reduction potential at pH 7. The standard re-
duction potentials given in Table 13–7 and used through-
out this book are values for E
and are therefore valid
only for systems at neutral pH. Each value represents
the potential difference when the conjugate redox pair,
at 1 M concentrations and pH 7, is connected with the
standard (pH 0) hydrogen electrode. Notice in Table
13–7 that when the conjugate pair 2H/H2 at pH 7 is
connected with the standard hydrogen electrode (pH
Reference cell of Test cell containing
known emf: the 1 M concentrations
0), electrons tend to flow from the pH 7 cell to the stan-
hydrogen electrode of the oxidized and dard (pH 0) cell; the measured E
for the 2H/H2 pair
in which H2 gas reduced species of is 0.414 V.
at 101.3 kPa is the redox pair to
equilibrated at be examined
the electrode Standard Reduction Potentials Can Be Used
with 1 M H to Calculate the Free-Energy Change
The usefulness of reduction potentials stems from the
FIGURE 13–14 Measurement of the standard reduction potential fact that when E values have been determined for any
(E
) of a redox pair. Electrons flow from the test electrode to the ref- two half-cells, relative to the standard hydrogen elec-
erence electrode, or vice versa. The ultimate reference half-cell is the trode, their reduction potentials relative to each other
hydrogen electrode, as shown here, at pH 0. The electromotive force
are also known. We can then predict the direction in
(emf) of this electrode is designated 0.00 V. At pH 7 in the test cell,
which electrons will tend to flow when the two half-cells
E
for the hydrogen electrode is 0.414 V. The direction of electron
are connected through an external circuit or when com-
flow depends on the relative electron “pressure” or potential of the
ponents of both half-cells are present in the same solu-
two cells. A salt bridge containing a saturated KCl solution provides
tion. Electrons tend to flow to the half-cell with the more
a path for counter-ion movement between the test cell and the refer-
ence cell. From the observed emf and the known emf of the reference
positive E, and the strength of that tendency is pro-
cell, the experimenter can find the emf of the test cell containing the
portional to the difference in reduction potentials, E.
redox pair. The cell that gains electrons has, by convention, the more The energy made available by this spontaneous
positive reduction potential. electron flow (the free-energy change for the oxidation-
reduction reaction) is proportional to  E:
G  n E or G
 n  E
(13–6)

lower standard reduction potential to the half-cell of Here n represents the number of electrons transferred
higher standard reduction potential. By convention, the in the reaction. With this equation we can calculate the
half-cell with the stronger tendency to acquire electrons free-energy change for any oxidation-reduction reaction
is assigned a positive value of E . from the values of E
in a table of reduction potentials
The reduction potential of a half-cell depends not (Table 13–7) and the concentrations of the species par-
only on the chemical species present but also on their ticipating in the reaction.
activities, approximated by their concentrations. About Consider the reaction in which acetaldehyde is
a century ago, Walther Nernst derived an equation that reduced by the biological electron carrier NADH:
relates standard reduction potential (E ) to the reduc-
Acetaldehyde  NADH  H 88n ethanol  NAD
tion potential (E) at any concentration of oxidized and
reduced species in the cell: The relevant half-reactions and their E
values are:
RT [electron acceptor] (1) Acetaldehyde  2H  2e 88n ethanol
E  E   ln  (13–4)
nℑ [electron donor] E
 0.197 V
13.3 Biological Oxidation-Reduction Reactions 511

(2) NAD  2H  2e 88n NADH  H the values of E for both reductants are determined
E
 0.320 V (Eqn 13–4):
By convention, E
is expressed as E
of the electron RT [acetaldehyde]
Eacetaldehyde  E   ln 
acceptor minus E
of the electron donor. Because ac- nℑ [ethanol]
etaldehyde is accepting electrons from NADH in our 0.026 V 1.00
 0.197 V   ln   0.167 V
example, E
 0.197 V  (0.320 V)  0.123 V, 2 0.100
and n is 2. Therefore, RT [NAD]
ENADH  E   ln 
nℑ [NADH]
G
 n E
 2(96.5 kJ/V  mol)(0.123 V) 0.026 V 1.00
 23.7 kJ/mol  0.320 V   ln   0.350 V
2 0.100
This is the free-energy change for the oxidation- Then E is used to calculate G (Eqn 13–5):
reduction reaction at pH 7, when acetaldehyde, ethanol,
NAD, and NADH are all present at 1.00 M concentra- E  0.167 V  (0.350) V  0.183 V
tions. If, instead, acetaldehyde and NADH were present G  n E
at 1.00 M but ethanol and NAD were present at 0.100 M,  2(96.5 kJ/V  mol)(0.183 V)
the value for G would be calculated as follows. First,  35.3 kJ/mol

TABLE 13–7 Standard Reduction Potentials of Some Biologically


Important Half-Reactions, at pH 7.0 and 25 C (298 K)
Half-reaction E
(V)
1
O2  2H  2e 88n H2O

2 0.816
Fe3  e 88n Fe2 0.771
NO3  2H  2e 88n NO 2  H2O 0.421
Cytochrome f (Fe3)  e 88n cytochrome f (Fe2) 0.365
Fe(CN)63 (ferricyanide)  e 88n Fe(CN)64 0.36
Cytochrome a3 (Fe3)  e 88n cytochrome a3 (Fe2) 0.35
O2  2H  2e 88n H2O2 0.295
Cytochrome a (Fe3)  e 88n cytochrome a (Fe2) 0.29
Cytochrome c (Fe3)  e 88n cytochrome c (Fe2) 0.254
Cytochrome c1 (Fe3)  e 88n cytochrome c1 (Fe2) 0.22
Cytochrome b (Fe3)  e 88n cytochrome b (Fe2) 0.077
Ubiquinone  2H  2e 88n ubiquinol  H2 0.045
Fumarate2  2H  2e 88n succinate2 0.031
2H  2e 88n H2 (at standard conditions, pH 0) 0.000
Crotonyl-CoA  2H  2e 88n butyryl-CoA 0.015
Oxaloacetate2  2H  2e 88n malate2 0.166
Pyruvate  2H  2e 88n lactate 0.185
Acetaldehyde  2H  2e 88n ethanol 0.197
FAD  2H  2e 88n FADH2 0.219*
Glutathione  2H  2e 88n 2 reduced glutathione 0.23
S  2H  2e 88n H2S 0.243
Lipoic acid  2H  2e 88n dihydrolipoic acid 0.29
NAD  H  2e 88n NADH 0.320
NADP  H  2e 88n NADPH 0.324
Acetoacetate  2H  2e 88n -hydroxybutyrate 0.346
-Ketoglutarate  CO2  2H  2e 88n isocitrate 0.38
2H  2e 88n H2 (at pH 7) 0.414
Ferredoxin (Fe3)  e 88n ferredoxin (Fe2) 0.432

Source: Data mostly from Loach, P.A. (1976) In Handbook of Biochemistry and Molecular Biology, 3rd edn (Fasman, G.D., ed.),
Physical and Chemical Data, Vol. I, pp. 122–130, CRC Press, Boca Raton, FL.
* This is the value for free FAD; FAD bound to a specific flavoprotein (for example succinate dehydrogenase) has a different E

that depends on its protein environments.
512 Chapter 13 Principles of Bioenergetics

It is thus possible to calculate the free-energy change NADH and NADPH Act with Dehydrogenases
for any biological redox reaction at any concentrations as Soluble Electron Carriers
of the redox pairs.
Nicotinamide adenine dinucleotide (NAD in its oxi-
dized form) and its close analog nicotinamide adenine
Cellular Oxidation of Glucose to Carbon Dioxide
dinucleotide phosphate (NADP) are composed of two
Requires Specialized Electron Carriers nucleotides joined through their phosphate groups by a
The principles of oxidation-reduction energetics de- phosphoanhydride bond (Fig. 13–15a). Because the
scribed above apply to the many metabolic reactions that nicotinamide ring resembles pyridine, these compounds
involve electron transfers. For example, in many organ- are sometimes called pyridine nucleotides. The vita-
isms, the oxidation of glucose supplies energy for the min niacin is the source of the nicotinamide moiety in
production of ATP. The complete oxidation of glucose: nicotinamide nucleotides.
Both coenzymes undergo reversible reduction of
C6H12O6  6O2 8n 6CO2  6H2O
the nicotinamide ring (Fig. 13–15). As a substrate mol-
has a G
of 2,840 kJ/mol. This is a much larger re- ecule undergoes oxidation (dehydrogenation), giving up
lease of free energy than is required for ATP synthesis two hydrogen atoms, the oxidized form of the nucleotide
(50 to 60 kJ/mol; see Box 13–1). Cells convert glucose (NAD or NADP) accepts a hydride ion (:H, the
to CO2 not in a single, high-energy-releasing reaction, equivalent of a proton and two electrons) and is trans-
but rather in a series of controlled reactions, some of formed into the reduced form (NADH or NADPH). The
which are oxidations. The free energy released in these second proton removed from the substrate is released
oxidation steps is of the same order of magnitude as that to the aqueous solvent. The half-reaction for each type
required for ATP synthesis from ADP, with some energy of nucleotide is therefore
to spare. Electrons removed in these oxidation steps are NAD  2e  2H 8n NADH  H
transferred to coenzymes specialized for carrying elec- NADP  2e  2H 8n NADPH  H
trons, such as NAD and FAD (described below).
Reduction of NAD or NADP converts the benzenoid
A Few Types of Coenzymes and Proteins Serve ring of the nicotinamide moiety (with a fixed positive
charge on the ring nitrogen) to the quinonoid form (with
as Universal Electron Carriers
no charge on the nitrogen). Note that the reduced nu-
The multitude of enzymes that catalyze cellular oxida- cleotides absorb light at 340 nm; the oxidized forms do
tions channel electrons from their hundreds of different not (Fig. 13–15b). The plus sign in the abbreviations
substrates into just a few types of universal electron car- NAD and NADP does not indicate the net charge on
riers. The reduction of these carriers in catabolic these molecules (they are both negatively charged);
processes results in the conservation of free energy re- rather, it indicates that the nicotinamide ring is in its
leased by substrate oxidation. NAD, NADP, FMN, and oxidized form, with a positive charge on the nitrogen
FAD are water-soluble coenzymes that undergo re- atom. In the abbreviations NADH and NADPH, the “H”
versible oxidation and reduction in many of the electron- denotes the added hydride ion. To refer to these nu-
transfer reactions of metabolism. The nucleotides NAD cleotides without specifying their oxidation state, we
and NADP move readily from one enzyme to another; use NAD and NADP.
the flavin nucleotides FMN and FAD are usually very The total concentration of NAD  NADH in most
tightly bound to the enzymes, called flavoproteins, for tissues is about 105 M; that of NADP  NADPH is
which they serve as prosthetic groups. Lipid-soluble about 106 M. In many cells and tissues, the ratio of
quinones such as ubiquinone and plastoquinone act as NAD (oxidized) to NADH (reduced) is high, favoring
electron carriers and proton donors in the nonaqueous hydride transfer from a substrate to NAD to form
environment of membranes. Iron-sulfur proteins and cy- NADH. By contrast, NADPH (reduced) is generally pres-
tochromes, which have tightly bound prosthetic groups ent in greater amounts than its oxidized form, NADP,
that undergo reversible oxidation and reduction, also favoring hydride transfer from NADPH to a substrate.
serve as electron carriers in many oxidation-reduction This reflects the specialized metabolic roles of the two
reactions. Some of these proteins are water-soluble, but coenzymes: NAD generally functions in oxidations—
others are peripheral or integral membrane proteins (see usually as part of a catabolic reaction; and NADPH is
Fig. 11–6). the usual coenzyme in reductions—nearly always as
We conclude this chapter by describing some chem- part of an anabolic reaction. A few enzymes can use ei-
ical features of nucleotide coenzymes and some of the ther coenzyme, but most show a strong preference for
enzymes (dehydrogenases and flavoproteins) that use one over the other. The processes in which these two
them. The oxidation-reduction chemistry of quinones, cofactors function are also segregated in specific or-
iron-sulfur proteins, and cytochromes is discussed in ganelles of eukaryotic cells: oxidations of fuels such as
Chapter 19. pyruvate, fatty acids, and -keto acids derived from
13.3 Biological Oxidation-Reduction Reactions 513

O O O
H B H H B H H B
C 2e
? C ? CH
H H
NH2 NH2 or NH2  H

N 2H N N
O CH2 O A A
R A side R B side
H H
OP PO O H H NADH
(reduced)

O OH OH
NH2
N
OP PO O N Adenine
1.0
Oxidized
N N (NAD)
O CH2 O 0.8
Absorbance

H H 0.6
H H

NAD Reduced
(oxidized) 0.4
OH OH (NADH)

In NADP this hydroxyl group 0.2


is esterified with phosphate.
0.0
(a) 220 240 260 280 300 320 340 360 380
Wavelength (nm)
(b)

FIGURE 13–15 NAD and NADP. (a) Nicotinamide adenine dinu- tra of NAD and NADH. Reduction of the nicotinamide ring produces
cleotide, NAD, and its phosphorylated analog NADP undergo re- a new, broad absorption band with a maximum at 340 nm. The pro-
duction to NADH and NADPH, accepting a hydride ion (two elec- duction of NADH during an enzyme-catalyzed reaction can be con-
trons and one proton) from an oxidizable substrate. The hydride ion veniently followed by observing the appearance of the absorbance at
is added to either the front (the A side) or the back (the B side) of the 340 nm (the molar extinction coefficient 340  6,200 M1cm1).
planar nicotinamide ring (see Table 13–8). (b) The UV absorption spec-

amino acids occur in the mitochondrial matrix, whereas When NAD or NADP is reduced, the hydride ion
reductive biosynthesis processes such as fatty acid syn- could in principle be transferred to either side of the
thesis take place in the cytosol. This functional and spa- nicotinamide ring: the front (A side) or the back (B
tial specialization allows a cell to maintain two distinct side), as represented in Figure 13–15a. Studies with iso-
pools of electron carriers, with two distinct functions. topically labeled substrates have shown that a given en-
More than 200 enzymes are known to catalyze re- zyme catalyzes either an A-type or a B-type transfer, but
actions in which NAD (or NADP) accepts a hydride not both. For example, yeast alcohol dehydrogenase and
ion from a reduced substrate, or NADPH (or NADH) do- lactate dehydrogenase of vertebrate heart transfer a hy-
nates a hydride ion to an oxidized substrate. The gen- dride ion to (or remove a hydride ion from) the A side
eral reactions are of the nicotinamide ring; they are classed as type A de-
hydrogenases to distinguish them from another group
AH2  NAD 8n A  NADH  H
of enzymes that transfer a hydride ion to (or remove a
A  NADPH  H 8n AH2  NADP
hydride ion from) the B side of the nicotinamide ring
where AH2 is the reduced substrate and A the oxidized (Table 13–8). The specificity for one side or another can
substrate. The general name for an enzyme of this type be very striking; lactate dehydrogenase, for example,
is oxidoreductase; they are also commonly called de- prefers the A side over the B side by a factor of 5  107!
hydrogenases. For example, alcohol dehydrogenase Most dehydrogenases that use NAD or NADP bind
catalyzes the first step in the catabolism of ethanol, in the cofactor in a conserved protein domain called the
which ethanol is oxidized to acetaldehyde: Rossmann fold (named for Michael Rossmann, who de-
duced the structure of lactate dehydrogenase and first
CH3CH2OH  NAD 8n CH3CHO  NADH  H
described this structural motif). The Rossmann fold typ-
Ethanol Acetaldehyde
ically consists of a six-stranded parallel  sheet and four
Notice that one of the carbon atoms in ethanol has lost associated helices (Fig. 13–16).
a hydrogen; the compound has been oxidized from an The association between a dehydrogenase and NAD
alcohol to an aldehyde (refer again to Fig. 13–13 for the or NADP is relatively loose; the coenzyme readily diffuses
oxidation states of carbon). from one enzyme to another, acting as a water-soluble
514 Chapter 13 Principles of Bioenergetics

TABLE 13–8 Stereospecificity of Dehydrogenases That Employ NAD or NADP as Coenzymes


Stereochemical
specificity for
nicotinamide
Enzyme Coenzyme ring (A or B) Text page(s)
Isocitrate dehydrogenase NAD A XXX–XXX
-Ketoglutarate dehydrogenase NAD B XXX
Glucose 6-phosphate dehydrogenase NADP B XXX
Malate dehydrogenase NAD A XXX
Glutamate dehydrogenase NAD or NADP B XXX
Glyceraldehyde 3-phosphate dehydrogenase NAD B XXX
Lactate dehydrogenase NAD A XXX
Alcohol dehydrogenase NAD A XXX

carrier of electrons from one metabolite to another. For (1) Glyceraldehyde 3-phosphate  NAD 8n
example, in the production of alcohol during fermenta- 3-phosphoglycerate  NADH  H
tion of glucose by yeast cells, a hydride ion is removed (2) Acetaldehyde  NADH  H 8n ethanol  NAD
from glyceraldehyde 3-phosphate by one enzyme (glyc-
Sum: Glyceraldehyde 3-phosphate  acetaldehyde 8n
eraldehyde 3-phosphate dehydrogenase, a type B en-
3-phosphoglycerate  ethanol
zyme) and transferred to NAD. The NADH produced
then leaves the enzyme surface and diffuses to another Notice that in the overall reaction there is no net pro-
enzyme (alcohol dehydrogenase, a type A enzyme), duction or consumption of NAD or NADH; the coen-
which transfers a hydride ion to acetaldehyde, produc- zymes function catalytically and are recycled repeatedly
ing ethanol: without a net change in the concentration of NAD 
NADH.

Dietary Deficiency of Niacin, the Vitamin Form


of NAD and NADP, Causes Pellagra
The pyridine-like rings of NAD and NADP are de-
rived from the vitamin niacin (nicotinic acid; Fig.
13–17), which is synthesized from tryptophan. Humans
generally cannot synthesize niacin in sufficient quanti-
ties, and this is especially so for those with diets low in
tryptophan (maize, for example, has a low tryptophan
content). Niacin deficiency, which affects all the
NAD(P)-dependent dehydrogenases, causes the serious
human disease pellagra (Italian for “rough skin”) and a
related disease in dogs, blacktongue. These diseases are
characterized by the “three Ds”: dermatitis, diarrhea, and
dementia, followed in many cases by death. A century
ago, pellagra was a common human disease; in the south-
ern United States, where maize was a dietary staple,
about 100,000 people were afflicted and about 10,000
died between 1912 and 1916. In 1920 Joseph Goldberger
showed pellagra to be caused by a dietary insufficiency,
FIGURE 13–16 The nucleotide binding domain of the enzyme lac- and in 1937 Frank Strong, D. Wayne Wolley, and Conrad
tate dehydrogenase. (a) The Rossmann fold is a structural motif found Elvehjem identified niacin as the curative agent for
in the NAD-binding site of many dehydrogenases. It consists of a blacktongue. Supplementation of the human diet with
six-stranded parallel  sheet and four  helices; inspection reveals this inexpensive compound led to the eradication of pel-
the arrangement to be a pair of structurally similar ---- motifs. lagra in the populations of the developed world—with
(b) The dinucleotide NAD binds in an extended conformation through one significant exception. Pellagra is still found among
hydrogen bonds and salt bridges (derived from PDB ID 3LDH). alcoholics, whose intestinal absorption of niacin is much
13.3 Biological Oxidation-Reduction Reactions 515

O
TABLE 13–9 Some Enzymes (Flavoproteins)
C O C That Employ Flavin Nucleotide Coenzymes


  CH3 Flavin Text


Niacin Nicotine Enzyme nucleotide page(s)
(nicotinic acid)
Acyl–CoA dehydrogenase FAD XXX

O NH3 Dihydrolipoyl dehydrogenase FAD XXX
Succinate dehydrogenase FAD XXX
C CH2 CH COO
NH2 Glycerol 3-phosphate dehydrogenase FAD XXX
Thioredoxin reductase FAD XXX–XXX
  NADH dehydrogenase (Complex I) FMN XXX

Nicotinamide Tryptophan
Glycolate oxidase FMN XXX

FIGURE 13–17 Structures of niacin (nicotinic acid) and its deriva-


tive nicotinamide. The biosynthetic precursor of these compounds is
loxazine ring is produced, abbreviated FADH• and
tryptophan. In the laboratory, nicotinic acid was first produced by ox-
FMNH•. Because flavoproteins can participate in either
idation of the natural product nicotine—thus the name. Both nicotinic
acid and nicotinamide cure pellagra, but nicotine (from cigarettes or
one- or two-electron transfers, this class of proteins
elsewhere) has no curative activity. is involved in a greater diversity of reactions than the
NAD (P)-linked dehydrogenases.
Like the nicotinamide coenzymes (Fig. 13–15), the
reduced, and whose caloric needs are often met with dis-
flavin nucleotides undergo a shift in a major absorption
tilled spirits that are virtually devoid of vitamins, in-
band on reduction. Flavoproteins that are fully reduced
cluding niacin. In a few places, including the Deccan
(two electrons accepted) generally have an absorption
Plateau in India, pellagra still occurs, especially among
maximum near 360 nm. When partially reduced (one
the poor. ■
electron), they acquire another absorption maximum at
about 450 nm; when fully oxidized, the flavin has max-
Flavin Nucleotides Are Tightly Bound in Flavoproteins
ima at 370 and 440 nm. The intermediate radical form,
Flavoproteins (Table 13–9) are enzymes that catalyze reduced by one electron, has absorption maxima at 380,
oxidation-reduction reactions using either flavin 480, 580, and 625 nm. These changes can be used to as-
mononucleotide (FMN) or flavin adenine dinucleotide say reactions involving a flavoprotein.
(FAD) as coenzyme (Fig. 13–18). These coenzymes, the The flavin nucleotide in most flavoproteins is bound
flavin nucleotides, are derived from the vitamin ri- rather tightly to the protein, and in some enzymes, such
boflavin. The fused ring structure of flavin nucleotides as succinate dehydrogenase, it is bound covalently. Such
(the isoalloxazine ring) undergoes reversible reduction, tightly bound coenzymes are properly called prosthetic
accepting either one or two electrons in the form of one groups. They do not transfer electrons by diffusing from
or two hydrogen atoms (each atom an electron plus a one enzyme to another; rather, they provide a means by
proton) from a reduced substrate. The fully reduced which the flavoprotein can temporarily hold electrons
forms are abbreviated FADH2 and FMNH2. When a fully while it catalyzes electron transfer from a reduced sub-
oxidized flavin nucleotide accepts only one electron strate to an electron acceptor. One important feature of
(one hydrogen atom), the semiquinone form of the isoal- the flavoproteins is the variability in the standard re-
duction potential (E
) of the
bound flavin nucleotide. Tight as-
sociation between the enzyme and
prosthetic group confers on the
flavin ring a reduction potential
typical of that particular flavopro-
tein, sometimes quite different
from the reduction potential of the
free flavin nucleotide. FAD bound
to succinate dehydrogenase, for
example, has an E
close to 0.0 V,
compared with 0.219 V for free
Frank Strong, D. Wayne Woolley, Conrad Elvehjem, FAD; E
for other flavoproteins
1908–1993 1914–1966 1901–1962 ranges from 0.40 V to 0.06 V.
516 Chapter 13 Principles of Bioenergetics

isoalloxazine ring
O H O H O

CH3 N H e  CH3 N 
H e 
CH3 N
NH • NH NH

CH3 N N O CH3 N N O CH3 N N O


CH2 R R H
HCOH FADH• (FMNH•) FADH2 (FMNH2)
FMN (semiquinone) (fully reduced)
HCOH
HCOH
CH2
FAD
O

O P O
O
NH2

O P O N
N
O
N N
CH2 O
FIGURE 13–18 Structures of oxidized and reduced FAD and FMN.
H H FMN consists of the structure above the dashed line on the FAD (ox-
H H
idized form). The flavin nucleotides accept two hydrogen atoms (two
OH OH electrons and two protons), both of which appear in the flavin ring
Flavin adenine dinucleotide (FAD) and system. When FAD or FMN accepts only one hydrogen atom, the semi-
flavin mononucleotide (FMN) quinone, a stable free radical, forms.

Flavoproteins are often very complex; some have, in ad- ■ Biological oxidation-reduction reactions can be
dition to a flavin nucleotide, tightly bound inorganic ions described in terms of two half-reactions, each
(iron or molybdenum, for example) capable of partici- with a characteristic standard reduction
pating in electron transfers. potential, E
.
Certain flavoproteins act in a quite different role as ■ When two electrochemical half-cells, each
light receptors. Cryptochromes are a family of flavo- containing the components of a half-reaction,
proteins, widely distributed in the eukaryotic phyla, that are connected, electrons tend to flow to the
mediate the effects of blue light on plant development half-cell with the higher reduction potential.
and the effects of light on mammalian circadian rhythms The strength of this tendency is proportional
(oscillations in physiology and biochemistry, with a to the difference between the two reduction
24-hour period). The cryptochromes are homologs of potentials (E) and is a function of the
another family of flavoproteins, the photolyases. Found concentrations of oxidized and reduced
in both prokaryotes and eukaryotes, photolyases use species.
the energy of absorbed light to repair chemical defects
in DNA. ■ The standard free-energy change for an
We examine the function of flavoproteins as elec- oxidation-reduction reaction is directly
tron carriers in Chapter 19, when we consider their roles proportional to the difference in standard
in oxidative phosphorylation (in mitochondria) and pho- reduction potentials of the two half-cells:
tophosphorylation (in chloroplasts), and we describe G
 n E
.
the photolyase reactions in Chapter 25. ■ Many biological oxidation reactions are
dehydrogenations in which one or two
SUMMARY 13.3 Biological Oxidation-Reduction hydrogen atoms (H  e) are transferred
Reactions from a substrate to a hydrogen acceptor.
Oxidation-reduction reactions in living cells
■ In many organisms, a central energy-conserving involve specialized electron carriers.
process is the stepwise oxidation of glucose to ■ NAD and NADP are the freely diffusible
CO2, in which some of the energy of oxidation is coenzymes of many dehydrogenases. Both
conserved in ATP as electrons are passed to O2. NAD and NADP accept two electrons and
Chapter 13 Further Reading 517

one proton. NAD and NADP are bound to flavoproteins. They can accept either one or
dehydrogenases in a widely conserved two electrons. Flavoproteins also serve as light
structural motif called the Rossmann fold. receptors in cryptochromes and photolyases.
■ FAD and FMN, the flavin nucleotides, serve as
tightly bound prosthetic groups of

Key Terms
Terms in bold are defined in the glossary.
autotroph XXX thioester XXX dehydrogenation XXX
heterotroph XXX adenylylation XXX dehydrogenases XXX
metabolism XXX inorganic pyrophosphatase XXX reducing equivalent XXX
metabolic pathways XXX nucleoside diphosphate standard reduction potential
metabolite XXX kinase XXX (E ) XXX
intermediary metabolism XXX adenylate kinase XXX pyridine nucleotide XXX
catabolism XXX creatine kinase XXX oxidoreductase XXX
anabolism XXX phosphagens XXX flavoprotein XXX
standard transformed constants XXX polyphosphate kinase-1, -2 XXX flavin nucleotides XXX
phosphorylation potential electromotive force (emf) XXX cryptochrome XXX
(Gp) XXX conjugate redox pair XXX photolyase XXX

Further Reading
Bioenergetics and Thermodynamics Morowitz, H.J. (1978) Foundations of Bioenergetics, Academic
Atkins, P.W. (1984) The Second Law, Scientific American Books, Press, Inc., New York. [Out of print.]
Inc., New York. Clear, rigorous description of thermodynamics in biology.
A well-illustrated and elementary discussion of the second law Nicholls, D.G. & Ferguson, S.J. (2002) Bioenergetics 3,
and its implications. Academic Press, Inc., New York.
Becker, W.M. (1977) Energy and the Living Cell: An Introduc- Clear, well-illustrated intermediate-level discussion of the
tion to Bioenergetics, J. B. Lippincott Company, Philadelphia. theory of bioenergetics and the mechanisms of energy
A clear introductory account of cellular metabolism, in terms of transductions.
energetics. Tinoco, I., Jr., Sauer, K., & Wang, J.C. (1996) Physical Chem-
Bergethon, P.R. (1998) The Physical Basis of Biochemistry, istry: Principles and Applications in Biological Sciences, 3rd
Springer Verlag, New York. edn, Prentice-Hall, Inc., Upper Saddle River, NJ.
Chapters 11 through 13 of this book, and the books by Tinoco Chapters 2 through 5 cover thermodynamics.
et al. and van Holde et al. (below), are excellent general refer- van Holde, K.E., Johnson, W.C., & Ho, P.S. (1998) Principles
ences for physical biochemistry, with good discussions of the of Physical Biochemistry, Prentice-Hall, Inc., Upper Saddle River,
applications of thermodynamics to biochemistry. NJ.
Edsall, J.T. & Gutfreund, H. (1983) Biothermodynamics: The Chapters 2 and 3 are especially relevant.
Study of Biochemical Processes at Equilibrium, John Wiley &
Sons, Inc., New York.
Phosphoryl Group Transfers and ATP
Harold, F.M. (1986) The Vital Force: A Study of Bioenergetics, Alberty, R.A. (1994) Biochemical thermodynamics. Biochim.
W. H. Freeman and Company, New York. Biophys. Acta 1207, 1–11.
A beautifully clear discussion of thermodynamics in biological Explains the distinction between biochemical and chemical
processes. equations, and the calculation and meaning of transformed
Harris, D.A. (1995) Bioenergetics at a Glance, Blackwell thermodynamic properties for ATP and other phosphorylated
Science, Oxford. compounds.
A short, clearly written account of cellular energetics, including Bridger, W.A. & Henderson, J.F. (1983) Cell ATP, John Wiley &
introductory chapters on thermodynamics. Sons, Inc., New York.
Loewenstein, W.R. (1999) The Touchstone of Life: Molecular The chemistry of ATP, its role in metabolic regulation, and its
Information, Cell Communication, and the Foundations of catabolic and anabolic roles.
Life, Oxford University Press, New York. Frey, P.A. & Arabshahi, A. (1995) Standard free-energy change
Beautifully written discussion of the relationship between for the hydrolysis of the –-phosphoanhydride bridge in ATP.
entropy and information. Biochemistry 34, 11,307–11,310.
518 Chapter 13 Principles of Bioenergetics

Hanson, R.W. (1989) The role of ATP in metabolism. Biochem. Biological Oxidation-Reduction Reactions
Educ. 17, 86–92. Cashmore, A.R., Jarillo, J.A., Wu, Y.J., & Liu D. (1999)
Excellent summary of the chemistry and biology of ATP. Cryptochromes: blue light receptors for plants and animals.
Kornberg, A. (1999) Inorganic polyphosphate: a molecule of Science 284, 760–765.
many functions. Annu. Rev. Biochem. 68, 89–125. Dolphin, D., Avramovic, O., & Poulson, R. (eds) (1987)
Lipmann, F. (1941) Metabolic generation and utilization of Pyridine Nucleotide Coenzymes: Chemical, Biochemical, and
phosphate bond energy. Adv. Enzymol. 11, 99–162. Medical Aspects, John Wiley & Sons, Inc., New York.
The classic description of the role of high-energy phosphate An excellent two-volume collection of authoritative reviews.
compounds in biology. Among the most useful are the chapters by Kaplan,
Westheimer, Veech, and Ohno and Ushio.
Pullman, B. & Pullman, A. (1960) Electronic structure of
energy-rich phosphates. Radiat. Res., Suppl. 2, 160–181. Fraaije, M.W. & Mattevi, A. (2000) Flavoenzymes: diverse cata-
An advanced discussion of the chemistry of ATP and other lysts with recurrent features. Trends Biochem. Sci. 25, 126–132.
“energy-rich” compounds. Massey, V. (1994) Activation of molecular oxygen by flavins and
Veech, R.L., Lawson, J.W.R., Cornell, N.W., & Krebs, H.A. flavoproteins. J. Biol. Chem. 269, 22,459–22,462.
(1979) Cytosolic phosphorylation potential. J. Biol. Chem. 254, A short review of the chemistry of flavin–oxygen interactions in
6538–6547. flavoproteins.
Experimental determination of ATP, ADP, and Pi concentrations Rees, D.C. (2002) Great metalloclusters in enzymology. Annu.
in brain, muscle, and liver, and a discussion of the problems in Rev. Biochem. 71, 221–246.
determining the real free-energy change for ATP synthesis in Advanced review of the types of metal ion clusters found in
cells. enzymes and their modes of action.
Westheimer, F.H. (1987) Why nature chose phosphates. Science Williams, R.E. & Bruce, N.C. (2002) New uses for an old
235, 1173–1178. enzyme—the old yellow enzyme family of flavoenzymes.
A chemist’s description of the unique suitability of phosphate Microbiology 148, 1607–1614.
esters and anhydrides for metabolic transformations.

Problems
1. Entropy Changes during Egg Development Con- lowing reactions at pH 7.0 and 25 C, using the G
values
sider a system consisting of an egg in an incubator. The white in Table 13–4.
and yolk of the egg contain proteins, carbohydrates, and
glucose
lipids. If fertilized, the egg is transformed from a single cell 6-phosphatase
to a complex organism. Discuss this irreversible process in (a) Glucose 6-phosphate  H2O 3::::::::::4
terms of the entropy changes in the system, surroundings, glucose  Pi
and universe. Be sure that you first clearly define the system
and surroundings. b-galactosidase
(b) Lactose  H2O 3::::::::::4 glucose  galactose
2. Calculation of G from an Equilibrium Constant
Calculate the standard free-energy changes of the following fumarase
metabolically important enzyme-catalyzed reactions at 25 C (c) Malate 3::::::4 fumarate  H2O
and pH 7.0, using the equilibrium constants given. 4. Experimental Determination of Keq and G If a 0.1
aspartate M solution of glucose 1-phosphate is incubated with a catalytic
aminotransferase amount of phosphoglucomutase, the glucose 1-phosphate is
(a) Glutamate  oxaloacetate 3::::::::::::4 transformed to glucose 6-phosphate. At equilibrium, the con-
aspartate  -ketoglutarate K
eq  6.8 centrations of the reaction components are
triose phosphate Glucose 1-phosphate 34 glucose 6-phosphate
isomerase 4.5  103 M 9.6  102 M
(b) Dihydroxyacetone phosphate 3:::::::::::4
glyceraldehyde 3-phosphate K
eq  0.0475 Calculate K
eq and G
for this reaction at 25 C.
phosphofructokinase 5. Experimental Determination of G for ATP Hy-
(c) Fructose 6-phosphate  ATP 3:::::::::::::::4 drolysis A direct measurement of the standard free-energy
fructose 1,6-bisphosphate  ADP K
eq  254 change associated with the hydrolysis of ATP is technically
demanding because the minute amount of ATP remaining at
3. Calculation of the Equilibrium Constant from G
equilibrium is difficult to measure accurately. The value of
Calculate the equilibrium constants K
eq for each of the fol- G
can be calculated indirectly, however, from the equilib-
Chapter 13 Problems 519

rium constants of two other enzymatic reactions having less (c) The phosphorylation of glucose in the cell is coupled
favorable equilibrium constants: to the hydrolysis of ATP; that is, part of the free energy of
ATP hydrolysis is used to phosphorylate glucose:
Glucose 6-phosphate  H2O 8n glucose  Pi K
eq  270
ATP  glucose 8n ADP  glucose 6-phosphate (1) Glucose  Pi 8n glucose 6-phosphate  H2O
K
eq  890 G
 13.8 kJ/mol
(2) ATP  H2O 8n ADP  Pi
Using this information, calculate the standard free energy of G
 30.5 kJ/mol
hydrolysis of ATP at 25 C.
Sum: Glucose  ATP 8n glucose 6-phosphate  ADP
6. Difference between G and G Consider the fol-
lowing interconversion, which occurs in glycolysis (Chapter Calculate K
eq for the overall reaction. For the ATP-dependent
14): phosphorylation of glucose, what concentration of glucose is
needed to achieve a 250 M intracellular concentration of glu-
Fructose 6-phosphate 34 glucose 6-phosphate
cose 6-phosphate when the concentrations of ATP and ADP
K
eq  1.97 are 3.38 mM and 1.32 mM, respectively? Does this coupling
process provide a feasible route, at least in principle, for the
(a) What is G
for the reaction (at 25 C)?
phosphorylation of glucose in the cell? Explain.
(b) If the concentration of fructose 6-phosphate is ad-
(d) Although coupling ATP hydrolysis to glucose phos-
justed to 1.5 M and that of glucose 6-phosphate is adjusted
phorylation makes thermodynamic sense, we have not yet
to 0.50 M, what is G?
specified how this coupling is to take place. Given that cou-
(c) Why are G
and G different?
pling requires a common intermediate, one conceivable route
7. Dependence of G on pH The free energy released is to use ATP hydrolysis to raise the intracellular concentra-
by the hydrolysis of ATP under standard conditions at pH tion of Pi and thus drive the unfavorable phosphorylation of
7.0 is 30.5 kJ/mol. If ATP is hydrolyzed under standard con- glucose by Pi. Is this a reasonable route? (Think about the
ditions but at pH 5.0, is more or less free energy released? solubility products of metabolic intermediates.)
Explain. (e) The ATP-coupled phosphorylation of glucose is cat-
alyzed in hepatocytes by the enzyme glucokinase. This en-
8. The G for Coupled Reactions Glucose 1-phos-
zyme binds ATP and glucose to form a glucose-ATP-enzyme
phate is converted into fructose 6-phosphate in two succes-
complex, and the phosphoryl group is transferred directly
sive reactions:
from ATP to glucose. Explain the advantages of this route.
Glucose 1-phosphate 88n glucose 6-phosphate
10. Calculations of G for ATP-Coupled Reactions
Glucose 6-phosphate 88n fructose 6-phosphate
From data in Table 13–6 calculate the G
value for the
Using the G
values in Table 13–4, calculate the equilibrium reactions
constant, K
eq, for the sum of the two reactions at 25 C: (a) Phosphocreatine  ADP 8n creatine  ATP
(b) ATP  fructose 8n ADP  fructose 6-phosphate
Glucose 1-phosphate 88n fructose 6-phosphate
11. Coupling ATP Cleavage to an Unfavorable Reaction
9. Strategy for Overcoming an Unfavorable Reaction: To explore the consequences of coupling ATP hydrolysis under
ATP-Dependent Chemical Coupling The phosphoryla- physiological conditions to a thermodynamically unfavorable
tion of glucose to glucose 6-phosphate is the initial step in biochemical reaction, consider the hypothetical transformation
the catabolism of glucose. The direct phosphorylation of glu- X n Y, for which G
 20 kJ/mol.
cose by Pi is described by the equation (a) What is the ratio [Y]/[X] at equilibrium?
Glucose  Pi 88n glucose 6-phosphate  H2O (b) Suppose X and Y participate in a sequence of reac-
G
 13.8 kJ/mol tions during which ATP is hydrolyzed to ADP and Pi. The
overall reaction is
(a) Calculate the equilibrium constant for the above re-
action. In the rat hepatocyte the physiological concentrations X  ATP  H2O 8n Y  ADP  Pi
of glucose and Pi are maintained at approximately 4.8 mM. Calculate [Y]/[X] for this reaction at equilibrium. Assume that
What is the equilibrium concentration of glucose 6-phosphate the equilibrium concentrations of ATP, ADP, and Pi are 1 M.
obtained by the direct phosphorylation of glucose by Pi? Does (c) We know that [ATP], [ADP], and [Pi] are not 1 M un-
this reaction represent a reasonable metabolic step for the der physiological conditions. Calculate [Y]/[X] for the ATP-
catabolism of glucose? Explain. coupled reaction when the values of [ATP], [ADP], and [Pi]
(b) In principle, at least, one way to increase the con- are those found in rat myocytes (Table 13–5).
centration of glucose 6-phosphate is to drive the equilibrium
reaction to the right by increasing the intracellular concen- 12. Calculations of G at Physiological Concentrations
trations of glucose and Pi. Assuming a fixed concentration of Calculate the physiological G (not G
) for the reaction
Pi at 4.8 mM, how high would the intracellular concentration
Phosphocreatine  ADP 8n creatine  ATP
of glucose have to be to give an equilibrium concentration of
glucose 6-phosphate of 250 M (the normal physiological con- at 25 C, as it occurs in the cytosol of neurons, with phos-
centration)? Would this route be physiologically reasonable, phocreatine at 4.7 mM, creatine at 1.0 mM, ADP at 0.73 mM,
given that the maximum solubility of glucose is less than 1 M? and ATP at 2.6 mM.
520 Chapter 13 Principles of Bioenergetics

13. Free Energy Required for ATP Synthesis under 18. Standard Reduction Potentials The standard re-
Physiological Conditions In the cytosol of rat hepato- duction potential, E
, of any redox pair is defined for the
cytes, the mass-action ratio, Q, is half-cell reaction:
[ATP] 1 Oxidizing agent  n electrons 8n reducing agent
  5.33  102 M
[ADP][Pi]
The E
values for the NAD/NADH and pyruvate/lactate con-
Calculate the free energy required to synthesize ATP in a rat jugate redox pairs are 0.32 V and 0.19 V, respectively.
hepatocyte. (a) Which conjugate pair has the greater tendency to
lose electrons? Explain.
14. Daily ATP Utilization by Human Adults
(b) Which is the stronger oxidizing agent? Explain.
(a) A total of 30.5 kJ/mol of free energy is needed to
(c) Beginning with 1 M concentrations of each reactant
synthesize ATP from ADP and Pi when the reactants and
and product at pH 7, in which direction will the following re-
products are at 1 M concentrations (standard state). Because
action proceed?
the actual physiological concentrations of ATP, ADP, and Pi
are not 1 M, the free energy required to synthesize ATP un- Pyruvate  NADH  H 34 lactate  NAD
der physiological conditions is different from G
. Calculate
the free energy required to synthesize ATP in the human he- (d) What is the standard free-energy change (G
) at
patocyte when the physiological concentrations of ATP, ADP, 25 C for the conversion of pyruvate to lactate?
and Pi are 3.5, 1.50, and 5.0 mM, respectively. (e) What is the equilibrium constant (K
eq) for this
(b) A 68 kg (150 lb) adult requires a caloric intake of reaction?
2,000 kcal (8,360 kJ) of food per day (24 h). The food is me- 19. Energy Span of the Respiratory Chain Electron
tabolized and the free energy is used to synthesize ATP, which transfer in the mitochondrial respiratory chain may be rep-
then provides energy for the body’s daily chemical and me- resented by the net reaction equation
chanical work. Assuming that the efficiency of converting
food energy into ATP is 50%, calculate the weight of ATP NADH  H 
1
 O2 34 H2O  NAD
2
used by a human adult in 24 h. What percentage of the body
weight does this represent?
(a) Calculate the value of E
for the net reaction of
(c) Although adults synthesize large amounts of ATP
mitochondrial electron transfer. Use E
values from Table
daily, their body weight, structure, and composition do not
13–7.
change significantly during this period. Explain this apparent
(b) Calculate G
for this reaction.
contradiction.
(c) How many ATP molecules can theoretically be gen-
15. Rates of Turnover of  and  Phosphates of ATP erated by this reaction if the free energy of ATP synthesis un-
If a small amount of ATP labeled with radioactive phospho- der cellular conditions is 52 kJ/mol?
rus in the terminal position, [-32P]ATP, is added to a yeast
20. Dependence of Electromotive Force on Concen-
extract, about half of the 32P activity is found in Pi within a
trations Calculate the electromotive force (in volts) regis-
few minutes, but the concentration of ATP remains un-
tered by an electrode immersed in a solution containing the
changed. Explain. If the same experiment is carried out us-
following mixtures of NAD and NADH at pH 7.0 and 25 C,
ing ATP labeled with 32P in the central position, [-32P]ATP,
with reference to a half-cell of E
0.00 V.
the 32P does not appear in Pi within such a short time. Why?
(a) 1.0 mM NAD and 10 mM NADH
16. Cleavage of ATP to AMP and PPi during Metabo- (b) 1.0 mM NAD and 1.0 mM NADH
lism The synthesis of the activated form of acetate (acetyl- (c) 10 mM NAD and 1.0 mM NADH
CoA) is carried out in an ATP-dependent process:
21. Electron Affinity of Compounds List the following
Acetate  CoA  ATP 8n acetyl-CoA  AMP  PPi substances in order of increasing tendency to accept elec-
trons: (a) -ketoglutarate  CO2 (yielding isocitrate); (b) ox-
(a) The G
for the hydrolysis of acetyl-CoA to acetate aloacetate; (c) O2; (d) NADP.
and CoA is 32.2 kJ/mol and that for hydrolysis of ATP to
AMP and PPi is 30.5 kJ/mol. Calculate G
for the ATP- 22. Direction of Oxidation-Reduction Reactions Which
dependent synthesis of acetyl-CoA. of the following reactions would you expect to proceed in the
(b) Almost all cells contain the enzyme inorganic py- direction shown, under standard conditions, assuming that
rophosphatase, which catalyzes the hydrolysis of PPi to Pi. the appropriate enzymes are present to catalyze them?
What effect does the presence of this enzyme have on the (a) Malate  NAD 8n oxaloacetate  NADH  H
synthesis of acetyl-CoA? Explain. (b) Acetoacetate  NADH  H 8n
-hydroxybutyrate  NAD
17. Energy for H Pumping The parietal cells of the (c) Pyruvate  NADH  H 8n lactate  NAD
stomach lining contain membrane “pumps” that transport hy- (d) Pyruvate  -hydroxybutyrate 8n
drogen ions from the cytosol of these cells (pH 7.0) into the lactate  acetoacetate
stomach, contributing to the acidity of gastric juice (pH 1.0). (e) Malate  pyruvate 8n oxaloacetate  lactate
Calculate the free energy required to transport 1 mol of hy- (f) Acetaldehyde  succinate On ethanol  fumarate
drogen ions through these pumps. (Hint: See Chapter 11.)
Assume a temperature of 25 C.

You might also like