chapter
O
xidative phosphorylation is the culmination of energy-
yielding metabolism in aerobic organisms. All oxi-
dative steps in the degradation of carbohydrates, fats,
and amino acids converge at this final stage of cellular
respiration, in which the energy of oxidation drives the
synthesis of ATP. Photophosphorylation is the means
by which photosynthetic organisms capture the energy
of sunlight—the ultimate source of energy in the bio-
sphere—and harness it to make ATP. Together, oxida-
tive phosphorylation and photophosphorylation account
for most of the ATP synthesized by most organisms most
of the time.
In eukaryotes, oxidative phosphorylation occurs in
mitochondria, photophosphorylation in chloroplasts.
Oxidative phosphorylation involves the reduction of O
2
to H
2
O with electrons donated by NADH and FADH
2
; it
occurs equally well in light or darkness. Photophosphor-
ylation involves the oxidation of H
2
O to O
2
, with
NADP
H11001
as ultimate electron acceptor; it is absolutely
dependent on the energy of light. Despite their differ-
ences, these two highly efficient energy-converting
processes have fundamentally similar mechanisms.
Our current understanding of ATP synthesis in mi-
tochondria and chloroplasts is based on the hypothesis,
introduced by Peter Mitchell in 1961, that transmem-
brane differences in proton concentration are the reser-
voir for the energy extracted from biological oxidation
reactions. This chemiosmotic theory has been ac-
cepted as one of the great unifying principles of twen-
tieth century biology. It provides insight into the
processes of oxidative phosphorylation and photophos-
phorylation, and into such apparently disparate energy
transductions as active transport across membranes and
the motion of bacterial flagella.
Oxidative phosphorylation and photophosphoryla-
tion are mechanistically similar in three respects. (1) Both
19
690
OXIDATIVE PHOSPHORYLATION
AND PHOTOPHOSPHORYLATION
If an idea presents itself to us, we must not reject it
simply because it does not agree with the logical
deductions of a reigning theory.
—Claude Bernard, An Introduction to the Study of
Experimental Medicine, 1813
The aspect of the present position of consensus that I
find most remarkable and admirable, is the altruism and
generosity with which former opponents of the
chemiosmotic hypothesis have not only come to accept it,
but have actively promoted it to the status of a theory.
—Peter Mitchell, Nobel Address, 1978
OXIDATIVE PHOSPHORYLATION
19.1 Electron-Transfer Reactions in Mitochondria 691
19.2 ATP Synthesis 704
19.3 Regulation of Oxidative Phosphorylation 716
19.4 Mitochondrial Genes: Their Origin and the Effects
of Mutations 719
19.5 The Role of Mitochondria in Apoptosis and
Oxidative Stress 721
PHOTOSYNTHESIS: HARVESTING LIGHT ENERGY
19.6 General Features of Photophosphorylation 723
19.7 Light Absorption 725
19.8 The Central Photochemical Event: Light-Driven
Electron Flow 730
19.9 ATP Synthesis by Photophosphorylation 740
8885d_c19_690-750 3/1/04 11:32 AM Page 690 mac76 mac76:385_reb:
processes involve the flow of electrons through a chain
of membrane-bound carriers. (2) The free energy made
available by this “downhill” (exergonic) electron flow
is coupled to the “uphill” transport of protons across a
proton-impermeable membrane, conserving the free
energy of fuel oxidation as a transmembrane electro-
chemical potential (p. 391). (3) The transmembrane
flow of protons down their concentration gradient
through specific protein channels provides the free
energy for synthesis of ATP, catalyzed by a membrane
protein complex (ATP synthase) that couples proton
flow to phosphorylation of ADP.
We begin this chapter with oxidative phosphoryla-
tion. We first describe the components of the electron-
transfer chain, their organization into large functional
complexes in the inner mitochondrial membrane, the
path of electron flow through them, and the proton
movements that accompany this flow. We then consider
the remarkable enzyme complex that, by “rotational
catalysis,” captures the energy of proton flow in ATP,
and the regulatory mechanisms that coordinate oxida-
tive phosphorylation with the many catabolic pathways
by which fuels are oxidized. With this understanding of
mitochondrial oxidative phosphorylation, we turn to
photophosphorylation, looking first at the absorption of
light by photosynthetic pigments, then at the light-
driven flow of electrons from H
2
O to NADP
H11001
and the
molecular basis for coupling electron and proton flow.
We also consider the similarities of structure and mech-
anism between the ATP synthases of chloroplasts and
mitochondria, and the evolutionary basis for this con-
servation of mechanism.
OXIDATIVE PHOSPHORYLATION
19.1 Electron-Transfer Reactions
in Mitochondria
The discovery in 1948 by Eugene Kennedy and Albert
Lehninger that mitochondria are the site of oxidative
phosphorylation in eukaryotes marked the beginning
of the modern phase of studies
in biological energy transduc-
tions. Mitochondria, like gram-
negative bacteria, have two
membranes (Fig. 19–1). The
outer mitochondrial membrane
is readily permeable to small
molecules (M
r
H110215,000) and
ions, which move freely
through transmembrane chan-
nels formed by a family of inte-
gral membrane proteins called
porins. The inner membrane is
impermeable to most small
molecules and ions, including protons (H
H11001
); the only
species that cross this membrane do so through specific
transporters. The inner membrane bears the compo-
nents of the respiratory chain and the ATP synthase.
The mitochondrial matrix, enclosed by the inner
membrane, contains the pyruvate dehydrogenase com-
plex and the enzymes of the citric acid cycle, the fatty
19.1 Electron-Transfer Reactions in Mitochondria 691
Outer membrane
Freely permeable to
small molecules and ions
ATP synthase
(F
o
F
1
)
Cristae
Impermeable to most
small molecules and ions,
including H
H11001
Contains:
Contains:
Ribosomes
Porin channels
? Respiratory electron
carriers (Complexes I–IV)
? ADP-ATP translocase
? ATP synthase (F
o
F
1
)
? Other membrane
transporters
? Pyruvate
dehydrogenase
complex
? Citric acid
cycle enzymes
? Amino acid
oxidation
enzymes
? DNA, ribosomes
? Many other enzymes
? ATP, ADP, P
i
, Mg
2H11001
, Ca
2H11001
, K
H11001
? Many soluble metabolic
intermediates
Inner membrane
Matrix
? Fatty acid
-oxidation
enzymes
H9252
Albert L. Lehninger,
1917–1986
FIGURE 19–1 Biochemical anatomy of a mitochondrion. The convo-
lutions (cristae) of the inner membrane provide a very large surface
area. The inner membrane of a single liver mitochondrion may have
more than 10,000 sets of electron-transfer systems (respiratory chains)
and ATP synthase molecules, distributed over the membrane surface.
Heart mitochondria, which have more profuse cristae and thus a much
larger area of inner membrane, contain more than three times as many
sets of electron-transfer systems as liver mitochondria. The mitochon-
drial pool of coenzymes and intermediates is functionally separate from
the cytosolic pool. The mitochondria of invertebrates, plants, and mi-
crobial eukaryotes are similar to those shown here, but with much vari-
ation in size, shape, and degree of convolution of the inner membrane.
8885d_c19_690-750 3/1/04 11:32 AM Page 691 mac76 mac76:385_reb:
acid H9252-oxidation pathway, and the pathways of amino
acid oxidation—all the pathways of fuel oxidation ex-
cept glycolysis, which takes place in the cytosol. The
selectively permeable inner membrane segregates the
intermediates and enzymes of cytosolic metabolic path-
ways from those of metabolic processes occurring in the
matrix. However, specific transporters carry pyruvate,
fatty acids, and amino acids or their H9251-keto derivatives
into the matrix for access to the machinery of the citric
acid cycle. ADP and P
i
are specifically transported into
the matrix as newly synthesized ATP is transported out.
Electrons Are Funneled to Universal
Electron Acceptors
Oxidative phosphorylation begins with the entry of elec-
trons into the respiratory chain. Most of these electrons
arise from the action of dehydrogenases that collect
electrons from catabolic pathways and funnel them into
universal electron acceptors—nicotinamide nucleotides
(NAD
H11001
or NADP
H11001
) or flavin nucleotides (FMN or FAD).
Nicotinamide nucleotide–linked dehydroge-
nases catalyze reversible reactions of the following gen-
eral types:
Reduced substrate H11001 NAD
H11001
oxidized substrate H11001 NADH H11001 H
H11001
Reduced substrate H11001 NADP
H11001
oxidized substrate H11001 NADPH H11001 H
H11001
Most dehydrogenases that act in catabolism are specific
for NAD
H11001
as electron acceptor (Table 19–1). Some are
z
y
z
y
in the cytosol, others are in mitochondria, and still oth-
ers have mitochondrial and cytosolic isozymes.
NAD-linked dehydrogenases remove two hydrogen
atoms from their substrates. One of these is transferred
as a hydride ion (
:
H
H11002
) to NAD
H11001
; the other is released
as H
H11001
in the medium (see Fig. 13–15). NADH and
NADPH are water-soluble electron carriers that associ-
ate reversibly with dehydrogenases. NADH carries elec-
trons from catabolic reactions to their point of entry into
the respiratory chain, the NADH dehydrogenase com-
plex described below. NADPH generally supplies elec-
trons to anabolic reactions. Cells maintain separate
pools of NADPH and NADH, with different redox po-
tentials. This is accomplished by holding the ratios of
[reduced form]/[oxidized form] relatively high for
NADPH and relatively low for NADH. Neither NADH nor
NADPH can cross the inner mitochondrial membrane,
but the electrons they carry can be shuttled across in-
directly, as we shall see.
Flavoproteins contain a very tightly, sometimes
covalently, bound flavin nucleotide, either FMN or FAD
(see Fig. 13–18). The oxidized flavin nucleotide can ac-
cept either one electron (yielding the semiquinone
form) or two (yielding FADH
2
or FMNH
2
). Electron
transfer occurs because the flavoprotein has a higher
reduction potential than the compound oxidized. The
standard reduction potential of a flavin nucleotide, un-
like that of NAD or NADP, depends on the protein with
which it is associated. Local interactions with functional
groups in the protein distort the electron orbitals in the
flavin ring, changing the relative stabilities of oxidized
and reduced forms. The relevant standard reduction
Chapter 19 Oxidative Phosphorylation and Photophosphorylation692
TABLE 19–1 Some Important Reactions Catalyzed by NAD(P)H-Linked Dehydrogenases
Reaction
*
Location
?
NAD-linked
H9251-Ketoglutarate H11001 CoA H11001NAD
H11001
succinyl-CoA H11001 CO
2
H11001 NADH H11001 H
H11001
M
L-Malate H11001 NAD
H11001
oxaloacetate H11001 NADH H11001 H
H11001
M and C
Pyruvate H11001 CoA H11001 NAD
H11001
acetyl-CoA H11001 CO
2
H11001 NADH H11001 H
H11001
M
Glyceraldehyde 3-phosphate H11001 P
i
H11001 NAD
H11001
1,3-bisphosphoglycerate H11001 NADH H11001 H
H11001
C
Lactate H11001 NAD
H11001
pyruvate H11001 NADH H11001 H
H11001
C
H9252-Hydroxyacyl-CoA H11001 NAD
H11001
H9252-ketoacyl-CoA H11001 NADH H11001 H
H11001
M
NADP-linked
Glucose 6-phosphate H11001 NADP
H11001
6-phosphogluconate H11001 NADPH H11001 H
H11001
C
NAD- or NADP-linked
L-Glutamate H11001 H
2
O H11001 NAD(P)
H11001
H9251-ketoglutarate H11001 NH
4
H11001
H11001 NAD(P)H M
Isocitrate H11001 NAD(P)
H11001
H9251-ketoglutarate H11001 CO
2
H11001 NAD(P)H H11001 H
H11001
M and C
z
y
z
y
z
y
z
y
z
y
z
y
z
y
z
y
z
y
*
These reactions and their enzymes are discussed in Chapters 14 through 18.
?
M designates mitochondria; C, cytosol.
8885d_c19_690-750 3/1/04 11:32 AM Page 692 mac76 mac76:385_reb:
potential is therefore that of the particular flavoprotein,
not that of isolated FAD or FMN. The flavin nucleotide
should be considered part of the flavoprotein’s active
site rather than a reactant or product in the electron-
transfer reaction. Because flavoproteins can participate
in either one- or two-electron transfers, they can serve
as intermediates between reactions in which two elec-
trons are donated (as in dehydrogenations) and those in
which only one electron is accepted (as in the reduction
of a quinone to a hydroquinone, described below).
Electrons Pass through a Series
of Membrane-Bound Carriers
The mitochondrial respiratory chain consists of a series
of sequentially acting electron carriers, most of which
are integral proteins with prosthetic groups capable of
accepting and donating either one or two electrons.
Three types of electron transfers occur in oxidative
phosphorylation: (1) direct transfer of electrons, as in
the reduction of Fe
3H11001
to Fe
2H11001
; (2) transfer as a hydro-
gen atom (H
H11001
H11001 e
H11002
); and (3) transfer as a hydride ion
(:H
H11002
), which bears two electrons. The term reducing
equivalent is used to designate a single electron equiv-
alent transferred in an oxidation-reduction reaction.
In addition to NAD and flavoproteins, three other
types of electron-carrying molecules function in the res-
piratory chain: a hydrophobic quinone (ubiquinone) and
two different types of iron-containing proteins (cyto-
chromes and iron-sulfur proteins). Ubiquinone (also
called coenzyme Q, or simply Q) is a lipid-soluble ben-
zoquinone with a long isoprenoid side chain (Fig. 19–2).
The closely related compounds plastoquinone (of plant
chloroplasts) and menaquinone (of bacteria) play roles
analogous to that of ubiquinone, carrying electrons in
membrane-associated electron-transfer chains. Ubiqui-
none can accept one electron to become the semi-
quinone radical (
H11080
QH) or two electrons to form ubiquinol
(QH
2
) (Fig. 19–2) and, like flavoprotein carriers, it can
act at the junction between a two-electron donor and a
one-electron acceptor. Because ubiquinone is both small
and hydrophobic, it is freely diffusible within the lipid
bilayer of the inner mitochondrial membrane and can
shuttle reducing equivalents between other, less mobile
electron carriers in the membrane. And because it car-
ries both electrons and protons, it plays a central role
in coupling electron flow to proton movement.
The cytochromes are proteins with characteristic
strong absorption of visible light, due to their iron-
containing heme prosthetic groups (Fig. 19–3). Mito-
chondria contain three classes of cytochromes, desig-
nated a, b, and c, which are distinguished by differences
in their light-absorption spectra. Each type of cyto-
chrome in its reduced (Fe
2H11001
) state has three absorp-
tion bands in the visible range (Fig. 19–4). The longest-
wavelength band is near 600 nm in type a cytochromes,
near 560 nm in type b, and near 550 nm in type c. To
distinguish among closely related cytochromes of one
type, the exact absorption maximum is sometimes used
in the names, as in cytochrome b
562
.
The heme cofactors of a and b cytochromes are
tightly, but not covalently, bound to their associated pro-
teins; the hemes of c-type cytochromes are covalently
attached through Cys residues (Fig. 19–3). As with the
flavoproteins, the standard reduction potential of the
heme iron atom of a cytochrome depends on its inter-
action with protein side chains and is therefore differ-
ent for each cytochrome. The cytochromes of type a
and b and some of type c are integral proteins of the
inner mitochondrial membrane. One striking exception
is the cytochrome c of mitochondria, a soluble protein
that associates through electrostatic interactions with
the outer surface of the inner
membrane. We encountered
cytochrome c in earlier dis-
cussions of protein structure
(see Fig. 4–18).
In iron-sulfur proteins,
first discovered by Helmut
Beinert, the iron is present not
in heme but in association
with inorganic sulfur atoms or
with the sulfur atoms of Cys
residues in the protein, or
both. These iron-sulfur (Fe-S)
19.1 Electron-Transfer Reactions in Mitochondria 693
?
O
HCCH
R
OH
CH
3
CH
3
O
(CH
2
O
CH
3
O
CH
3
CH
2
)
10
Ubiquinone (Q)
(fully oxidized)
Semiquinone radical
(
?
QH)
Ubiquinol (QH
2
)
(fully reduced)
H
H11001
H11001 e
H11002
O
CH
3
CH
3
O
CH
3
O
H
H11001
H11001 e
H11002
OH
OH
R
CH
3
CH
3
O
CH
3
O
FIGURE 19–2 Ubiquinone (Q, or coenzyme Q). Complete reduction
of ubiquinone requires two electrons and two protons, and occurs in
two steps through the semiquinone radical intermediate.
Helmut Beinert
8885d_c19_690-750 3/1/04 11:32 AM Page 693 mac76 mac76:385_reb:
centers range from simple structures with a single Fe
atom coordinated to four Cys OSH groups to more com-
plex Fe-S centers with two or four Fe atoms (Fig. 19–5).
Rieske iron-sulfur proteins (named after their dis-
coverer, John S. Rieske) are a variation on this theme,
in which one Fe atom is coordinated to two His residues
rather than two Cys residues. All iron-sulfur proteins
participate in one-electron transfers in which one iron
atom of the iron-sulfur cluster is oxidized or reduced.
At least eight Fe-S proteins function in mitochondrial
electron transfer. The reduction potential of Fe-S pro-
teins varies from H110020.65 V to H110010.45 V, depending on the
microenvironment of the iron within the protein.
In the overall reaction catalyzed by the mitochon-
drial respiratory chain, electrons move from NADH, suc-
cinate, or some other primary electron donor through
flavoproteins, ubiquinone, iron-sulfur proteins, and cy-
tochromes, and finally to O
2
. A look at the methods used
to determine the sequence in which the carriers act is
instructive, as the same general approaches have been
used to study other electron-transfer chains, such as
those of chloroplasts.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation694
Fe N
N
N
N
CH
3
CH
3
CH
2
CH
2
COO
H11002
CH
2
CH
CH
2
CH
CH
3
Heme A
(in a-type cytochromes)
Fe N
N
N
N
CH
3
CH
3
CH
3
CH
3
CH
2
CH
2
COO
H11002
CH
2
CH
2
CHO
CH
2
CHCH
2
Iron protoporphyrin IX
(in b-type cytochromes)
COO
H11002
CH
3
OH
Cys
S Cys
Fe N
N
N
N
CH
3
CH
3
CH
3
CH
3
CH
2
CH
2
COO
H11002
CH
2
CH
2
Heme C
(in c-type cytochromes)
COO
H11002
CH
3
CH
2
CH
S
CH
CH
2
CH
CH
3
CH
3
CH
3
CH
3
COO
H11002
FIGURE 19–3 Prosthetic groups of cytochromes.
Each group consists of four five-membered,
nitrogen-containing rings in a cyclic structure
called a porphyrin. The four nitrogen atoms are
coordinated with a central Fe ion, either Fe
2H11001
or
Fe
3H11001
. Iron protoporphyrin IX is found in b-type
cytochromes and in hemoglobin and myoglobin
(see Fig. 4–17). Heme c is covalently bound to
the protein of cytochrome c through thioether
bonds to two Cys residues. Heme a, found in the
a-type cytochromes, has a long isoprenoid tail
attached to one of the five-membered rings. The
conjugated double-bond system (shaded pink) of
the porphyrin ring accounts for the absorption of
visible light by these hemes.
100
Relative light absorption (%)
50
300 400 500 600
Wavelength (nm)
Oxidized
cyt c
Reduced
cyt c
H9251
H9252
H9253
0
FIGURE 19–4 Absorption spectra of cytochrome c (cyt c) in its oxi-
dized (red) and reduced (blue) forms. Also labeled are the character-
istic H9251, H9252, and H9253 bands of the reduced form.
8885d_c19_690-750 3/1/04 11:32 AM Page 694 mac76 mac76:385_reb:
First, the standard reduction potentials of the in-
dividual electron carriers have been determined ex-
perimentally (Table 19–2). We would expect the carri-
ers to function in order of increasing reduction
potential, because electrons tend to flow spontaneously
from carriers of lower EH11032H11034 to carriers of higher EH11032H11034. The
order of carriers deduced by this method is NADH →
Q → cytochrome b → cytochrome c
1
→ cytochrome
c → cytochrome a → cytochrome a
3
→ O
2
. Note, how-
ever, that the order of standard reduction potentials is
not necessarily the same as the order of actual reduc-
tion potentials under cellular conditions, which depend
on the concentration of reduced and oxidized forms
(p. 510). A second method for determining the sequence
19.1 Electron-Transfer Reactions in Mitochondria 695
S
(c)
Cys
Cys
Fe
S
S
S
S
S
S
S
Fe
Fe
Fe Cys
Cys
S
(b)
SCys Cys
Fe
S
Fe
S
S CysSCysS Cys
S
S
S
Cys
(a)
Cys Cys
Fe
Protein
(d)
FIGURE 19–5 Iron-sulfur centers. The Fe-S centers of iron-sulfur
proteins may be as simple as (a), with a single Fe ion surrounded
by the S atoms of four Cys residues. Other centers include both
inorganic and Cys S atoms, as in (b) 2Fe-2S or (c) 4Fe-4S centers.
(d) The ferredoxin of the cyanobacterium Anabaena 7120 has one
2Fe-2S center (PDB ID 1FRD); Fe is red, inorganic S
2
is yellow, and
the S of Cys is orange. (Note that in these designations only the
inorganic S atoms are counted. For example, in the 2Fe-2S center
(b), each Fe ion is actually surrounded by four S atoms.) The exact
standard reduction potential of the iron in these centers depends on
the type of center and its interaction with the associated protein.
TABLE 19–2 Standard Reduction Potentials of Respiratory Chain and Related Electron Carriers
Redox reaction (half-reaction) E H11032H11034 (V)
2H
H11001
H11001 2e
H11002
8n H
2
H110020.414
NAD
H11001
H11001 H
H11001
H11001 2e
H11002
8n NADH H110020.320
NADP
H11001
H11001 H
H11001
H11001 2e
H11002
8n NADPH H110020.324
NADH dehydrogenase (FMN) H11001 2H
H11001
H11001 2e
H11002
8n NADH dehydrogenase (FMNH
2
) H110020.30
Ubiquinone H11001 2H
H11001
H11001 2e
H11002
8n ubiquinol 0.045
Cytochrome b (Fe
3H11001
) H11001 e
H11002
8n cytochrome b (Fe
2H11001
) 0.077
Cytochrome c
1
(Fe
3H11001
) H11001 e
H11002
8n cytochrome c
1
(Fe
2H11001
) 0.22
Cytochrome c (Fe
3H11001
) H11001 e
H11002
8n cytochrome c (Fe
2H11001
) 0.254
Cytochrome a (Fe
3H11001
) H11001 e
H11002
8n cytochrome a (Fe
2H11001
) 0.29
Cytochrome a
3
(Fe
3H11001
) H11001 e
H11002
8n cytochrome a
3
(Fe
2H11001
) 0.35
H5007
1
2
H5007
O
2
H11001 2H
H11001
H11001 2e
H11002
8n H
2
O 0.8166
8885d_c19_690-750 3/1/04 11:32 AM Page 695 mac76 mac76:385_reb:
of electron carriers involves reducing the entire chain
of carriers experimentally by providing an electron
source but no electron acceptor (no O
2
). When O
2
is
suddenly introduced into the system, the rate at which
each electron carrier becomes oxidized (measured
spectroscopically) reveals the order in which the car-
riers function. The carrier nearest O
2
(at the end of the
chain) gives up its electrons first, the second carrier
from the end is oxidized next, and so on. Such exper-
iments have confirmed the sequence deduced from
standard reduction potentials.
In a final confirmation, agents that inhibit the flow
of electrons through the chain have been used in com-
bination with measurements of the degree of oxidation
of each carrier. In the presence of O
2
and an electron
donor, carriers that function before the inhibited step
become fully reduced, and those that function after this
step are completely oxidized (Fig. 19–6). By using sev-
eral inhibitors that block different steps in the chain, in-
vestigators have determined the entire sequence; it is
the same as deduced in the first two approaches.
Electron Carriers Function in Multienzyme Complexes
The electron carriers of the respiratory chain are or-
ganized into membrane-embedded supramolecular
complexes that can be physically separated. Gentle
treatment of the inner mitochondrial membrane with
detergents allows the resolution of four unique electron-
carrier complexes, each capable of catalyzing electron
transfer through a portion of the chain (Table 19–3; Fig.
19–7). Complexes I and II catalyze electron transfer to
ubiquinone from two different electron donors: NADH
(Complex I) and succinate (Complex II). Complex III
carries electrons from reduced ubiquinone to cyto-
chrome c, and Complex IV completes the sequence by
transferring electrons from cytochrome c to O
2
.
We now look in more detail at the structure and
function of each complex of the mitochondrial respira-
tory chain.
Complex I: NADH to Ubiquinone Figure 19–8 illustrates the
relationship between Complexes I and II and ubiquinone.
Complex I, also called NADH:ubiquinone oxidore-
ductase or NADH dehydrogenase, is a large enzyme
composed of 42 different polypeptide chains, including
an FMN-containing flavoprotein and at least six iron-
sulfur centers. High-resolution electron microscopy
shows Complex I to be L-shaped, with one arm of the L
in the membrane and the other extending into the ma-
trix. As shown in Figure 19–9, Complex I catalyzes two
simultaneous and obligately coupled processes: (1) the
Chapter 19 Oxidative Phosphorylation and Photophosphorylation696
NADH Q
Cyt c
1
Cyt (a H11001 a
3
) O
2
rotenone
antimycin A
CN
H11002
or CO
Cyt cCyt bNADH Q O
2
NADH Q O
2
Cyt c
1
Cyt cCyt b
Cyt c
1
Cyt Cyt cCyt b
Cyt (a H11001 a
3
)
a H11001 a
3
)(
FIGURE 19–6 Method for determining the
sequence of electron carriers. This method
measures the effects of inhibitors of electron
transfer on the oxidation state of each carrier.
In the presence of an electron donor and O
2
,
each inhibitor causes a characteristic pattern
of oxidized/reduced carriers: those before the
block become reduced (blue), and those after
the block become oxidized (pink).
TABLE 19–3 The Protein Components of the Mitochondrial Electron-Transfer Chain
Enzyme complex/protein Mass (kDa) Number of subunits
*
Prosthetic group(s)
I NADH dehydrogenase 850 43 (14) FMN, Fe-S
II Succinate dehydrogenase 140 4 FAD, Fe-S
III Ubiquinone cytochrome c oxidoreductase 250 11 Hemes, Fe-S
Cytochrome c
?
13 1 Heme
IV Cytochrome oxidase 160 13 (3–4) Hemes; Cu
A
,Cu
B
*
Numbers of subunits in the bacterial equivalents in parentheses.
?
Cytochrome c is not part of an enzyme complex; it moves between Complexes III and IV as a freely soluble protein.
8885d_c19_696 3/1/04 1:58 PM Page 696 mac76 mac76:385_reb:
exergonic transfer to ubiquinone of a hydride ion from
NADH and a proton from the matrix, expressed by
NADH H11001 H
H11001
H11001 Q On NAD
H11001
H11001 QH
2
(19–1)
and (2) the endergonic transfer of four protons from the
matrix to the intermembrane space. Complex I is there-
fore a proton pump driven by the energy of electron
transfer, and the reaction it catalyzes is vectorial: it
moves protons in a specific direction from one location
(the matrix, which becomes negatively charged with the
departure of protons) to another (the intermembrane
space, which becomes positively charged). To empha-
size the vectorial nature of the process, the overall re-
action is often written with subscripts that indicate the
location of the protons: P for the positive side of the in-
ner membrane (the intermembrane space), N for the
negative side (the matrix):
NADH H11001 5H
H11001
N
H11001 Q On NAD
H11001
H11001 QH
2
H11001 4H
H11001
P
(19–2)
19.1 Electron-Transfer Reactions in Mitochondria 697
Osmotic rupture
Inner
membrane
fragments
Outer membrane
fragments
discarded
ATP
synthase
IV
III
II
I
I II III IV ATP
synthase
NADH Q Suc-
cinate
Q Q Cyt c Cyt c O
2
ATP ADP
H11001
P
i
Reactions catalyzed by isolated
fractions in vitro
Solubilization with detergent
followed by ion-exchange chromatography
Treatment with digitonin
FIGURE 19–7 Separation of functional complexes of the respiratory
chain. The outer mitochondrial membrane is first removed by treat-
ment with the detergent digitonin. Fragments of inner membrane are
then obtained by osmotic rupture of the mitochondria, and the frag-
ments are gently dissolved in a second detergent. The resulting mix-
ture of inner membrane proteins is resolved by ion-exchange chro-
matography into different complexes (I through IV) of the respiratory
chain, each with its unique protein composition (see Table 19–3), and
the enzyme ATP synthase (sometimes called Complex V). The isolated
Complexes I through IV catalyze transfers between donors (NADH
and succinate), intermediate carriers (Q and cytochrome c), and O
2
,
as shown. In vitro, isolated ATP synthase has only ATP-hydrolyzing
(ATPase), not ATP-synthesizing, activity.
III
Intermembrane
space
Matrix
Fe-S
Fe-S
FAD
Glycerol
3-phosphate
(cytosolic)
glycerol
3-phosphate
dehydrogenase
FAD
FMN
NADH NAD
+
Succinate
ETF:Q
oxidoreductase
acyl-CoA
dehydrogenase
ETF
(FAD)
Fe-S
(FAD)
Fatty acyl–CoA
FAD
Q
FIGURE 19–8 Path of electrons from NADH, succinate, fatty
acyl–CoA, and glycerol 3-phosphate to ubiquinone. Electrons from
NADH pass through a flavoprotein to a series of iron-sulfur proteins
(in Complex I) and then to Q. Electrons from succinate pass through
a flavoprotein and several Fe-S centers (in Complex II) on the way to
Q. Glycerol 3-phosphate donates electrons to a flavoprotein (glycerol
3-phosphate dehydrogenase) on the outer face of the inner mito-
chondrial membrane, from which they pass to Q. Acyl-CoA dehydro-
genase (the first enzyme of H9252 oxidation) transfers electrons to electron-
transferring flavoprotein (ETF), from which they pass to Q via
ETF:ubiquinone oxidoreductase.
8885d_c19_690-750 3/1/04 11:32 AM Page 697 mac76 mac76:385_reb:
Amytal (a barbiturate drug), rotenone (a plant product
commonly used as an insecticide), and piericidin A (an
antibiotic) inhibit electron flow from the Fe-S centers
of Complex I to ubiquinone (Table 19–4) and therefore
block the overall process of oxidative phosphorylation.
Ubiquinol (QH
2
, the fully reduced form; Fig. 19–2)
diffuses in the inner mitochondrial membrane from
Complex I to Complex III, where it is oxidized to Q in a
process that also involves the outward movement of H
H11001
.
Complex II: Succinate to Ubiquinone We encountered
Complex II in Chapter 16 as succinate dehydroge-
nase, the only membrane-bound enzyme in the citric
acid cycle (p. 612). Although smaller and simpler than
Complex I, it contains five prosthetic groups of two
types and four different protein subunits (Fig. 19–10).
Subunits C and D are integral membrane proteins, each
with three transmembrane helices. They contain a heme
group, heme b, and a binding site for ubiquinone, the
final electron acceptor in the reaction catalyzed by
Complex II. Subunits A and B extend into the matrix (or
the cytosol of a bacterium); they contain three 2Fe-2S
centers, bound FAD, and a binding site for the substrate,
succinate. The path of electron transfer from the
succinate-binding site to FAD, then through the Fe-S
centers to the Q-binding site, is more than 40 ? long,
but none of the individual electron-transfer distances
exceeds about 11 ?—a reasonable distance for rapid
electron transfer (Fig. 19–10).
Chapter 19 Oxidative Phosphorylation and Photophosphorylation698
Complex I Intermembrane
space (P side)
Matrix (N side)
2H
+
4H
+
Fe-S
FMN
NADH NAD
+
H
+
H11001
2e
–
2e
–
N-2
Q
QH
2
Matrix
arm
Membrane
arm
FIGURE 19–9 NADH:ubiquinone oxidoreductase (Complex I). Com-
plex I catalyzes the transfer of a hydride ion from NADH to FMN, from
which two electrons pass through a series of Fe-S centers to the iron-
sulfur protein N-2 in the matrix arm of the complex. Electron transfer
from N-2 to ubiquinone on the membrane arm forms QH
2
, which dif-
fuses into the lipid bilayer. This electron transfer also drives the ex-
pulsion from the matrix of four protons per pair of electrons. The de-
tailed mechanism that couples electron and proton transfer in Complex
I is not yet known, but probably involves a Q cycle similar to that in
Complex III in which QH
2
participates twice per electron pair (see
Fig. 19–12). Proton flux produces an electrochemical potential across
the inner mitochondrial membrane (N side negative, P side positive),
which conserves some of the energy released by the electron-transfer
reactions. This electrochemical potential drives ATP synthesis.
TABLE 19–4 Agents That Interfere with Oxidative Phosphorylation or Photophosphorylation
Type of interference Compound
*
Target/mode of action
Inhibition of electron transfer Cyanide
Carbon monoxide
Antimycin A Blocks electron transfer from cytochrome b to cytochrome c
1
Myxothiazol
Rotenone
Amytal
Piericidin A
DCMU Competes with Q
B
for binding site in PSII
Inhibition of ATP synthase Aurovertin Inhibits F
1
Oligomycin
Venturicidin
DCCD Blocks proton flow through F
o
and CF
o
Uncoupling of phosphorylation FCCP
from electron transfer DNP
Hydrophobic proton carriers
Valinomycin K
H11001
ionophore
Thermogenin In brown fat, forms proton-conducting pores in inner mitochondrial
membrane
Inhibition of ATP-ADP exchange Atractyloside Inhibits adenine nucleotide translocase
*
DCMU is 3-(3,4-dichlorophenyl)-1,1-dimethylurea; DCCD, dicyclohexylcarbodiimide; FCCP, cyanide-p-trifluoromethoxyphenylhydrazone; DNP,
2,4-dinitrophenol.
Inhibit cytochrome oxidase
Prevent electron transfer from Fe-S center to ubiquinone
Inhibit F
o
and CF
o
?
?
?
?
?
?
?
?
?
?
?
?
?
?
8885d_c19_690-750 3/1/04 11:32 AM Page 698 mac76 mac76:385_reb:
flavoprotein acyl-CoA dehydrogenase (see Fig. 17–8),
involves transfer of electrons from the substrate to the
FAD of the dehydrogenase, then to electron-transferring
flavoprotein (ETF), which in turn passes its electrons to
ETF : ubiquinone oxidoreductase (Fig. 19–8). This
enzyme transfers electrons into the respiratory chain by
reducing ubiquinone. Glycerol 3-phosphate, formed ei-
ther from glycerol released by triacylglycerol breakdown
or by the reduction of dihydroxyacetone phosphate
from glycolysis, is oxidized by glycerol 3-phosphate
dehydrogenase (see Fig. 17–4). This enzyme is a flavo-
protein located on the outer face of the inner mito-
chondrial membrane, and like succinate dehydrogenase
and acyl-CoA dehydrogenase it channels electrons into
the respiratory chain by reducing ubiquinone (Fig.
19–8). The important role of glycerol 3-phosphate de-
hydrogenase in shuttling reducing equivalents from
cytosolic NADH into the mitochondrial matrix is de-
scribed in Section 19.2 (see Fig. 19–28). The effect of
each of these electron-transferring enzymes is to con-
tribute to the pool of reduced ubiquinone. QH
2
from all
these reactions is reoxidized by Complex III.
Complex III: Ubiquinone to Cytochrome c The next respi-
ratory complex, Complex III, also called cytochrome
bc
1
complex or ubiquinone:cytochrome c oxidore-
ductase, couples the transfer of electrons from
ubiquinol (QH
2
) to cytochrome c with the vectorial
transport of protons from the matrix to the intermem-
brane space. The determination of the complete struc-
ture of this huge complex (Fig. 19–11) and of Complex
IV (below) by x-ray crystallography, achieved between
1995 and 1998, were landmarks in the study of mito-
chondrial electron transfer, providing the structural
framework to integrate the many biochemical observa-
tions on the functions of the respiratory complexes.
Based on the structure of Complex III and detailed
biochemical studies of the redox reactions, a reasonable
model has been proposed for the passage of electrons
19.1 Electron-Transfer Reactions in Mitochondria 699
The heme b of Complex II is apparently not in
the direct path of electron transfer; it may serve
instead to reduce the frequency with which electrons
“leak” out of the system, moving from succinate to mo-
lecular oxygen to produce the reactive oxygen species
(ROS) hydrogen peroxide (H
2
O
2
) and the superoxide
radical (
H11080
O
2
H11002
) described in Section 19.5. Humans with
point mutations in Complex II subunits near heme b or
the quinone-binding site suffer from hereditary para-
ganglioma. This inherited condition is characterized by
benign tumors of the head and neck, commonly in the
carotid body, an organ that senses O
2
levels in the blood.
These mutations result in greater production of ROS
and perhaps greater tissue damage during succinate
oxidation. ■
Other substrates for mitochondrial dehydrogenases
pass electrons into the respiratory chain at the level of
ubiquinone, but not through Complex II. The first step
in the H9252 oxidation of fatty acyl–CoA, catalyzed by the
FIGURE 19–10 Structure of Complex II (succinate dehydrogenase)
of E. coli (PDB ID 1NEK). The enzyme has two transmembrane sub-
units, C (green) and D (blue); the cytoplasmic extensions contain sub-
units B (orange) and A (purple). Just behind the FAD in subunit A (gold)
is the binding site for succinate (occupied in this crystal structure by
the inhibitor oxaloacetate, green). Subunit B has three sets of Fe-S cen-
ters (yellow and red); ubiquinone (yellow) is bound to subunit C; and
heme b (purple) is sandwiched between subunits C and D. A cardi-
olipin molecule is so tightly bound to subunit C that it shows up in
the crystal structure (gray spacefilling). Electrons move (blue arrows)
from succinate to FAD, then through the three Fe-S centers to
ubiquinone. The heme b is not on the main path of electron transfer
but protects against the formation of reactive oxygen species (ROS) by
electrons that go astray.
Substrate
binding
site
Cytoplasm
(N side)
C
B
D
A
FAD
Fe-S
centers
Periplasm
(P side)
Cardiolipin
Ubiquinone
Heme bQH
2
8885d_c19_690-750 3/1/04 11:32 AM Page 699 mac76 mac76:385_reb:
and protons through the complex. The net equation for
the redox reactions of this Q cycle (Fig. 19–12) is
QH
2
H11001 2 cyt c
1
(oxidized) H11001 2H
H11001
N O
n
Q H11001 2 cyt c
1
(reduced) H11001 4H
H11001
P
(19–3)
The Q cycle accommodates the switch between the two-
electron carrier ubiquinone and the one-electron carri-
ers—cytochromes b
562
, b
566
, c
1
, and c—and explains the
measured stoichiometry of four protons translocated
per pair of electrons passing through the Complex III to
cytochrome c. Although the path of electrons through
this segment of the respiratory chain is complicated, the
net effect of the transfer is simple: QH
2
is oxidized to Q
and two molecules of cytochrome c are reduced.
Cytochrome c (see Fig. 4–18) is a soluble protein of
the intermembrane space. After its single heme accepts
an electron from Complex III, cytochrome c moves to
Complex IV to donate the electron to a binuclear cop-
per center.
Complex IV: Cytochrome c to O
2
In the final step of the
respiratory chain, Complex IV, also called cytochrome
oxidase, carries electrons from cytochrome c to mo-
lecular oxygen, reducing it to H
2
O. Complex IV is a large
enzyme (13 subunits; M
r
204,000) of the inner mito-
chondrial membrane. Bacteria contain a form that is
much simpler, with only three or four subunits, but still
capable of catalyzing both electron transfer and proton
pumping. Comparison of the mitochondrial and bacter-
ial complexes suggests that three subunits are critical
to the function (Fig. 19–13).
Mitochondrial subunit II contains two Cu ions com-
plexed with the OSH groups of two Cys residues in a
binuclear center (Cu
A
; Fig. 19–13b) that resembles the
2Fe-2S centers of iron-sulfur proteins. Subunit I con-
tains two heme groups, designated a and a
3
, and an-
other copper ion (Cu
B
). Heme a
3
and Cu
B
form a sec-
ond binuclear center that accepts electrons from heme
a and transfers them to O
2
bound to heme a
3
.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation700
FIGURE 19–11 Cytochrome bc
1
complex (Complex III). The com-
plex is a dimer of identical monomers, each with 11 different sub-
units. (a) Structure of a monomer. The functional core is three sub-
units: cytochrome b (green) with its two hemes (b
H
and b
L
, light red);
the Rieske iron-sulfur protein (purple) with its 2Fe-2S centers (yellow);
and cytochrome c
1
(blue) with its heme (red) (PDB ID 1BGY). (b) The
dimeric functional unit. Cytochrome c
1
and the Rieske iron-sulfur pro-
tein project from the P surface and can interact with cytochrome c
(not part of the functional complex) in the intermembrane space. The
complex has two distinct binding sites for ubiquinone, Q
N
and Q
P
,
which correspond to the sites of inhibition by two drugs that block
oxidative phosphorylation. Antimycin A, which blocks electron flow
from heme b
H
to Q, binds at Q
N
, close to heme b
H
on the N (matrix)
side of the membrane. Myxothiazol, which prevents electron flow from
QH
2
to the Rieske iron-sulfur protein, binds at Q
P
, near the 2Fe-2S
center and heme b
L
on the P side. The dimeric structure is essential
to the function of Complex III. The interface between monomers forms
two pockets, each containing a Q
P
site from one monomer and a Q
N
site from the other. The ubiquinone intermediates move within these
sheltered pockets.
Complex III crystallizes in two distinct conformations (not shown).
In one, the Rieske Fe-S center is close to its electron acceptor, the
heme of cytochrome c
1
, but relatively distant from cytochrome b and
the QH
2
-binding site at which the Rieske Fe-S center receives elec-
trons. In the other, the Fe-S center has moved away from cytochrome
c
1
and toward cytochrome b. The Rieske protein is thought to oscil-
late between these two conformations as it is reduced, then oxidized.
(a)
Intermembrane
space (P side)
Matrix
(N side)
Cytochrome c
1
Cytochrome b
Rieske iron-
sulfur protein
2Fe-2S
(b)
Cytochrome c
1
Cytochrome c
Rieske iron-
sulfur protein
2Fe-2S
center
Cytochrome b
(P side)
(N side)
b
L
Q
P
Q
N
b
H
c
1
Heme
8885d_c19_690-750 3/1/04 11:32 AM Page 700 mac76 mac76:385_reb:
Electron transfer through Complex IV is from cyto-
chrome c to the Cu
A
center, to heme a, to the heme
a
3
–Cu
B
center, and finally to O
2
(Fig. 19–14). For every
four electrons passing through this complex, the enzyme
consumes four “substrate” H
H11001
from the matrix (N side)
in converting O
2
to 2H
2
O. It also uses the energy of this
redox reaction to pump one proton outward into the in-
termembrane space (P side) for each electron that
passes through, adding to the electrochemical potential
produced by redox-driven proton transport through
Complexes I and III. The overall reaction catalyzed by
Complex IV is
4 Cyt c (reduced) H11001 8H
H11001
N
H11001 O
2 O
n
4 cyt c (oxidized) H11001 4H
H11001
P
H11001 2H
2
O (19–4)
This four-electron reduction of O
2
involves redox cen-
ters that carry only one electron at a time, and it must
occur without the release of incompletely reduced
intermediates such as hydrogen peroxide or hydroxyl
free radicals—very reactive species that would damage
cellular components. The intermediates remain tightly
bound to the complex until completely converted to
water.
The Energy of Electron Transfer Is Efficiently
Conserved in a Proton Gradient
The transfer of two electrons from NADH through the
respiratory chain to molecular oxygen can be written as
NADH H11001 H
H11001
H11001
H5007
1
2
H5007
O
2 O
n NAD
H11001
H11001 H
2
O (19–5)
This net reaction is highly exergonic. For the redox pair
NAD
H11001
/NADH, EH11032H11034 is H110020.320 V, and for the pair O
2
/H
2
O,
EH11032H11034 is 0.816 V. The H9004EH11032H11034 for this reaction is therefore
1.14 V, and the standard free-energy change (see Eqn
13–6, p. 510) is
H9004GH11032H11034 H11005 H11002n H9004EH11032H11034 (19–6)
H11005H110022(96.5 kJ/V
H11080
mol)(1.14 V)
H11005H11002220 kJ/mol (of NADH)
This standard free-energy change is based on the as-
sumption of equal concentrations (1 M) of NADH and
19.1 Electron-Transfer Reactions in Mitochondria 701
b
H
b
L
Cyt c
1
Cyt c
Oxidation of
first QH
2
Fe-S
Q
2H
+
?
Q
–
QH
2
Q
b
H
b
L
Cyt c
1
Cyt c
Fe-S
2H
+
Q
Oxidation of
second QH
2
Matrix (N side)
Intermembrane
space (P side)
QH
2
2H
+
?
Q
–
QH
2
QH
2
cyt c
1
(oxidized)H11001
QH
2
2 cyt c
1
(oxidized)H11001
2H
P
H11001
H11001
cyt c
1
(reduced)H11001
Q 2 cyt c
1
(reduced)H11001H11001
H11001
?
Q
H11002
?
Q
H11002
H11001
4H
P
H11001
2H
P
H11001
2H
N
H11001
H11001 2H
N
H11001
QH
2
QH
2
cyt c
1
(oxidized)H11001
QH11001H11001cyt c
1
(reduced)H11001
Net equation:
FIGURE 19–12 The Q cycle. The path of electrons through Complex
III is shown by blue arrows. On the P side of the membrane, two mol-
ecules of QH
2
are oxidized to Q near the P side, releasing two pro-
tons per Q (four protons in all) into the intermembrane space. Each
QH
2
donates one electron (via the Rieske Fe-S center) to cytochrome
c
1
, and one electron (via cytochrome b) to a molecule of Q near the
N side, reducing it in two steps to QH
2
. This reduction also uses two
protons per Q, which are taken up from the matrix.
8885d_c19_690-750 3/1/04 11:32 AM Page 701 mac76 mac76:385_reb:
NAD
H11001
. In actively respiring mitochondria, the actions of
many dehydrogenases keep the actual [NADH]/[NAD
H11001
]
ratio well above unity, and the real free-energy change
for the reaction shown in Equation 19–5 is therefore
substantially greater (more negative) than H11002220 kJ/mol.
A similar calculation for the oxidation of succinate
shows that electron transfer from succinate (EH11032H11034 for
fumarate/succinate H11005 0.031 V) to O
2
has a smaller, but
still negative, standard free-energy change of about
H11002150 kJ/mol.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation702
(a) (b)
FIGURE 19–13 Critical subunits of cytochrome oxidase (Complex
IV). The bovine complex is shown here (PDB ID 1OCC). (a) The core
of Complex IV, with three subunits. Subunit I (yellow) has two heme
groups, a and a
3
(red), and a copper ion, Cu
B
(green sphere). Heme
a
3
and Cu
B
form a binuclear Fe-Cu center. Subunit II (blue) contains
two Cu ions (green spheres) complexed with the OSH groups of two
Cys residues in a binuclear center, Cu
A
, that resembles the 2Fe-2S cen-
ters of iron-sulfur proteins. This binuclear center and the cytochrome
c–binding site are located in a domain of subunit II that protrudes
from the P side of the inner membrane (into the intermembrane space).
Subunit III (green) seems to be essential for Complex IV function, but
its role is not well understood. (b) The binuclear center of Cu
A
. The
Cu ions (green spheres) share electrons equally. When the center is
reduced they have the formal charges Cu
1H11001
Cu
1H11001
; when oxidized,
Cu
1.5H11001
Cu
1.5H11001
. Ligands around the Cu ions include two His (dark blue),
two Cys (yellow), an Asp (red), and Met (orange) residues.
Subunit
II
Subunit
I
4H
+
4H
+
(pumped)
Subunit
III
4H
+
(substrate)
4e
–
2H
2
O
O
2
4Cyt c
Cu
B
Cu
A
Fe-Cu
center
a
3
a
Intermembrane
space
(P side)
Matrix
(N side)
FIGURE 19–14 Path of electrons through Complex IV. The three pro-
teins critical to electron flow are subunits I, II, and III. The larger green
structure includes the other ten proteins in the complex. Electron trans-
fer through Complex IV begins when two molecules of reduced cy-
tochrome c (top) each donate an electron to the binuclear center Cu
A
.
From here electrons pass through heme a to the Fe-Cu center (cy-
tochrome a
3
and Cu
B
). Oxygen now binds to heme a
3
and is reduced
to its peroxy derivative (O
2
2H11002
) by two electrons from the Fe-Cu center.
Delivery of two more electrons from cytochrome c (making four elec-
trons in all) converts the O
2
2H11002
to two molecules of water, with con-
sumption of four “substrate” protons from the matrix. At the same time,
four more protons are pumped from the matrix by an as yet unknown
mechanism.
8885d_c19_690-750 3/1/04 11:32 AM Page 702 mac76 mac76:385_reb:
Much of this energy is used to pump protons out of
the matrix. For each pair of electrons transferred to O
2
,
four protons are pumped out by Complex I, four by Com-
plex III, and two by Complex IV (Fig. 19–15). The vec-
torial equation for the process is therefore
NADH H11001 11H
H11001
N
H11001
H5007
1
2
H5007
O
2 O
n NAD
H11001
H11001 10H
H11001
P
H11001 H
2
O (19–7)
The electrochemical energy inherent in this difference
in proton concentration and separation of charge rep-
resents a temporary conservation of much of the energy
of electron transfer. The energy stored in such a gradi-
ent, termed the proton-motive force, has two com-
ponents: (1) the chemical potential energy due to the
difference in concentration of a chemical species (H
H11001
)
in the two regions separated by the membrane, and (2)
the electrical potential energy that results from the
separation of charge when a proton moves across the
membrane without a counterion (Fig. 19–16).
As we showed in Chapter 11, the free-energy
change for the creation of an electrochemical gradient
by an ion pump is
H9004G H11005 RT ln
H20898H20899
H11001 Z H9004H9274 (19–8)
where C
2
and C
1
are the concentrations of an ion in two
regions, and C
2
H11022 C
1
; Z is the absolute value of its elec-
trical charge (1 for a proton), and H9004H9274 is the transmem-
brane difference in electrical potential, measured in volts.
For protons at 25 H11034C,
ln
H20898H20899
H11005 2.3(log [H
H11001
]
P
H11002 log [H
H11001
]
N
)
H11005 2.3(pH
N
H11002 pH
P
) H11005 2.3 H9004pH
and Equation 19–8 reduces to
H9004G H11005 2.3RT H9004pH H11001H9004H9274 (19–9)
H11005 (5.70 kJ/mol)H9004pH H11001 (96.5 kJ/V
H11080
mol)?H9274
In actively respiring mitochondria, the measured ?H9274 is
0.15 to 0.20 V and the pH of the matrix is about 0.75
C
2
H5007
C
1
C
2
H5007
C
1
units more alkaline than that of the intermembrane
space, so the calculated free-energy change for pump-
ing protons outward is about 20 kJ/mol (of H
H11001
), most
of which is contributed by the electrical portion of the
electrochemical potential. Because the transfer of two
electrons from NADH to O
2
is accompanied by the out-
ward pumping of 10 H
H11001
(Eqn 19–7), roughly 200 kJ of
the 220 kJ released by oxidation of a mole of NADH is
conserved in the proton gradient.
When protons flow spontaneously down their elec-
trochemical gradient, energy is made available to do
work. In mitochondria, chloroplasts, and aerobic bacte-
ria, the electrochemical energy in the proton gradient
drives the synthesis of ATP from ADP and P
i
. We return
to the energetics and stoichiometry of ATP synthesis
driven by the electrochemical potential of the proton
gradient in Section 19.2.
19.1 Electron-Transfer Reactions in Mitochondria 703
Intermembrane
space (P side)
Matrix (N side)
4H
+
1
–
2
O
2
+ 2H
+
H
2
O
Succinate Fumarate
4H
+
II
NADH
+
H
+
NAD
+
Cyt c
IV
2H
+
III
I
Q
FIGURE 19–15 Summary of the flow of electrons and protons
through the four complexes of the respiratory chain. Electrons reach
Q through Complexes I and II. QH
2
serves as a mobile carrier of elec-
trons and protons. It passes electrons to Complex III, which passes
them to another mobile connecting link, cytochrome c. Complex IV
then transfers electrons from reduced cytochrome c to O
2
. Electron
flow through Complexes I, III, and IV is accompanied by proton flow
from the matrix to the intermembrane space. Recall that electrons from
H9252 oxidation of fatty acids can also enter the respiratory chain through
Q (see Fig. 19–8).
N side
[H
H11001
]
N
H11005 C
1
OH
H11002
OH
H11002
OH
H11002
OH
H11002
OH
H11002
OH
H11002
OH
H11002
H
H11001
H
H11001
H
H11001
H
H11001
H
H11001
H
H11001
H
H11001
P side
[H
H11001
]
P
H11005 C
2
H9004G H11005 RT ln (C
2
/C
1
) H11001 Z? H9004
H9274
H11005 2.3RT H9004pH H11001 ? ?
H9274
H
H11001
Proton
pump
H9274
FIGURE 19–16 Proton-motive force. The inner mitochondrial mem-
brane separates two compartments of different [H
H11001
], resulting in dif-
ferences in chemical concentration (H9004pH) and charge distribution
(H9004H9274) across the membrane. The net effect is the proton-motive force
(H9004G), which can be calculated as shown here. This is explained more
fully in the text.
8885d_c19_690-750 3/1/04 11:32 AM Page 703 mac76 mac76:385_reb:
Plant Mitochondria Have Alternative Mechanisms
for Oxidizing NADH
Plant mitochondria supply the cell with ATP during
periods of low illumination or darkness by mechanisms
entirely analogous to those used by nonphotosynthetic
organisms. In the light, the principal source of mito-
chondrial NADH is a reaction in which glycine, produced
by a process known as photorespiration, is converted to
serine (see Fig. 20–21):
2 Glycine H11001 NAD
H11001
88n serine H11001 CO
2
H11001 NH
3
H11001 NADH H11001 H
H11001
For reasons discussed in Chapter 20, plants must carry
out this reaction even when they do not need NADH for
ATP production. To regenerate NAD
H11001
from unneeded
NADH, plant mitochondria transfer electrons from NADH
directly to ubiquinone and from ubiquinone directly to
O
2
, bypassing Complexes III and IV and their proton
pumps. In this process the energy in NADH is dissipated
as heat, which can sometimes be of value to the plant
(Box 19–1). Unlike cytochrome oxidase (Complex IV),
the alternative QH
2
oxidase is not inhibited by cyanide.
Cyanide-resistant NADH oxidation is therefore the hall-
mark of this unique plant electron-transfer pathway.
SUMMARY 19.1 Electron-Transfer Reactions
in Mitochondria
■ Chemiosmotic theory provides the intellectual
framework for understanding many biological
energy transductions, including oxidative
phosphorylation and photophosphorylation.
The mechanism of energy coupling is similar in
both cases: the energy of electron flow is
conserved by the concomitant pumping of
protons across the membrane, producing an
electrochemical gradient, the proton-motive
force.
■ In mitochondria, hydride ions removed from
substrates by NAD-linked dehydrogenases
donate electrons to the respiratory
(electron-transfer) chain, which transfers the
electrons to molecular O
2
, reducing it to H
2
O.
■ Shuttle systems convey reducing equivalents
from cytosolic NADH to mitochondrial NADH.
Reducing equivalents from all NAD-linked
dehydrogenations are transferred to mito-
chondrial NADH dehydrogenase (Complex I).
■ Reducing equivalents are then passed through
a series of Fe-S centers to ubiquinone, which
transfers the electrons to cytochrome b, the
first carrier in Complex III. In this complex,
electrons take two separate paths through two
b-type cytochromes and cytochrome c
1
to an
Fe-S center. The Fe-S center passes electrons,
one at a time, through cytochrome c and into
Complex IV, cytochrome oxidase. This
copper-containing enzyme, which also contains
cytochromes a and a
3
, accumulates electrons,
then passes them to O
2
, reducing it to H
2
O.
■ Some electrons enter this chain of carriers
through alternative paths. Succinate is oxidized
by succinate dehydrogenase (Complex II),
which contains a flavoprotein that passes
electrons through several Fe-S centers to
ubiquinone. Electrons derived from the
oxidation of fatty acids pass to ubiquinone via
the electron-transferring flavoprotein.
■ Plants also have an alternative, cyanide-resistant
NADH oxidation pathway.
19.2 ATP Synthesis
How is a concentration gradient of protons transformed
into ATP? We have seen that electron transfer releases,
and the proton-motive force conserves, more than
enough free energy (about 200 kJ) per “mole” of elec-
tron pairs to drive the formation
of a mole of ATP, which requires
about 50 kJ (see Box 13–1). Mi-
tochondrial oxidative phospho-
rylation therefore poses no
thermodynamic problem. But
what is the chemical mechanism
that couples proton flux with
phosphorylation?
The chemiosmotic model,
proposed by Peter Mitchell, is
the paradigm for this mecha-
nism. According to the model
(Fig. 19–17), the electrochemi-
cal energy inherent in the difference in proton concen-
tration and separation of charge across the inner mito-
chondrial membrane—the proton-motive force—drives
the synthesis of ATP as protons flow passively back into
the matrix through a proton pore associated with ATP
synthase. To emphasize this crucial role of the proton-
motive force, the equation for ATP synthesis is some-
times written
ADP H11001 P
i
H11001 nH
H11001
P O
n ATP H11001 H
2
O H11001 nH
H11001
N
(19–10)
Mitchell used “chemiosmotic” to describe enzymatic re-
actions that involve, simultaneously, a chemical reaction
and a transport process. The operational definition of
“coupling” is shown in Figure 19–18. When isolated mi-
tochondria are suspended in a buffer containing ADP,
P
i
, and an oxidizable substrate such as succinate, three
easily measured processes occur: (1) the substrate is
oxidized (succinate yields fumarate), (2) O
2
is consumed,
and (3) ATP is synthesized. Oxygen consumption and
ATP synthesis depend on the presence of an oxidizable
substrate (succinate in this case) as well as ADP and P
i
.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation704
Peter Mitchell,
1920–1992
8885d_c19_690-750 3/1/04 11:32 AM Page 704 mac76 mac76:385_reb:
Because the energy of substrate oxidation drives
ATP synthesis in mitochondria, we would expect in-
hibitors of the passage of electrons to O
2
(such as
cyanide, carbon monoxide, and antimycin A) to block
ATP synthesis (Fig. 19–18a). More surprising is the find-
ing that the converse is also true: inhibition of ATP syn-
thesis blocks electron transfer in intact mitochondria.
This obligatory coupling can be demonstrated in isolated
mitochondria by providing O
2
and oxidizable substrates,
but not ADP (Fig. 19–18b). Under these conditions, no
ATP synthesis can occur and electron transfer to O
2
does not proceed. Coupling of oxidation and phosphor-
ylation can also be demonstrated using oligomycin or
venturicidin, toxic antibiotics that bind to the ATP syn-
thase in mitochondria. These compounds are potent in-
hibitors of both ATP synthesis and the transfer of elec-
trons through the chain of carriers to O
2
(Fig. 19–18b).
Because oligomycin is known to interact not directly with
the electron carriers but with ATP synthase, it follows
that electron transfer and ATP synthesis are obligately
coupled; neither reaction occurs without the other.
Chemiosmotic theory readily explains the depend-
ence of electron transfer on ATP synthesis in mitochon-
dria. When the flow of protons into the matrix through
the proton channel of ATP synthase is blocked (with
oligomycin, for example), no path exists for the return
of protons to the matrix, and the continued extrusion
of protons driven by the activity of the respiratory chain
generates a large proton gradient. The proton-motive
force builds up until the cost (free energy) of pumping
19.2 ATP Synthesis 705
NADH + H
+
NAD
+
Succinate
Fumarate
Cyt c
+
–
ADP + P
i
ATP
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
–
–
–
–
–
– –
–
–
–
–
–
4H
+
4H
+
2H
+
H
+
Chemical
potential
?pΗ
(inside
alkaline)
ATP
synthesis
driven by
proton-motive
force
Electrical
potential
?w
(inside
negative)
+
+
–
–O
2
+
H
2
2H
+
O
2
1
II
IV
I
III
F
o
F
1
Intermembrane
space
Matrix
Q
FIGURE 19–17 Chemiosmotic model. In this
simple representation of the chemiosmotic
theory applied to mitochondria, electrons from
NADH and other oxidizable substrates pass
through a chain of carriers arranged asymmet-
rically in the inner membrane. Electron flow is
accompanied by proton transfer across the
membrane, producing both a chemical
gradient (H9004pH) and an electrical gradient (H9004H9274).
The inner mitochondrial membrane is imper-
meable to protons; protons can reenter the
matrix only through proton-specific channels
(F
o
). The proton-motive force that drives
protons back into the matrix provides the
energy for ATP synthesis, catalyzed by the F
1
complex associated with F
o
.
O
2
consumed
Add
ADP H11001 P
i
Add
succinate
Time(b)
ATP synthesized
Add
venturicidin
or
oligomycin
Add
DNP
Uncoupled
O
2
consumed
Add
ADP H11001 P
i
Add
succinate
Time(a)
ATP synthesized
Add
CN
H11002
FIGURE 19–18 Coupling of electron transfer and ATP synthesis in
mitochondria. In experiments to demonstrate coupling, mitochondria
are suspended in a buffered medium and an O
2
electrode monitors O
2
consumption. At intervals, samples are removed and assayed for the
presence of ATP. (a) Addition of ADP and P
i
alone results in little or no
increase in either respiration (O
2
consumption; black) or ATP synthe-
sis (red). When succinate is added, respiration begins immediately and
ATP is synthesized. Addition of cyanide (CN
H11002
), which blocks electron
transfer between cytochrome oxidase and O
2
, inhibits both respiration
and ATP synthesis. (b) Mitochondria provided with succinate respire
and synthesize ATP only when ADP and P
i
are added. Subsequent ad-
dition of venturicidin or oligomycin, inhibitors of ATP synthase, blocks
both ATP synthesis and respiration. Dinitrophenol (DNP) is an un-
coupler, allowing respiration to continue without ATP synthesis.
8885d_c19_690-750 3/1/04 11:32 AM Page 705 mac76 mac76:385_reb:
Chapter 19 Oxidative Phosphorylation and Photophosphorylation706
BOX 19–1 THE WORLD OF BIOCHEMISTRY
Hot, Stinking Plants and Alternative Respiratory
Pathways
Many flowering plants attract insect pollinators by re-
leasing odorant molecules that mimic an insect’s nat-
ural food sources or potential egg-laying sites. Plants
pollinated by flies or beetles that normally feed on or
lay their eggs in dung or carrion sometimes use foul-
smelling compounds to attract these insects.
One family of stinking plants is the Araceae, which
includes philodendrons, arum lilies, and skunk cab-
bages. These plants have tiny flowers densely packed
on an erect structure, the spadix, surrounded by a
modified leaf, the spathe. The spadix releases odors
of rotting flesh or dung. Before pollination the spadix
also heats up, in some species to as much as 20 to
40 H11034C above the ambient temperature. Heat produc-
tion (thermogenesis) helps evaporate odorant mole-
cules for better dispersal, and because rotting flesh
and dung are usually warm from the hyperactive me-
tabolism of scavenging microbes, the heat itself might
also attract insects. In the case of the eastern skunk
cabbage (Fig. 1), which flowers in late winter or early
spring when snow still covers the ground, thermogen-
esis allows the spadix to grow up through the snow.
How does a skunk cabbage heat its spadix? The
mitochondria of plants, fungi, and unicellular eukary-
otes have electron-transfer systems that are essen-
tially the same as those in animals, but they also
have an alternative respiratory pathway. A cyanide-
resistant QH
2
oxidase transfers electrons from the
ubiquinone pool directly to oxygen, bypassing the two
proton-translocating steps of Complexes III and IV
(Fig. 2). Energy that might have been conserved as
ATP is instead released as heat. Plant mitochondria
also have an alternative NADH dehydrogenase, insen-
sitive to the Complex I inhibitor rotenone (see Table
19–4), that transfers electrons from NADH in the ma-
trix directly to ubiquinone, bypassing Complex I and
its associated proton pumping. And plant mitochon-
dria have yet another NADH dehydrogenase, on the
external face of the inner membrane, that transfers
electrons from NADPH or NADH in the intermem-
brane space to ubiquinone, again bypassing Complex
I. Thus when electrons enter the alternative respira-
tory pathway through the rotenone-insensitive NADH
dehydrogenase, the external NADH dehydrogenase,
or succinate dehydrogenase (Complex II), and pass to
O
2
via the cyanide-resistant alternative oxidase, en-
ergy is not conserved as ATP but is released as heat.
A skunk cabbage can use the heat to melt snow, pro-
duce a foul stench, or attract beetles or flies.
FIGURE 1 Eastern skunk cabbage.
FIGURE 2 Electron carriers of the inner membrane of plant mitochondria. Electrons can flow
through Complexes I, III, and IV, as in animal mitochondria, or through plant-specific alterna-
tive carriers by the paths shown with blue arrows.
Heat
IV
I
Intermembrane
space
Q
Matrix
Cyt c
NAD
+
NAD(P)
+
External NAD(P)H
dehydrogenase
Alternative
oxidase
Alternative
NADH
dehydrogenase
III
H
2
O
NADH
NAD(P)H
1
–
2
O
2
8885d_c19_706 3/1/04 1:59 PM Page 706 mac76 mac76:385_reb:
protons out of the matrix against this gradient equals or
exceeds the energy released by the transfer of electrons
from NADH to O
2
. At this point electron flow must stop;
the free energy for the overall process of electron flow
coupled to proton pumping becomes zero, and the sys-
tem is at equilibrium.
Certain conditions and reagents, however, can un-
couple oxidation from phosphorylation. When intact mi-
tochondria are disrupted by treatment with detergent or
by physical shear, the resulting membrane fragments can
still catalyze electron transfer from succinate or NADH
to O
2
, but no ATP synthesis is coupled to this respiration.
Certain chemical compounds cause uncoupling without
disrupting mitochondrial structure. Chemical uncouplers
include 2,4-dinitrophenol (DNP) and carbonylcyanide-p-
trifluoromethoxyphenylhydrazone (FCCP) (Table 19–4;
Fig. 19–19), weak acids with hydrophobic properties
that permit them to diffuse readily across mitochondrial
membranes. After entering the matrix in the protonated
form, they can release a proton, thus dissipating the
proton gradient. Ionophores such as valinomycin (see
Fig. 11–45) allow inorganic ions to pass easily through
membranes. Ionophores uncouple electron transfer from
oxidative phosphorylation by dissipating the electrical
contribution to the electrochemical gradient across the
mitochondrial membrane.
19.2 ATP Synthesis 707
A prediction of the chemiosmotic theory is that, be-
cause the role of electron transfer in mitochondrial ATP
synthesis is simply to pump protons to create the elec-
trochemical potential of the proton-motive force, an ar-
tificially created proton gradient should be able to re-
place electron transfer in driving ATP synthesis. This
has been experimentally confirmed (Fig. 19–20). Mito-
chondria manipulated so as to impose a difference of
proton concentration and a separation of charge across
the inner membrane synthesize ATP in the absence of
an oxidizable substrate; the proton-motive force alone
suffices to drive ATP synthesis.
FIGURE 19–19 Two chemical uncouplers of oxidative phosphoryla-
tion. Both DNP and FCCP have a dissociable proton and are very
hydrophobic. They carry protons across the inner mitochondrial mem-
brane, dissipating the proton gradient. Both also uncouple photo-
phosphorylation (see Fig. 19–57).
FIGURE 19–20 Evidence for the role of a proton gradient in ATP syn-
thesis. An artificially imposed electrochemical gradient can drive ATP
synthesis in the absence of an oxidizable substrate as electron donor.
In this two-step experiment, (a) isolated mitochondria are first incu-
bated in a pH 9 buffer containing 0.1 M KCl. Slow leakage of buffer
and KCl into the mitochondria eventually brings the matrix into equi-
librium with the surrounding medium. No oxidizable substrates are
present. (b) Mitochondria are now separated from the pH 9 buffer and
resuspended in pH 7 buffer containing valinomycin but no KCl. The
change in buffer creates a difference of two pH units across the inner
mitochondrial membrane. The outward flow of K
H11001
, carried (by vali-
nomycin) down its concentration gradient without a counterion, cre-
ates a charge imbalance across the membrane (matrix negative). The
sum of the chemical potential provided by the pH difference and the
electrical potential provided by the separation of charges is a proton-
motive force large enough to support ATP synthesis in the absence of
an oxidizable substrate.
N
NH
H11001 H
H11001
N
H11002
2,4-Dinitrophenol
(DNP)
Carbonylcyanide-p-
trifluoromethoxyphenylhydrazone
(FCCP)
OH
NO
2
O
H11002
H11001 H
H11001
NO
2
NO
2
NO
2
N
C
CC
NN
C
CC
NN
O
CFF
F
O
CFF
F
[K
H11001
] H11005 [Cl
H11002
] H11005 0.1 M
[H
H11001
] H11005 10
H110029
M
[H
H11001
] H11005 10
H110029
M
F
o
F
1
pH lowered from 9 to 7;
valinomycin present; no K
H11001
(a)
[K
H11001
] H11021 [Cl
H11002
]
[H
H11001
] H11005 10
H110029
M
(b)
[H
H11001
] H11005
10
H11002
7
M
ADPH11001 P
i
K
H11001
K
H11001
H11001
H11001
H11001
H11001
H11001
H11001
H11001
H11002
H11002
H11002
H11002
H11002
H11002
H11002 H11002
ATP
Matrix
Intermembrane
space
8885d_c19_690-750 3/1/04 11:32 AM Page 707 mac76 mac76:385_reb:
ATP Synthase Has Two Functional Domains, F
o
and F
1
Mitochondrial ATP synthase is an F-type ATPase (see
Fig. 11–39; Table 11–3) similar in structure and mech-
anism to the ATP synthases of chloroplasts and eubac-
teria. This large enzyme complex of the inner mito-
chondrial membrane catalyzes the formation of ATP
from ADP and P
i
, accompanied
by the flow of protons from the
P to the N side of the mem-
brane (Eqn 19–10). ATP syn-
thase, also called Complex V,
has two distinct components:
F
1
, a peripheral membrane
protein, and F
o
(o denoting
oligomycin-sensitive), which is
integral to the membrane. F
1
,
the first factor recognized as
essential for oxidative phos-
phorylation, was identified and
purified by Efraim Racker and
his colleagues in the early 1960s.
In the laboratory, small membrane vesicles formed
from inner mitochondrial membranes carry out ATP syn-
thesis coupled to electron transfer. When F
1
is gently
extracted, the “stripped” vesicles still contain intact res-
piratory chains and the F
o
portion of ATP synthase. The
vesicles can catalyze electron transfer from NADH to O
2
but cannot produce a proton gradient: F
o
has a proton
pore through which protons leak as fast as they are
pumped by electron transfer, and without a proton gra-
dient the F
1
-depleted vesicles cannot make ATP. Iso-
lated F
1
catalyzes ATP hydrolysis (the reversal of syn-
thesis) and was therefore originally called F
1
ATPase.
When purified F
1
is added back to the depleted vesicles,
it reassociates with F
o
, plugging its proton pore and
restoring the membrane’s capacity to couple electron
transfer and ATP synthesis.
ATP Is Stabilized Relative to ADP on the
Surface of F
1
Isotope exchange experiments with purified F
1
reveal
a remarkable fact about the enzyme’s catalytic mecha-
nism: on the enzyme surface, the reaction ADP H11001 P
i
ATP H11001 H
2
O is readily reversible—the free-energy
change for ATP synthesis is close to zero! When ATP is
hydrolyzed by F
1
in the presence of
18
O-labeled water,
the P
i
released contains an
18
O atom. Careful meas-
urement of the
18
O content of P
i
formed in vitro by F
1
-
catalyzed hydrolysis of ATP reveals that the P
i
has not
one, but three or four
18
O atoms (Fig. 19–21). This in-
dicates that the terminal pyrophosphate bond in ATP
is cleaved and re-formed repeatedly before P
i
leaves the
enzyme surface. With P
i
free to tumble in its binding
site, each hydrolysis inserts
18
O randomly at one of the
z
y
Chapter 19 Oxidative Phosphorylation and Photophosphorylation708
Efraim Racker,
1913–1991
18
O
18
O
ADP
ATP H11001 H
2
18
O
H11001
18
OP
18
O
Enzyme
(F
1
)
(a)
ADP
a-Arg
376
b-Arg
182
b-Glu
181
b-Lys
155
Mg
2+
FIGURE 19–21 Catalytic mechanism of F
1
. (a)
18
O-exchange exper-
iment. F
1
solubilized from mitochondrial membranes is incubated with
ATP in the presence of
18
O-labeled water. At intervals, a sample of
the solution is withdrawn and analyzed for the incorporation of
18
O
into the P
i
produced from ATP hydrolysis. In minutes, the P
i
contains
three or four
18
O atoms, indicating that both ATP hydrolysis and ATP
synthesis have occurred several times during the incubation. (b) The
likely transition state complex for ATP hydrolysis and synthesis in ATP
synthase (derived from PDB ID 1BMF). The H9251 subunit is shown in
green, H9252 in gray. The positively charged residues H9252-Arg
182
and H9251-Arg
376
coordinate two oxygens of the pentavalent phosphate intermediate; H9252-
Lys
155
interacts with a third oxygen, and the Mg
2H11001
ion (green sphere)
further stabilizes the intermediate. The blue sphere represents the leav-
ing group (H
2
O). These interactions result in the ready equilibration
of ATP and ADP H11001 P
i
in the active site.
(b)
8885d_c19_690-750 3/1/04 11:32 AM Page 708 mac76 mac76:385_reb:
four positions in the molecule. This exchange reaction
occurs in unenergized F
o
F
1
complexes (with no proton
gradient) and with isolated F
1
—the exchange does not
require the input of energy.
Kinetic studies of the initial rates of ATP synthesis
and hydrolysis confirm the conclusion that H9004GH11032H11034 for ATP
synthesis on the enzyme is near zero. From the meas-
ured rates of hydrolysis (k
1
H11005 10 s
H110021
) and synthesis
(k
H110021
H11005 24 s
H110021
), the calculated equilibrium constant for
the reaction
Enz-ATP Enz–(ADP H11001 P
i
)
is
KH11032
eq
H11005H11005 H110052.4
From this KH11032
eq
, the calculated apparent H9004GH11032H11034 is close to
zero. This is much different from the KH11032
eq
of about 10
5
(H9004GH11032H11034 H11005 H1100230.5 kJ/mol) for the hydrolysis of ATP free in
solution (not on the enzyme surface).
What accounts for the huge difference? ATP syn-
thase stabilizes ATP relative to ADP H11001 P
i
by binding ATP
more tightly, releasing enough energy to counterbalance
the cost of making ATP. Careful measurements of the
binding constants show that F
o
F
1
binds ATP with very
high affinity (K
d
≤ 10
H1100212
M) and ADP with much lower
affinity (K
d
≈ 10
H110025
M). The difference in K
d
corresponds
to a difference of about 40 kJ/mol in binding energy, and
this binding energy drives the equilibrium toward for-
mation of the product ATP.
The Proton Gradient Drives the Release of ATP
from the Enzyme Surface
Although ATP synthase equilibrates ATP with ADP H11001
P
i
, in the absence of a proton gradient the newly syn-
thesized ATP does not leave the surface of the enzyme.
24 s
H110021
H5007
10 s
H110021
k
H110021
H5007
k
1
z
y
It is the proton gradient that causes the enzyme to re-
lease the ATP formed on its surface. The reaction co-
ordinate diagram of the process (Fig. 19–22) illustrates
the difference between the mechanism of ATP synthase
and that of many other enzymes that catalyze ender-
gonic reactions.
For the continued synthesis of ATP, the enzyme
must cycle between a form that binds ATP very tightly
and a form that releases ATP. Chemical and crystallo-
graphic studies of the ATP synthase have revealed the
structural basis for this alternation in function.
Each H9252 Subunit of ATP Synthase Can Assume Three
Different Conformations
Mitochondrial F
1
has nine subunits of five different
types, with the composition H9251
3
H9252
3
H9253H9254H9255. Each of the three
H9252 subunits has one catalytic site for ATP synthesis. The
crystallographic determination of the F
1
structure by
John E. Walker and colleagues revealed structural de-
tails very helpful in explaining the catalytic mechanism
of the enzyme. The knoblike portion of F
1
is a flattened
sphere, 8 nm high and 10 nm across, consisting of al-
ternating H9251 and H9252 subunits arranged like the sections of
an orange (Fig. 19–23a–c). The polypeptides that make
up the stalk in the F
1
crystal structure are asymmetri-
cally arranged, with one domain of the single H9253 subunit
making up a central shaft that passes through F
1
, and
another domain of H9253 associated primarily with one of
the three H9252 subunits, designated H9252-empty (Fig. 19–23c).
Although the amino acid sequences of the three H9252 sub-
units are identical, their conformations differ, in part
because of the association of the H9253 subunit with just one
of the three. The structures of the H9254 and H9255 subunits are
not revealed in these crystallographic studies.
The conformational differences among H9252 subunits
extend to differences in their ATP/ADP-binding sites.
19.2 ATP Synthesis 709
G
(kJ/mol)
Reaction coordinate
80
60
40
20
0
?
P
ADPH11001P
i
ES
E H11001 S
E ADPH11001P
i
[E ATP]
Typical enzyme ATP synthase
ATP
(in solution)
FIGURE 19–22 Reaction coordinate diagrams for
ATP synthase and for a more typical enzyme. In a
typical enzyme-catalyzed reaction (left), reaching
the transition state (?) between substrate and
product is the major energy barrier to overcome.
In the reaction catalyzed by ATP synthase (right),
release of ATP from the enzyme, not formation of
ATP, is the major energy barrier. The free-energy
change for the formation of ATP from ADP and P
i
in aqueous solution is large and positive, but on
the enzyme surface, the very tight binding of ATP
provides sufficient binding energy to bring the free
energy of the enzyme-bound ATP close to that of
ADP H11001 P
i
, so the reaction is readily reversible.
The equilibrium constant is near 1. The free
energy required for the release of ATP is provided
by the proton-motive force.
8885d_c19_690-750 3/1/04 11:32 AM Page 709 mac76 mac76:385_reb:
When researchers crystallized the protein in the pres-
ence of ADP and App(NH)p, a close structural analog
of ATP that cannot be hydrolyzed by the ATPase activ-
ity of F
1
, the binding site of one of the three H9252 subunits
was filled with App(NH)p, the second was filled with
Chapter 19 Oxidative Phosphorylation and Photophosphorylation710
b-ATP
b-ADP
b-empty
a-ADP
a-empty
a-ATP
(c)
ATP
ADP
(b)
H9251
H9251H9251
H9252
H9252
H9252
H9253
(a)
FIGURE 19–23 Mitochondrial ATP synthase complex. (a) Structure
of the F
1
complex, deduced from crystallographic and biochemical
studies. In F
1
, three H9251 and three H9252 subunits are arranged like the seg-
ments of an orange, with alternating H9251 (shades of gray) and H9252 (shades
of purple) subunits around a central shaft, the H9253 subunit (green). (b)
Crystal structure of bovine F
1
(PDB ID 1BMF), viewed from the side.
Two H9251 subunits and one H9252 subunit have been omitted to reveal the
central shaft (H9253 subunit) and the binding sites for ATP (red) and ADP
(yellow) on the H9252 subunits. The H9254 and H9255 subunits are not shown here.
(c) F
1
viewed from above (that is, from the N side of the membrane),
showing the three H9252 and three H9251 subunits and the central shaft (H9253 sub-
unit, green). Each H9252 subunit, near its interface with the neighboring H9251
subunit, has a nucleotide-binding site critical to the catalytic activity.
The single H9253 subunit associates primarily with one of the three H9251H9252
pairs, forcing each of the three H9252 subunits into slightly different con-
formations, with different nucleotide-binding sites. In the crystalline
enzyme, one subunit (H9252-ADP) has ADP (yellow) in its binding site, the
next (H9252-ATP) has ATP (red), and the third (H9252-empty) has no bound nu-
cleotide. (d) Side view of the F
o
F
1
structure. This is a composite, in
which the crystallographic coordinates of bovine mitochondrial F
1
(shades of purple and gray) have been combined with those of yeast
mitochondrial F
o
(shades of yellow and orange) (PDB ID 1QO1). Sub-
units a, b, H9254, and H9255 were not part of the crystal structure shown here.
(e) The F
o
F
1
structure, viewed end-on in the direction P side to N side.
The major structures visible in this cross section are the two trans-
membrane helices of each of ten c subunits arranged in concentric
circles. (f) Diagram of the F
o
F
1
complex, deduced from biochemical
and crystallographic studies. The two b subunits of F
o
associate firmly
with the H9251 and H9252 subunits of F
1
, holding them fixed relative to the
membrane. In F
o
, the membrane-embedded cylinder of c subunits is
attached to the shaft made up of F
1
subunits H9253 and H9255. As protons flow
through the membrane from the P side to the N side through F
o
, the
cylinder and shaft rotate, and the H9252 subunits of F
1
change conforma-
tion as the H9253 subunit associates with each in turn.
John E. Walker
Nonhydrolyzable
CH
2
O P P
H
NO
O OO
O
H
N
N
NH
2
N
N
H
OHOH
HH
P O
H11002
O
H11002
O
H11002
O
H11002
H9252H9253H9251
H9252-H9253 bondH9252
App(NH)p (H9252,H9253-imidoadenosine 5H11032-triphosphate)
8885d_c19_690-750 3/1/04 11:32 AM Page 710 mac76 mac76:385_reb:
F
1
F
o
(d)
ADP, and the third was empty. The corresponding H9252
subunit conformations are designated H9252-ATP, H9252-ADP,
and H9252-empty (Fig. 19–23c). This difference in nucleo-
tide binding among the three subunits is critical to the
mechanism of the complex.
The F
o
complex making up the proton pore is
composed of three subunits, a, b, and c, in the propor-
tion ab
2
c
10–12
. Subunit c is a small (M
r
8,000), very
hydrophobic polypeptide, consisting almost entirely of
two transmembrane helices, with a small loop extend-
ing from the matrix side of the membrane. The crystal
structure of the yeast F
o
F
1
, solved in 1999, shows the
arrangement of the c subunits. The yeast complex has
ten c subunits, each with two transmembrane helices
roughly perpendicular to the plane of the membrane and
arranged in two concentric circles (Fig. 19–23d, e). The
inner circle is made up of the amino-terminal helices of
each c subunit; the outer circle, about 55 ? in diame-
ter, is made up of the carboxyl-terminal helices. The H9255
and H9253 subunits of F
1
form a leg-and-foot that projects
from the bottom (membrane) side of F
1
and stands
firmly on the ring of c subunits. The schematic drawing
in Figure 19–23f combines the structural information
from studies of bovine F
1
and yeast F
o
F
1
.
Rotational Catalysis Is Key to the Binding-Change
Mechanism for ATP Synthesis
On the basis of detailed kinetic and binding studies of
the reactions catalyzed by F
o
F
1
, Paul Boyer proposed
a rotational catalysis mechanism in which the three
active sites of F
1
take turns catalyzing ATP synthesis
19.2 ATP Synthesis 711
(e)
H9251
H9280
H9251H9251
H9252
H9252
H9252
H9253
H9254
b
2
c
10
H
+
a
N side
P side
ADP + P
i
ATP
(f)
8885d_c19_690-750 3/1/04 11:32 AM Page 711 mac76 mac76:385_reb:
(Fig. 19–24). A given H9252 sub-
unit starts in the H9252-ADP con-
formation, which binds ADP
and P
i
from the surrounding
medium. The subunit now
changes conformation, assum-
ing the H9252-ATP form that
tightly binds and stabilizes
ATP, bringing about the ready
equilibration of ADP H11001 P
i
with
ATP on the enzyme surface.
Finally, the subunit changes to
the H9252-empty conformation, which has very low affinity
for ATP, and the newly synthesized ATP leaves the en-
zyme surface. Another round of catalysis begins when
this subunit again assumes the H9252-ADP form and binds
ADP and P
i
.
The conformational changes central to this mecha-
nism are driven by the passage of protons through the
F
o
portion of ATP synthase. The streaming of protons
through the F
o
“pore” causes the cylinder of c subunits
and the attached H9253 subunit to rotate about the long axis
of H9253, which is perpendicular to the plane of the mem-
brane. The H9253 subunit passes through the center of the
H9251
3
H9252
3
spheroid, which is held stationary relative to the
membrane surface by the b
2
and H9254 subunits (Fig.
19–23f). With each rotation of 120H11034, H9253 comes into con-
tact with a different H9252 subunit, and the contact forces
that H9252 subunit into the H9252-empty conformation.
The three H9252 subunits interact in such a way that
when one assumes the H9252-empty conformation, its neigh-
bor to one side must assume the H9252-ADP form, and the
other neighbor the H9252-ATP form. Thus one complete ro-
tation of the H9253 subunit causes each H9252 subunit to cycle
through all three of its possible conformations, and for
each rotation, three ATP are synthesized and released
from the enzyme surface.
One strong prediction of this binding-change model
is that the H9253 subunit should rotate in one direction when
F
o
F
1
is synthesizing ATP and in the opposite direction
when the enzyme is hydrolyzing ATP. This prediction
was confirmed in elegant experiments in the laborato-
ries of Masasuke Yoshida and Kazuhiko Kinosita, Jr. The
rotation of H9253 in a single F
1
molecule was observed mi-
croscopically by attaching a long, thin, fluorescent actin
polymer to H9253 and watching it move relative to H9251
3
H9252
3
im-
mobilized on a microscope slide, as ATP was hydrolyzed.
When the entire F
o
F
1
complex (not just F
1
) was used
in a similar experiment, the entire ring of c subunits ro-
tated with H9253 (Fig. 19–25). The “shaft” rotated in the pre-
dicted direction through 360H11034. The rotation was not
smooth, but occurred in three discrete steps of 120H11034. As
calculated from the known rate of ATP hydrolysis by
one F
1
molecule and from the frictional drag on the long
actin polymer, the efficiency of this mechanism in con-
verting chemical energy into motion is close to 100%. It
is, in Boyer’s words, “a splendid molecular machine!”
Chemiosmotic Coupling Allows Nonintegral
Stoichiometries of O
2
Consumption and ATP
Synthesis
Before the general acceptance of the chemiosmotic
model for oxidative phosphorylation, the assumption
was that the overall reaction equation would take the
following form:
xADP H11001 xP
i
H11001
H5007
1
2
H5007
O
2
H11001 H
H11001
H11001 NADH 88n
xATP H11001 H
2
O H11001 NAD
H11001
(19–11)
with the value of x—sometimes called the P/O ratio or
the P/2e
H11546
ratio—always an integer. When intact mito-
Chapter 19 Oxidative Phosphorylation and Photophosphorylation712
ATP
ATP
ADP
+P
i
H9252
H9251
H9252
H9251 H9252
H9251
ATP
ATP
ADP
+P
i
H9252
H9251
H9252
H9251
H9252
H9251
ATP
ATP
ADP
+P
i
H9252
H9251
H9252
H9251
H9252
H9251
3 H
P
+
3 H
P
+
3 H
N
+
3 H
N
+
3 H
N
+
3 H
P
+
FIGURE 19–24 Binding-change model for ATP synthase. The F
1
com-
plex has three nonequivalent adenine nucleotide–binding sites, one
for each pair of H9251 and H9252 subunits. At any given moment, one of these
sites is in the H9252-ATP conformation (which binds ATP tightly), a second
is in the H9252-ADP (loose-binding) conformation, and a third is in the H9252-
empty (very-loose-binding) conformation. The proton-motive force
causes rotation of the central shaft—the H9253 subunit, shown as a green
arrowhead—which comes into contact with each H9251H9252 subunit pair in
succession. This produces a cooperative conformational change in
which the H9252-ATP site is converted to the H9252-empty conformation, and
ATP dissociates; the H9252-ADP site is converted to the H9252-ATP conforma-
tion, which promotes condensation of bound ADP H11001 P
i
to form ATP;
and the H9252-empty site becomes a H9252-ADP site, which loosely binds ADP
H11001 P
i
entering from the solvent. This model, based on experimental
findings, requires that at least two of the three catalytic sites alternate
in activity; ATP cannot be released from one site unless and until ADP
and P
i
are bound at the other.
Paul Boyer
8885d_c19_690-750 3/1/04 11:32 AM Page 712 mac76 mac76:385_reb:
ADP + P
i
ATP
H9251
H9251
H9254 H9252
H9253
Ni complex
His
residues His residues
Avidin
F
o
F
1
a
b
c
Actin
filament
chondria are suspended in solution with an oxidizable
substrate such as succinate or NADH and are provided
with O
2
, ATP synthesis is readily measurable, as is the
decrease in O
2
. Measurement of P/O, however, is com-
plicated by the fact that intact mitochondria consume
ATP in many reactions taking place in the matrix, and
they consume O
2
for purposes other than oxidative
phosphorylation. Most experiments have yielded P/O
(ATP to H5007
1
2
H5007O
2
) ratios of between 2 and 3 when NADH
was the electron donor, and between 1 and 2 when suc-
cinate was the donor. Given the assumption that P/O
should have an integral value, most experimenters
agreed that the P/O ratios must be 3 for NADH and 2
for succinate, and for years those values have appeared
in research papers and textbooks.
With introduction of the chemiosmotic paradigm for
coupling ATP synthesis to electron transfer, there was
no theoretical requirement for P/O to be integral. The
relevant questions about stoichiometry became, how
many protons are pumped outward by electron transfer
from one NADH to O
2
, and how many protons must flow
inward through the F
o
F
1
complex to drive the synthe-
sis of one ATP? The measurement of proton fluxes is
technically complicated; the investigator must take into
account the buffering capacity of mitochondria, non-
productive leakage of protons across the inner mem-
brane, and use of the proton gradient for functions other
than ATP synthesis, such as driving the transport of sub-
strates across the inner mitochondrial membrane (de-
scribed below). The consensus values for number of pro-
tons pumped out per pair of electrons are 10 for NADH
and 6 for succinate. The most widely accepted experi-
mental value for number of protons required to drive
the synthesis of an ATP molecule is 4, of which 1 is used
in transporting P
i
, ATP, and ADP across the mitochon-
drial membrane (see below). If 10 protons are pumped
out per NADH and 4 must flow in to produce 1 ATP, the
proton-based P/O ratio is 2.5 for NADH as the electron
donor and 1.5 (6/4) for succinate. We use the P/O val-
ues of 2.5 and 1.5 throughout this book, but the values
3.0 and 2.0 are still common in the biochemical litera-
ture. The final word on proton stoichiometry will prob-
ably not be written until we know the full details of the
F
o
F
1
reaction mechanism.
The Proton-Motive Force Energizes Active Transport
Although the primary role of the proton gradient in mi-
tochondria is to furnish energy for the synthesis of ATP,
the proton-motive force also drives several transport
processes essential to oxidative phosphorylation. The
inner mitochondrial membrane is generally imperme-
able to charged species, but two specific systems trans-
port ADP and P
i
into the matrix and ATP out to the cy-
tosol (Fig. 19–26).
The adenine nucleotide translocase, integral to
the inner membrane, binds ADP
3H11002
in the intermembrane
space and transports it into the matrix in exchange for
an ATP
4H11002
molecule simultaneously transported outward
(see Fig. 13–1 for the ionic forms of ATP and ADP). Be-
cause this antiporter moves four negative charges out
for every three moved in, its activity is favored by the
19.2 ATP Synthesis 713
FIGURE 19–25 Rotation of F
o
and H9253 experimentally demonstrated.
F
1
genetically engineered to contain a run of His residues adheres
tightly to a microscope slide coated with a Ni complex; biotin is co-
valently attached to a c subunit of F
o
. The protein avidin, which binds
biotin very tightly, is covalently attached to long filaments of actin la-
beled with a fluorescent probe. Biotin-avidin binding now attaches
the actin filaments to the c subunit. When ATP is provided as sub-
strate for the ATPase activity of F
1
, the labeled filament is seen to ro-
tate continuously in one direction, proving that the F
o
cylinder of c
subunits rotates. In another experiment, a fluorescent actin filament
was attached directly to the H9253 subunit. The series of fluorescence mi-
crographs shows the position of the actin filament at intervals of
133 ms. Note that as the filament rotates, it makes a discrete jump
about every eleventh frame. Presumably the cylinder and shaft move
as one unit.
8885d_c19_690-750 3/1/04 11:32 AM Page 713 mac76 mac76:385_reb:
transmembrane electrochemical gradient, which gives
the matrix a net negative charge; the proton-motive
force drives ATP-ADP exchange. Adenine nucleotide
translocase is specifically inhibited by atractyloside, a
toxic glycoside formed by a species of thistle. If the
transport of ADP into and ATP out of mitochondria is
inhibited, cytosolic ATP cannot be regenerated from
ADP, explaining the toxicity of atractyloside.
A second membrane transport system essential to
oxidative phosphorylation is the phosphate translo-
case, which promotes symport of one H
2
PO
4
H11002
and one
H
H11001
into the matrix. This transport process, too, is fa-
vored by the transmembrane proton gradient (Fig.
19–26). Notice that the process requires movement of
one proton from the P to the N side of the inner mem-
brane, consuming some of the energy of electron trans-
fer. A complex of the ATP synthase and both translo-
cases, the ATP synthasome, can be isolated from
mitochondria by gentle dissection with detergents, sug-
gesting that the functions of these three proteins are
very tightly integrated.
Shuttle Systems Indirectly Convey Cytosolic NADH
into Mitochondria for Oxidation
The NADH dehydrogenase of the inner mitochondrial
membrane of animal cells can accept electrons only from
NADH in the matrix. Given that the inner membrane is
not permeable to NADH, how can the NADH generated
by glycolysis in the cytosol be reoxidized to NAD
H11001
by
O
2
via the respiratory chain? Special shuttle systems
carry reducing equivalents from cytosolic NADH into
mitochondria by an indirect route. The most active
NADH shuttle, which functions in liver, kidney, and
heart mitochondria, is the malate-aspartate shuttle
(Fig. 19–27). The reducing equivalents of cytosolic
NADH are first transferred to cytosolic oxaloacetate to
yield malate, catalyzed by cytosolic malate dehydroge-
nase. The malate thus formed passes through the inner
membrane via the malate–H9251-ketoglutarate transporter.
Within the matrix the reducing equivalents are passed
to NAD
H11001
by the action of matrix malate dehydrogenase,
forming NADH; this NADH can pass electrons directly
to the respiratory chain. About 2.5 molecules of ATP are
generated as this pair of electrons passes to O
2
. Cy-
tosolic oxaloacetate must be regenerated by transami-
nation reactions and the activity of membrane trans-
porters to start another cycle of the shuttle.
Skeletal muscle and brain use a different NADH
shuttle, the glycerol 3-phosphate shuttle (Fig.
19–28). It differs from the malate-aspartate shuttle in
that it delivers the reducing equivalents from NADH to
ubiquinone and thus into Complex III, not Complex I
(Fig. 19–8), providing only enough energy to synthesize
1.5 ATP molecules per pair of electrons.
The mitochondria of plants have an externally ori-
ented NADH dehydrogenase that can transfer electrons
directly from cytosolic NADH into the respiratory chain
at the level of ubiquinone. Because this pathway by-
passes the NADH dehydrogenase of Complex I and the
associated proton movement, the yield of ATP from cy-
tosolic NADH is less than that from NADH generated in
the matrix (Box 19–1).
SUMMARY 19.2 ATP Synthesis
■ The flow of electrons through Complexes I, III,
and IV results in pumping of protons across the
inner mitochondrial membrane, making the
matrix alkaline relative to the intermembrane
space. This proton gradient provides the energy
(in the form of the proton-motive force) for
ATP synthesis from ADP and P
i
by ATP synthase
(F
o
F
1
complex) in the inner membrane.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation714
Intermembrane
space
Matrix
Adenine
nucleotide
translocase
(antiporter)
ATP
synthase
Phosphate
translocase
(symporter)
ATP
4H11002
H11001
H11001
H11001
H11001
H11002
H11002
H11002
H11001
H11001
H11001
H11002
H11002
H11002
H11001
H11001
H11002
H11002
H11001
H11001
H11002
H11002
H11002
H
H11001
H
2
PO
4
H
H11001
ATP
4H11002
ADP
3H11002
H
H11001
H
2
PO
4
H
H11001
ADP
3H11002
H11002
H11002
FIGURE 19–26 Adenine nucleotide and phosphate translocases.
Transport systems of the inner mitochondrial membrane carry ADP
and P
i
into the matrix and newly synthesized ATP into the cytosol.
The adenine nucleotide translocase is an antiporter; the same protein
moves ADP into the matrix and ATP out. The effect of replacing ATP
4H11002
with ADP
3H11002
is the net efflux of one negative charge, which is favored
by the charge difference across the inner membrane (outside positive).
At pH 7, P
i
is present as both HPO
4
2H11002
and H
2
PO
4
H11002
; the phosphate
translocase is specific for H
2
PO
4
H11002
. There is no net flow of charge dur-
ing symport of H
2
PO
4
H11002
and H
H11001
, but the relatively low proton con-
centration in the matrix favors the inward movement of H
H11001
. Thus the
proton-motive force is responsible both for providing the energy for
ATP synthesis and for transporting substrates (ADP and P
i
) in and prod-
uct (ATP) out of the mitochondrial matrix. All three of these transport
systems can be isolated as a single membrane-bound complex (ATP
synthasome).
8885d_c19_690-750 3/1/04 11:32 AM Page 714 mac76 mac76:385_reb:
Intermembrane
space
Oxaloacetate
Aspartate
Oxaloacetate
malate
dehydrogenase
malate
dehydrogenase
aspartate
aminotransferase
aspartate
aminotransferase
Glutamate
H9251-Ketoglutarate
Glutamate
H9251-Ketoglutarate
Glutamate-aspartate
transporter
Malate–
H9251-ketoglutarate
transporter
2
3
NAD
+
NADH
4
5
Matrix
NAD
+
NADHH
+
+
6
1
Aspartate
Malate Malate
OH
C COO
H11002
CH
2
OOC
H11002
H
NH
3
H11001
NH
3
H11001
NH
3
H11001
NH
3
H11001
C COO
H11002
CH
2
CH
2
H
O
C COO
H11002
CH
2
CH
2
C COO
H11002
CH
2
H
C COO
H11002
CH
2
CH
2
H
C COO
H11002
CH
2
H
OH
C COO
H11002
CH
2
H
O
C COO
H11002
CH
2
CH
2
O
C COO
H11002
CH
2
OOC
H11002
O
C COO
H11002
CH
2
OOC
H11002
OOC
H11002
OOC
H11002
OOC
H11002
OOC
H11002
OOC
H11002 OOC
H11002
OOC
H11002
H
+
+
FIGURE 19–27 Malate-aspartate shuttle. This shuttle for transporting
reducing equivalents from cytosolic NADH into the mitochondrial ma-
trix is used in liver, kidney, and heart. 1 NADH in the cytosol (in-
termembrane space) passes two reducing equivalents to oxaloacetate,
producing malate. 2 Malate crosses the inner membrane via the
malate–H9251-ketoglutarate transporter. 3 In the matrix, malate passes
two reducing equivalents to NAD
H11001
, and the resulting NADH is oxi-
dized by the respiratory chain. The oxaloacetate formed from malate
cannot pass directly into the cytosol. 4 It is first transaminated to as-
partate, which 5 can leave via the glutamate-aspartate transporter.
6 Oxaloacetate is regenerated in the cytosol, completing the cycle.
19.2 ATP Synthesis 715
Q
Matrix
NADH
+
H
+
NAD
+
Glycolysis
Dihydroxyacetone
phosphate
Glycerol 3-
phosphate
FAD
FADH
2
III
CH
2
OH
CH
2
CHOH
O P
P
–
–
––
CH
2
OH
CH
2
C
O
–
–
–
–
–
–
O
mitochondrial
glycerol 3-phosphate
dehydrogenase
cytosolic
glycerol 3-phosphate
dehydrogenase
FIGURE 19–28 Glycerol 3-phosphate shuttle. This alternative
means of moving reducing equivalents from the cytosol to the
mitochondrial matrix operates in skeletal muscle and the
brain. In the cytosol, dihydroxyacetone phosphate accepts
two reducing equivalents from NADH in a reaction catalyzed
by cytosolic glycerol 3-phosphate dehydrogenase. An isozyme
of glycerol 3-phosphate dehydrogenase bound to the outer
face of the inner membrane then transfers two reducing
equivalents from glycerol 3-phosphate in the intermembrane
space to ubiquinone. Note that this shuttle does not involve
membrane transport systems.
8885d_c19_690-750 3/1/04 11:32 AM Page 715 mac76 mac76:385_reb:
■ ATP synthase carries out “rotational catalysis,”
in which the flow of protons through F
o
causes
each of three nucleotide-binding sites in F
1
to
cycle from (ADP H11001 P
i
)–bound to ATP-bound to
empty conformations.
■ ATP formation on the enzyme requires little
energy; the role of the proton-motive force is
to push ATP from its binding site on the
synthase.
■ The ratio of ATP synthesized per H5007
1
2
H5007O
2
reduced
to H
2
O (the P/O ratio) is about 2.5 when elec-
trons enter the respiratory chain at Complex I,
and 1.5 when electrons enter at CoQ.
■ Energy conserved in a proton gradient can
drive solute transport uphill across a membrane.
■ The inner mitochondrial membrane is
impermeable to NADH and NAD
H11001
, but NADH
equivalents are moved from the cytosol to the
matrix by either of two shuttles. NADH
equivalents moved in by the malate-aspartate
shuttle enter the respiratory chain at Complex
I and yield a P/O ratio of 2.5; those moved in
by the glycerol 3-phosphate shuttle enter at
CoQ and give a P/O ratio of 1.5.
19.3 Regulation of Oxidative
Phosphorylation
Oxidative phosphorylation produces most of the ATP
made in aerobic cells. Complete oxidation of a molecule
of glucose to CO
2
yields 30 or 32 ATP (Table 19–5).
By comparison, glycolysis under anaerobic conditions
(lactate fermentation) yields only 2 ATP per glucose.
Clearly, the evolution of oxidative phosphorylation pro-
vided a tremendous increase in the energy efficiency of
catabolism. Complete oxidation to CO
2
of the coenzyme
A derivative of palmitate (16:0), which also occurs in
the mitochondrial matrix, yields 108 ATP per palmitoyl-
CoA (see Table 17–1). A similar calculation can be made
for the ATP yield from oxidation of each of the amino
acids (Chapter 18). Aerobic oxidative pathways that
result in electron transfer to O
2
accompanied by oxida-
tive phosphorylation therefore account for the vast
majority of the ATP produced in catabolism, so the reg-
ulation of ATP production by oxidative phosphorylation
to match the cell’s fluctuating needs for ATP is ab-
solutely essential.
Oxidative Phosphorylation Is Regulated by Cellular
Energy Needs
The rate of respiration (O
2
consumption) in mitochon-
dria is tightly regulated; it is generally limited by the
availability of ADP as a substrate for phosphorylation.
Dependence of the rate of O
2
consumption on the avail-
ability of the P
i
acceptor ADP (Fig. 19–18b), the ac-
ceptor control of respiration, can be remarkable. In
some animal tissues, the acceptor control ratio, the
ratio of the maximal rate of ADP-induced O
2
consump-
tion to the basal rate in the absence of ADP, is at least
ten.
The intracellular concentration of ADP is one meas-
ure of the energy status of cells. Another, related meas-
ure is the mass-action ratio of the ATP-ADP system,
[ATP]/([ADP][P
i
]). Normally this ratio is very high, so
the ATP-ADP system is almost fully phosphorylated.
When the rate of some energy-requiring process (pro-
tein synthesis, for example) increases, the rate of break-
down of ATP to ADP and P
i
increases, lowering the
mass-action ratio. With more ADP available for oxida-
tive phosphorylation, the rate of respiration increases,
causing regeneration of ATP. This continues until the
mass-action ratio returns to its normal high level, at
which point respiration slows again. The rate of oxida-
tion of cellular fuels is regulated with such sensitivity
and precision that the [ATP]/([ADP][P
i
]) ratio fluctuates
only slightly in most tissues, even during extreme vari-
ations in energy demand. In short, ATP is formed only
as fast as it is used in energy-requiring cellular activities.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation716
Process Direct product Final ATP
Glycolysis 2 NADH (cytosolic) 3 or 5
*
2 ATP 2
Pyruvate oxidation (two per glucose) 2 NADH (mitochondrial matrix) 5
Acetyl-CoA oxidation in citric acid cycle 6 NADH (mitochondrial matrix) 15
(two per glucose) 2 FADH
2
3
2 ATP or 2 GTP 2
Total yield per glucose 30 or 32
TABLE 19–5 ATP Yield from Complete Oxidation of Glucose
*
The number depends on which shuttle system transfers reducing equivalents into the mitochondrion.
8885d_c19_690-750 3/1/04 11:32 AM Page 716 mac76 mac76:385_reb:
An Inhibitory Protein Prevents ATP Hydrolysis
during Ischemia
We have already encountered ATP synthase as an ATP-
driven proton pump (see Fig. 11–39; Table 11–3), cat-
alyzing the reverse of ATP synthesis. When a cell is is-
chemic (deprived of oxygen), as in a heart attack or
stroke, electron transfer to oxygen ceases, and so does
the pumping of protons. The proton-motive force soon
collapses. Under these conditions, the ATP synthase
could operate in reverse, hydrolyzing ATP to pump pro-
tons outward and causing a disastrous drop in ATP lev-
els. This is prevented by a small (84 amino acids) pro-
tein inhibitor, IF
1
, which simultaneously binds to two
ATP synthase molecules, inhibiting their ATPase activ-
ity (Fig. 19–29). IF
1
is inhibitory only in its dimeric form,
which is favored at pH lower than 6.5. In a cell starved
for oxygen, the main source of ATP becomes glycolysis,
and the pyruvic or lactic acid thus formed lowers the
pH in the cytosol and the mitochondrial matrix. This fa-
vors IF
1
dimerization, leading to inhibition of the ATPase
activity of ATP synthase, thereby preventing wasteful
hydrolysis of ATP. When aerobic metabolism resumes,
production of pyruvic acid slows, the pH of the cytosol
rises, the IF
1
dimer is destabilized, and the inhibition of
ATP synthase is lifted.
Uncoupled Mitochondria in Brown Fat Produce Heat
There is a remarkable and instructive exception to the
general rule that respiration slows when the ATP supply
is adequate. Most newborn mammals, including humans,
have a type of adipose tissue called brown fat in which
fuel oxidation serves not to produce ATP but to gener-
ate heat to keep the newborn warm. This specialized
adipose tissue is brown because of the presence of large
numbers of mitochondria and thus large amounts of
cytochromes, whose heme groups are strong absorbers
of visible light.
The mitochondria of brown fat are like those of other
mammalian cells in all respects, except that they have
a unique protein in their inner membrane. Thermo-
genin, also called the uncoupling protein (Table 19–4),
provides a path for protons to return to the matrix
without passing through the F
o
F
1
complex (Fig. 19–30).
19.3 Regulation of Oxidative Phosphorylation 717
FIGURE 19–29 Structure of bovine F
1
-ATPase in a complex
with its regulatory protein IF
1
. (Derived from PDB ID 1OHH)
Two F
1
molecules are viewed here as in Figure 19–23c. The
inhibitor IF
1
(red) binds to the H9251H9252 interface of the subunits in
the diphosphate (ADP) conformation (H9251ADP and H9252ADP),
freezing the two F
1
complexes and thereby blocking ATP
hydrolysis (and synthesis). (Parts of IF
1
that failed to resolve
in crystals of F
1
are shown in white outline as they occur in
crystals of isolated IF
1
.) This complex is stable only at the low
cytosolic pH characteristic of cells that are producing ATP by
glycolysis; when aerobic metabolism resumes, the cytosolic
pH rises, the inhibitor is destabilized, and ATP synthase
becomes active.
I
II
III
IV
Cyt c
Uncoupling
protein
(thermogenin)
ADP + P
i
ATP
H
+
H
+
H
+
F
o
F
1
Intermembrane
space
Matrix
Heat
FIGURE 19–30 Heat generation by uncoupled mitochondria. The un-
coupling protein (thermogenin) of brown fat mitochondria, by pro-
viding an alternative route for protons to reenter the mitochondrial
matrix, causes the energy conserved by proton pumping to be dissi-
pated as heat.
8885d_c19_690-750 3/1/04 11:32 AM Page 717 mac76 mac76:385_reb:
As a result of this short-circuiting of protons, the en-
ergy of oxidation is not conserved by ATP formation but
is dissipated as heat, which contributes to maintaining
the body temperature of the newborn. Hibernating an-
imals also depend on uncoupled mitochondria of brown
fat to generate heat during their long dormancy (see
Box 17–1).
ATP-Producing Pathways Are Coordinately Regulated
The major catabolic pathways have interlocking and
concerted regulatory mechanisms that allow them to
function together in an economical and self-regulating
manner to produce ATP and biosynthetic precursors.
The relative concentrations of ATP and ADP control not
only the rates of electron transfer and oxidative phos-
phorylation but also the rates of the citric acid cycle,
pyruvate oxidation, and glycolysis (Fig. 19–31). When-
ever ATP consumption increases, the rate of electron
transfer and oxidative phosphorylation increases. Si-
multaneously, the rate of pyruvate oxidation via the cit-
ric acid cycle increases, increasing the flow of electrons
into the respiratory chain. These events can in turn
evoke an increase in the rate of glycolysis, increasing
the rate of pyruvate formation. When conversion of ADP
to ATP lowers the ADP concentration, acceptor control
slows electron transfer and thus oxidative phosphoryla-
tion. Glycolysis and the citric acid cycle are also slowed,
because ATP is an allosteric inhibitor of the glycolytic
enzyme phosphofructokinase-1 (see Fig. 15–18) and of
pyruvate dehydrogenase (see Fig. 16–18).
Phosphofructokinase-1 is also inhibited by citrate,
the first intermediate of the citric acid cycle. When the
cycle is “idling,” citrate accumulates within mitochon-
dria, then spills into the cytosol. When the concentra-
tions of both ATP and citrate rise, they produce a con-
certed allosteric inhibition of phosphofructokinase-1
that is greater than the sum of their individual effects,
slowing glycolysis.
SUMMARY 19.3 Regulation of Oxidative
Phosphorylation
■ Oxidative phosphorylation is regulated by
cellular energy demands. The intracellular [ADP]
and the mass-action ratio [ATP]/([ADP][P
i
]) are
measures of a cell’s energy status.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation718
Glucose
P
i
Glycolysis
Glucose 6-phosphate
Fructose 1,6-bisphosphate
AMP, ADP
ATP, citrate
hexokinase
phosphofructokinase-1
multistep
Phosphoenolpyruvate
ADP
ATP, NADH
Pyruvate
AMP, ADP, NAD
H11001
ATP, NADH
Acetyl-CoA
ADP
ATP, NADH
Citrate
ADP
ATP
pyruvate kinase
H9251-Ketoglutarate
ATP, NADH
Succinyl-CoA
Oxaloacetate
pyruvate
dehydrogenase
complex
citrate synthase
ADP, P
i
Respiratory chain
H
2
O
ADP H11001 P
i
ATP
O
2
1
2
NADH
NAD
H11001
isocitrate
dehydrogenase
multistep
-ketoglutarate
dehydrogenase
H9251
Citric
acid
cycle
Oxidative
phosphory-
lation
FIGURE 19–31 Regulation of the ATP-producing pathways. This di-
agram shows the interlocking regulation of glycolysis, pyruvate oxi-
dation, the citric acid cycle, and oxidative phosphorylation by the rel-
ative concentrations of ATP, ADP, and AMP, and by NADH. High [ATP]
(or low [ADP] and [AMP]) produces low rates of glycolysis, pyruvate
oxidation, acetate oxidation via the citric acid cycle, and oxidative
phosphorylation. All four pathways are accelerated when the use of
ATP and the formation of ADP, AMP, and P
i
increase. The interlock-
ing of glycolysis and the citric acid cycle by citrate, which inhibits
glycolysis, supplements the action of the adenine nucleotide system.
In addition, increased levels of NADH and acetyl-CoA also inhibit the
oxidation of pyruvate to acetyl-CoA, and a high [NADH]/[NAD
H11001
] ra-
tio inhibits the dehydrogenase reactions of the citric acid cycle (see
Fig. 16–18).
8885d_c19_690-750 3/1/04 11:32 AM Page 718 mac76 mac76:385_reb:
■ In ischemic (oxygen-deprived) cells, a protein
inhibitor blocks ATP hydrolysis by the ATP
synthase operating in reverse, preventing a
drastic drop in [ATP].
■ In brown fat, which is specialized for the
production of metabolic heat, electron transfer
is uncoupled from ATP synthesis and the
energy of fatty acid oxidation is dissipated as
heat.
■ ATP and ADP concentrations set the rate
of electron transfer through the respiratory
chain via a series of interlocking controls on
respiration, glycolysis, and the citric acid cycle.
19.4 Mitochondrial Genes: Their Origin and
the Effects of Mutations
Mitochondria contain their own genome, a circular,
double-stranded DNA molecule. Each of the hundreds
or thousands of mitochondria in a typical cell has about
five copies of this genome. The human mitochondrial
chromosome (Fig. 19–32) contains 37 genes (16,569 bp),
including 13 that encode subunits of proteins of the
respiratory chain (Table 19–6); the remaining genes
code for rRNA and tRNA molecules essential to the
protein-synthesizing machinery of mitochondria. About
900 different mitochondrial proteins are encoded by nu-
clear genes, synthesized on cytoplasmic ribosomes, then
imported and assembled within the mitochondria
(Chapter 27).
Mutations in Mitochondrial Genes Cause
Human Disease
A growing number of human diseases can be at-
tributed to mutations in mitochondrial genes.
Many of these diseases, those known as the mitochon-
drial encephalomyopathies, affect primarily the brain
and skeletal muscle (both heavily dependent on an
abundant supply of ATP). These diseases are invariably
inherited from the mother, because a developing em-
bryo derives all its mitochondria from the mother’s egg.
The rare disease Leber’s hereditary optic neuropa-
thy (LHON) affects the central nervous system, in-
cluding the optic nerves, causing bilateral loss of vision
in early adulthood. A single base change in the mito-
chondrial gene ND4 (Fig. 19–32a) changes an Arg
residue to a His residue in a polypeptide of Complex I,
and the result is mitochondria partially defective in elec-
tron transfer from NADH to ubiquinone. Although these
mitochondria can produce some ATP by electron trans-
fer from succinate, they apparently cannot supply suf-
ficient ATP to support the very active metabolism of
neurons. One result is damage to the optic nerve, lead-
ing to blindness. A single base change in the mitochon-
drial gene for cytochrome b, a component of Complex
III, also produces LHON, demonstrating that the pathol-
ogy results from a general reduction of mitochondrial
function, not specifically from a defect in electron trans-
fer through Complex I.
Myoclonic epilepsy and ragged-red fiber dis-
ease (MERRF) is caused by a mutation in the mito-
chondrial gene that encodes a transfer RNA specific for
lycine (lysyl-tRNA). This disease, characterized by un-
controllable muscular jerking, apparently results from
defective production of several of the proteins whose
synthesis involves mitochondrial tRNAs. Skeletal mus-
cle fibers of individuals with MERRF have abnormally
shaped mitochondria that sometimes contain paracrys-
talline structures (Fig. 19–32b). Mutations in the mito-
chondrial lysyl-tRNA gene are also one of the causes of
adult-onset (type II) diabetes mellitus. Other mutations
in mitochondrial genes are believed to be responsible
for the progressive muscular weakness that character-
izes mitochondrial myopathy and for enlargement and
deterioration of the heart muscle in hypertrophic cardio-
myopathy. According to one hypothesis on the progres-
sive changes that accompany aging, the accumulation
of mutations in mitochondrial DNA during a lifetime of
exposure to DNA-damaging agents such as
H11080
O
2
H11002
(see
below) results in mitochondria that cannot supply suf-
ficient ATP for normal cellular function. Mitochondrial
disease can also result from mutations in any of the 900
nuclear genes that encode mitochondrial proteins. ■
19.4 Mitochondrial Genes: Their Origin and the Effects of Mutations 719
Number Number of subunits encoded
Complex of subunits by mitochondrial DNA
I NADH dehydrogenase H1102243 7
II Succinate dehydrogenase 4 0
III Ubiquinone:cytochrome c oxidoreductase 11 1
IV Cytochrome oxidase 13 3
V ATP synthase 8 2
TABLE 19–6 Respiratory Proteins Encoded by Mitochondrial Genes in Humans
8885d_c19_690-750 3/1/04 11:32 AM Page 719 mac76 mac76:385_reb:
Chapter 19 Oxidative Phosphorylation and Photophosphorylation720
Menaquinone
n H11005 7–9
CH
3
(CH
2
CH CH
2
)
n
C
CH
3
H
O
O
(b)
Cytosol (N side)
Bacterial inner
(plasma)
membrane
Periplasmic
space (P side)
Cyt b
Cyt o
Cu
2Fe-2S
FAD
NADH, succinate,
or glycerol
4Fe-4S
Q
O
2
H
+
H
+
(a)
FIGURE 19–33 Bacterial respiratory chain. (a) Shown here are the
respiratory carriers of the inner membrane of E. coli. Eubacteria con-
tain a minimal form of Complex I, containing all the prosthetic groups
normally associated with the mitochondrial complex but only 14
polypeptides. This plasma membrane complex transfers electrons from
NADH to ubiquinone or to (b) menaquinone, the bacterial equivalent
of ubiquinone, while pumping protons outward and creating an elec-
trochemical potential that drives ATP synthesis.
FIGURE 19–32 Mitochondrial genes and mutations.
(a) Map of human mitochondrial DNA, showing the genes
that encode proteins of Complex I, the NADH dehydroge-
nase (ND1 to ND6); the cytochrome b of Complex III (Cyt
b); the subunits of cytochrome oxidase (Complex IV) (COI to
COIII); and two subunits of ATP synthase (ATPase6 and ATPase8).
The colors of the genes correspond to those of the complexes shown
in Figure 19–7. Also included here are the genes for ribosomal RNAs
(rRNA) and for a number of mitochondrion-specific transfer RNAs;
tRNA specificity is indicated by the one-letter codes for amino acids.
Arrows indicate the positions of mutations that cause Leber’s heredi-
tary optic neuropathy (LHON) and myoclonic epilepsy and ragged-
red fiber disease (MERRF). Numbers in parentheses indicate the posi-
tion of the altered nucleotides (nucleotide 1 is at the top of the circle
and numbering proceeds counterclockwise). (b) Electron micrograph
of an abnormal mitochondrion from the muscle of an individual with
MERRF, showing the paracrystalline protein inclusions sometimes pres-
ent in the mutant mitochondria.
0/16,569
12S
rRNA
16S
rRNA
ND1
ND2
COI
COII
COIII
ATPase6
ND3
ND4L
ND4
ND5
ND6
Cyt b
F
V
L
I
M
W
A
N
C
Y
S
D
K
G
(a)
R
H
L
S
E
P
T
LHON
(15,257)
Q
LHON
(3,460)
LHON
(4,160)
LHON
(11,778)
MERRF
(8,344)
ATPase8
Complex I
Complex III
Complex IV
ATP synthase
Transfer RNA
Ribosomal RNA
Control region of DNA
(b)
8885d_c19_690-750 3/1/04 11:32 AM Page 720 mac76 mac76:385_reb:
Mitochondria Evolved from Endosymbiotic Bacteria
The existence of mitochondrial DNA, ribosomes, and
tRNAs supports the hypothesis of the endosymbiotic
origin of mitochondria (see Fig. 1–36), which holds that
the first organisms capable of aerobic metabolism, in-
cluding respiration-linked ATP production, were prokar-
yotes. Primitive eukaryotes that lived anaerobically (by
fermentation) acquired the ability to carry out oxidative
phosphorylation when they established a symbiotic re-
lationship with bacteria living in their cytosol. After much
evolution and the movement of many bacterial genes into
the nucleus of the “host” eukaryote, the endosymbiotic
bacteria eventually became mitochondria.
This hypothesis presumes that early free-living
prokaryotes had the enzymatic machinery for oxidative
phosphorylation and predicts that their modern
prokaryotic descendants must have respiratory chains
closely similar to those of modern eukaryotes. They do.
Aerobic bacteria carry out NAD-linked electron trans-
fer from substrates to O
2
, coupled to the phosphoryla-
tion of cytosolic ADP. The dehydrogenases are located
in the bacterial cytosol and the respiratory chain in the
plasma membrane. The electron carriers are similar to
some mitochondrial electron carriers (Fig. 19–33). They
translocate protons outward across the plasma mem-
brane as electrons are transferred to O
2
. Bacteria such
as Escherichia coli have F
o
F
1
complexes in their
plasma membranes; the F
1
portion protrudes into the
cytosol and catalyzes ATP synthesis from ADP and P
i
as protons flow back into the cell through the proton
channel of F
o
.
The respiration-linked extrusion of protons across
the bacterial plasma membrane also provides the driv-
ing force for other processes. Certain bacterial trans-
port systems bring about uptake of extracellular nutri-
ents (lactose, for example) against a concentration
gradient, in symport with protons (see Fig. 11–42). And
the rotary motion of bacterial flagella is provided by
“proton turbines,” molecular rotary motors driven not
by ATP but directly by the transmembrane electro-
chemical potential generated by respiration-linked pro-
ton pumping (Fig. 19–34). It appears likely that the
chemiosmotic mechanism evolved early, before the
emergence of eukaryotes.
SUMMARY 19.4 Mitochondrial Genes: Their Origin
and the Effects of Mutations
■ A small proportion of human mitochondrial
proteins (13 proteins) are encoded in the
mitochondrial genome and synthesized within
mitochondria. About 900 mitochondrial
proteins are encoded in nuclear genes and
imported into mitochondria after their
synthesis.
■ Mutations in the genes that encode
components of the respiratory chain, whether
in the mitochondrial genes or in the nuclear
genes that encode mitochondrial proteins,
cause a variety of human diseases, which often
affect muscle and brain most severely.
■ Mitochondria most likely arose from aerobic
prokaryotes that entered into an endosymbiotic
relationship with ancestral eukaryotes.
19.5 The Role of Mitochondria in Apoptosis
and Oxidative Stress
Besides their central role in ATP synthesis, mitochon-
dria also participate in processes associated with cellu-
lar damage and death. Apoptosis is a controlled process
by which cells die for the good of the organism, while
the organism conserves the molecular components
(amino acids, nucleotides, and so forth) of the dead
cells. Apoptosis may be triggered by an external signal,
acting at a receptor in the plasma membrane, or by in-
ternal events such as a viral infection. When a cell re-
ceives a signal for apoptosis, one consequence is an in-
crease in the permeability of the outer mitochondrial
membrane, allowing escape of the cytochrome c nor-
mally confined in the intermembrane space (see Fig.
12–50). The released cytochrome c activates one of the
proteolytic enzymes (caspase 9) responsible for protein
degradation during apoptosis. This is a dramatic case of
19.5 The Role of Mitochondria in Apoptosis and Oxidative Stress 721
Flagellum
Outer membrane
Inner (plasma)
membrane
Rotary motor
Peptidoglycan
and periplasmic
space
Electron-transfer
chain
H
H11001
H
H11001
FIGURE 19–34 Rotation of bacterial flagella by proton-motive force.
The shaft and rings at the base of the flagellum make up a rotary mo-
tor that has been called a “proton turbine.” Protons ejected by elec-
tron transfer flow back into the cell through the turbine, causing ro-
tation of the shaft of the flagellum. This motion differs fundamentally
from the motion of muscle and of eukaryotic flagella and cilia, for
which ATP hydrolysis is the energy source.
8885d_c19_690-750 3/1/04 11:32 AM Page 721 mac76 mac76:385_reb:
one protein (cytochrome c) playing two very different
roles in the cell.
Mitochondria are also involved in the cell’s response
to oxidative stress. As we have seen, several steps in
the path of oxygen reduction in mitochondria have the
potential to produce highly reactive free radicals that
can damage cells. The passage of electrons from QH
2
to
cytochrome b
L
through Complex III, and passage of elec-
trons from Complex I to QH
2
, involve the radical
H11080
Q
H11002
as
an intermediate. The
H11080
Q
H11002
can, with a low probability,
pass an electron to O
2
in the reaction
O
2
H11001 e
H11002
On
H11080
O
2
H11002
The superoxide free radical thus generated,
H11080
O
2
H11002
, is very
reactive and can damage enzymes, membrane lipids,
and nucleic acids. Antimycin A, an inhibitor of Complex
III, may act by occupying the Q
N
site (Fig. 19–11), thus
blocking the Q cycle and prolonging the binding of
H11080
Q
H11002
to the Q
P
site; this would increase the likelihood of su-
peroxide radical formation and cellular damage. From
0.1% to as much as 4% of the O
2
used by actively respir-
ing mitochondria forms
H11080
O
2
H11002
—more than enough to have
lethal effects on a cell unless the free radical is quickly
disposed of.
To prevent oxidative damage by
H11080
O
2
H11002
, cells have sev-
eral forms of the enzyme superoxide dismutase,
which catalyzes the reaction
2
H11080
O
2
H11002
H11001 2H
H11001
88n H
2
O
2
H11001 O
2
The hydrogen peroxide (H
2
O
2
) generated by this reac-
tion is rendered harmless by the action of glutathione
peroxidase (Fig. 19–35). This enzyme is remarkable for
the presence of a selenocysteine residue (see Fig. 3–8a),
in which an atom of selenium replaces the sulfur atom
normally present in the thiol of the side chain. The se-
lenol group (OSeH) is more acidic than the thiol (OSH);
its pKa is about 5, so at neutral pH, the selenocysteine
side chain is essentially fully ionized (OCH
2
Se
H11002
). Gluta-
thione reductase recycles oxidized glutathione to its re-
duced form, using electrons from the NADPH formed
by nicotinamide nucleotide transhydrogenase or by the
pentose phosphate pathway (see Fig. 14–20). Reduced
glutathione also serves in keeping protein sulfhydryl
groups in their reduced state, preventing some of the
deleterious effects of oxidative stress (Fig. 19–35).
SUMMARY 19.5 The Role of Mitochondria in
Apoptosis and Oxidative Stress
■ Mitochondrial cytochrome c, released into the
cytosol, participates in activation of one of the
proteases (caspase 9) involved in apoptosis.
■ Reactive oxygen species produced in
mitochondria are inactivated by a set of
protective enzymes, including superoxide
dismutase and glutathione peroxidase.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation722
Nicotinamide
nucleotide
transhydrogenase
NADH
NADPH
GSSG
H
2
O
2
H
2
O
2 GSH
S
S
2 GSH
Enz
inactive
active
oxidative
stress
protein thiol
reduction
GSSG
NADP
+
glutathione
reductase
glutathione
peroxidase
superoxide
dismutase
NAD
+
O
2
O
2
NAD
+
Inner
mitochondrial
membrane
Q
III
IV
I
Cyt c
.
–
SH
SH
FIGURE 19–35 Mitochondrial production and disposal of super-
oxide. Superoxide radical, ?O
2
H11002
, is formed in side reactions at
Complexes I and III, as the partially reduced ubiquinone radical (?Q
H11002
)
donates an electron to O
2
. The reactions shown in blue defend the
cell against the damaging effects of superoxide. Reduced glutathione
(GSH; see Fig. 22–27) donates electrons for the reduction of hydrogen
peroxide (H
2
O
2
) and of oxidized Cys residues (OSOSO) in proteins,
and GSH is regenerated from the oxidized form (GSSG) by reduction
with NADPH.
8885d_c19_690-750 3/1/04 11:32 AM Page 722 mac76 mac76:385_reb:
FIGURE 19–37 The light reactions
of photosynthesis generate energy-
rich NADPH and ATP at the
expense of solar energy. These
products are used in the carbon-
assimilation reactions, which occur
in light or darkness, to reduce CO
2
to form trioses and more complex
compounds (such as glucose)
derived from trioses.
Light
reactions
NADP
+
Carbon-assimilation
reactions
ADP + P
i
NADPH
ATP
H
2
O O
2
CO
2
Carbohydrate
Photosynthetic
cells
CarbohydrateO
2
CO
2
H
2
O
Heterotrophic
cells
FIGURE 19–36 Solar
energy as the ultimate
source of all biological
energy. Photosynthetic
organisms use the energy
of sunlight to manufacture
glucose and other organic
products, which hetero-
trophic cells use as energy
and carbon sources.
PHOTOSYNTHESIS:
HARVESTING LIGHT ENERGY
We now turn to another reaction sequence in which the
flow of electrons is coupled to the synthesis of ATP:
light-driven phosphorylation. The capture of solar en-
ergy by photosynthetic organisms and its conversion to
the chemical energy of reduced organic compounds is
the ultimate source of nearly all biological energy. Pho-
tosynthetic and heterotrophic organisms live in a bal-
anced steady state in the biosphere (Fig. 19–36). Pho-
tosynthetic organisms trap solar energy and form ATP
and NADPH, which they use as energy sources to make
carbohydrates and other organic compounds from CO
2
and H
2
O; simultaneously, they release O
2
into the at-
mosphere. Aerobic heterotrophs (humans, for example,
as well as plants during dark periods) use the O
2
so
formed to degrade the energy-rich organic products of
photosynthesis to CO
2
and H
2
O, generating ATP. The
CO
2
returns to the atmosphere, to be used again by pho-
tosynthetic organisms. Solar energy thus provides the
driving force for the continuous cycling of CO
2
and O
2
through the biosphere and provides the reduced
substrates—fuels, such as glucose—on which nonpho-
tosynthetic organisms depend.
Photosynthesis occurs in a variety of bacteria and
in unicellular eukaryotes (algae) as well as in vascular
plants. Although the process in these organisms differs
in detail, the underlying mechanisms are remarkably
similar, and much of our understanding of photosyn-
thesis in vascular plants is derived from studies of sim-
pler organisms. The overall equation for photosynthesis
in vascular plants describes an oxidation-reduction re-
action in which H
2
O donates electrons (as hydrogen)
for the reduction of CO
2
to carbohydrate (CH
2
O):
light
CO
2
H11001 H
2
O 888n O
2
H11001 (CH
2
O)
19.6 General Features
of Photophosphorylation
Unlike NADH (the major electron donor in oxidative
phosphorylation), H
2
O is a poor donor of electrons; its
standard reduction potential is 0.816 V, compared with
H110020.320 V for NADH. Photophosphorylation differs from
oxidative phosphorylation in requiring the input of en-
ergy in the form of light to create a good electron donor
and a good electron acceptor. In photophosphorylation,
electrons flow through a series of membrane-bound car-
riers including cytochromes, quinones, and iron-sulfur
proteins, while protons are pumped across a membrane
to create an electrochemical potential. Electron trans-
fer and proton pumping are catalyzed by membrane
complexes homologous in structure and function to
Complex III of mitochondria. The electrochemical po-
tential they produce is the driving force for ATP syn-
thesis from ADP and P
i
, catalyzed by a membrane-bound
ATP synthase complex closely similar to that of oxida-
tive phosphorylation.
Photosynthesis in plants encompasses two pro-
cesses: the light-dependent reactions, or light re-
actions, which occur only when plants are illuminated,
and the carbon-assimilation reactions (or carbon-
fixation reactions), sometimes misleadingly called the
dark reactions, which are driven by products of the light
reactions (Fig. 19–37). In the light reactions, chlorophyll
and other pigments of photosynthetic cells absorb light
energy and conserve it as ATP and NADPH; simultane-
ously, O
2
is evolved. In the carbon-assimilation reac-
tions, ATP and NADPH are used to reduce CO
2
to form
triose phosphates, starch, and sucrose, and other prod-
ucts derived from them. In this chapter we are con-
cerned only with the light-dependent reactions that lead
to the synthesis of ATP and NADPH. The reduction of
CO
2
is described in Chapter 20.
19.6 General Features of Photophosphorylation 723
8885d_c19_690-750 3/1/04 11:32 AM Page 723 mac76 mac76:385_reb:
Photosynthesis in Plants Takes Place in Chloroplasts
In photosynthetic eukaryotic cells, both the light-de-
pendent and the carbon-assimilation reactions take
place in the chloroplasts (Fig. 19–38), membrane-
bounded intracellular organelles that are variable in
shape and generally a few micrometers in diameter. Like
mitochondria, they are surrounded by two membranes,
an outer membrane that is permeable to small molecules
and ions, and an inner membrane that encloses the in-
ternal compartment. This compartment contains many
flattened, membrane-surrounded vesicles or sacs, the
thylakoids, usually arranged in stacks called grana
(Fig. 19–38b). Embedded in the thylakoid membranes
(commonly called lamellae) are the photosynthetic
pigments and the enzyme complexes that carry out the
light reactions and ATP synthesis. The stroma (the
aqueous phase enclosed by the inner membrane) con-
tains most of the enzymes required for the carbon-
assimilation reactions.
Light Drives Electron Flow in Chloroplasts
In 1937 Robert Hill found that when leaf extracts con-
taining chloroplasts were illuminated, they (1) evolved
O
2
and (2) reduced a nonbiological electron acceptor
added to the medium, according to the Hill reaction:
light
2H
2
O H11001 2A 888n 2AH
2
H11001 O
2
where A is the artificial electron acceptor, or Hill
reagent. One Hill reagent, the dye 2,6-dichlorophenol-
indophenol, is blue when oxidized (A) and colorless
when reduced (AH
2
), making the reaction easy to fol-
low. When a leaf extract supplemented with the dye was
illuminated, the blue dye became colorless and O
2
was
evolved. In the dark, neither O
2
evolution nor dye re-
duction took place. This was the first evidence that
absorbed light energy causes electrons to flow from H
2
O
to an electron acceptor. Moreover, Hill found that CO
2
was neither required nor reduced to a stable form un-
der these conditions; O
2
evolution could be dissociated
from CO
2
reduction. Several years later Severo Ochoa
showed that NADP
H11001
is the biological electron acceptor
in chloroplasts, according to the equation
light
2H
2
O H11001 2NADP
H11001
888n 2NADPH H11001 2H
H11001
H11001 O
2
To understand this photochemical process, we must first
consider the more general topic of the effects of light
absorption on molecular structure.
SUMMARY 19.6 General Features
of Photophosphorylation
■ The light reactions of photosynthesis are those
directly dependent on the absorption of light;
the resulting photochemistry takes electrons
from H
2
O and drives them through a series of
membrane-bound carriers, producing NADPH
and ATP.
■ The carbon-assimilation reactions of
photosynthesis reduce CO
2
with electrons from
NADPH and energy from ATP.
OH
Reduced form
(colorless)
Oxidized form
(blue)
Dichlorophenolindophenol
OH
O OH
N
Cl ClCl Cl
NH
Chapter 19 Oxidative Phosphorylation and Photophosphorylation724
Grana (thylakoids)
Stroma
(b)
Outer membrane
Inner membrane
Thylakoids
(a)
FIGURE 19–38 Chloroplast. (a) Schematic diagram. (b) Electron mi-
crograph at high magnification showing grana, stacks of thylakoid
membranes.
8885d_c19_690-750 3/1/04 11:32 AM Page 724 mac76 mac76:385_reb:
19.7 Light Absorption
Visible light is electromagnetic radiation of wavelengths
400 to 700 nm, a small part of the electromagnetic spec-
trum (Fig. 19–39), ranging from violet to red. The en-
ergy of a single photon (a quantum of light) is greater
at the violet end of the spectrum than at the red end;
shorter wavelength (and higher frequency) corresponds
to higher energy. The energy, E, in a “mole” of photons
(1 einstein, or 6 H11003 10
23
photons) of visible light is 170
to 300 kJ, as given by the Planck equation:
E H11005 hH9263
where h is Planck’s constant (6.626 H11003 10
H1100234
J H11554 s) and H9263
is the wavelength. These amounts of energy are almost
an order of magnitude greater than the 30 to 50 kJ re-
quired to synthesize a mole of ATP from ADP and P
i
.
When a photon is absorbed, an electron in the ab-
sorbing molecule (chromophore) is lifted to a higher
energy level. This is an all-or-nothing event; to be ab-
sorbed, the photon must contain a quantity of energy (a
quantum) that exactly matches the energy of the elec-
tronic transition. A molecule that has absorbed a photon
is in an excited state, which is generally unstable. An
electron lifted into a higher-energy orbital usually re-
turns rapidly to its normal lower-energy orbital; the ex-
cited molecule decays to the stable ground state,
giving up the absorbed quantum as light or heat or us-
ing it to do chemical work. Light emission accompany-
ing decay of excited molecules (called fluorescence)
is always at a longer wavelength (lower energy) than
that of the absorbed light (see Box 12–2). An alterna-
tive mode of decay important in photosynthesis involves
direct transfer of excitation energy from an excited mol-
ecule to a neighboring molecule. Just as the photon is
a quantum of light energy, so the exciton is a quantum
of energy passed from an excited molecule to another
molecule in a process called exciton transfer.
Chlorophylls Absorb Light Energy for Photosynthesis
The most important light-absorbing pigments in the thy-
lakoid membranes are the chlorophylls, green pigments
with polycyclic, planar structures resembling the proto-
porphyrin of hemoglobin (see Fig. 5–1), except that Mg
2H11001
,
not Fe
2H11001
, occupies the central position (Fig. 19–40). The
four inward-oriented nitrogen atoms of chlorophyll are
coordinated with the Mg
2H11001
. All chlorophylls have a long
phytol side chain, esterified to a carboxyl-group sub-
stituent in ring IV, and chlorophylls also have a fifth five-
membered ring not present in heme.
The heterocyclic five-ring system that surrounds
the Mg
2H11001
has an extended polyene structure, with al-
ternating single and double bonds. Such polyenes char-
acteristically show strong absorption in the visible re-
gion of the spectrum (Fig. 19–41); the chlorophylls have
unusually high molar extinction coefficients (see Box
3–1) and are therefore particularly well-suited for ab-
sorbing visible light during photosynthesis.
Chloroplasts always contain both chlorophyll a and
chlorophyll b (Fig. 19–40a). Although both are green,
their absorption spectra are sufficiently different (Fig.
19–41) that they complement each other’s range of
light absorption in the visible region. Most plants con-
tain about twice as much chlorophyll a as chlorophyll
b. The pigments in algae and photosynthetic bacteria
include chlorophylls that differ only slightly from the
plant pigments.
Chlorophyll is always associated with specific
binding proteins, forming light-harvesting com-
plexes (LHCs) in which chlorophyll molecules are
fixed in relation to each other, to other protein com-
plexes, and to the membrane. The detailed structure
of one light-harvesting complex is known from x-ray
crystallography (Fig. 19–42). It contains seven mole-
cules of chlorophyll a, five of chlorophyll b, and two
of the accessory pigment lutein (see below).
19.7 Light Absorption 725
380
CyanBlueViolet Green
Yellow
Orange Red
Wavelength
(nm)
Energy
(kJ/einstein)
300
430 500 560 600 650 750
240 200 170
Wavelength
Type of
radiation
Gamma rays
Visible light
X rays UV Infrared Microwaves Radio waves
Thousands of metersH110211 millimeter 1 meterH110211 nm 100 nm
FIGURE 19–39 Electromagnetic radiation. The spectrum of electromagnetic radiation, and the energy
of photons in the visible range of the spectrum. One einstein is 6 H11003 10
23
photons.
8885d_c19_725 3/1/04 1:59 PM Page 725 mac76 mac76:385_reb:
Cyanobacteria and red algae employ phycobilins
such as phycoerythrobilin and phycocyanobilin (Fig.
19–40b) as their light-harvesting pigments. These open-
chain tetrapyrroles have the extended polyene system
found in chlorophylls, but not their cyclic structure or
central Mg
2H11001
. Phycobilins are covalently linked to spe-
cific binding proteins, forming phycobiliproteins,
which associate in highly ordered complexes called phy-
cobilisomes (Fig. 19–43) that constitute the primary
light-harvesting structures in these microorganisms.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation726
A
CH
2
N
Mg
H
O
C
III
IIIIV
G D
B
M
0
;
;
H
H
CH
2
CH
3
CH
2
CH
3
CH CH
3
N
NN
D
CH
3
O
CH
3
O
CH
2
B
D
O
C
O
CH
2
CH
3
CH
3
CH
3
CH
3
G
CH
3
phytol side chain
b-Carotene
O
C
CH
3
in bacteriochlorophyll
CHO in chlorophyll b
Saturated bond in
bacteriochlorophyll
A
GJ
CH
3
CH
3
CH
3
H
3
CCH
3
CH
3
CH
3
CH
3
CH
3
CH
3
Phycoerythrobilin
A
A
A
A
COO
H11002
N
CH
3
CH
3
CH
2
COO
H11002
CH
2
CH
2
CH
3
CH
2
N
H
N
H
CH
3
CH
O
A
CH
3
N
H
CH
3
CH
O
Unsaturated bond
in phycocyanobilin
in phycocyanobilinB
CH
2
CH
3
Chlorophyll a
G
G
(a)
(b)
(c)
Lutein (xanthophyll)
CH
OH
3
HC
3
HC
3
H
3
C
CH
3
CH
3
CH
3
CHHO
3
CH
3
CH
3
(d)
FIGURE 19–40 Primary and secondary photopigments. (a) Chloro-
phylls a and b and bacteriochlorophyll are the primary gatherers of
light energy. (b) Phycoerythrobilin and phycocyanobilin (phycobilins)
are the antenna pigments in cyanobacteria and red algae. (c) H9252-
Carotene (a carotenoid) and (d) lutein (a xanthophyll) are accessory
pigments in plants. The areas shaded pink are the conjugated systems
(alternating single and double bonds) that largely account for the ab-
sorption of visible light.
8885d_c19_690-750 3/1/04 11:32 AM Page 726 mac76 mac76:385_reb:
19.7 Light Absorption 727
Absorption
300
Sunlight
reaching
the earth
Wavelength (nm)
400 500 600 700
Chlorophyll b
-CaroteneH9252
Phycocyanin
Chlorophyll a
Phycoerythrin
Lutein
800
FIGURE 19–41 Absorption of visible light by photopigments. Plants
are green because their pigments absorb light from the red and blue
regions of the spectrum, leaving primarily green light to be reflected
or transmitted. Compare the absorption spectra of the pigments with
the spectrum of sunlight reaching the earth’s surface; the combination
of chlorophylls (a and b) and accessory pigments enables plants to
harvest most of the energy available in sunlight.
The relative amounts of chlorophylls and accessory pigments are
characteristic of a particular plant species. Variation in the proportions
of these pigments is responsible for the range of colors of photosyn-
thetic organisms, from the deep blue-green of spruce needles, to the
greener green of maple leaves, to the red, brown, or purple color of
some species of multicellular algae and the leaves of some foliage
plants favored by gardeners.
FIGURE 19–42 A light-harvesting complex, LHCII. The functional
unit is an LHC trimer, with 36 chlorophyll and 6 lutein molecules.
Shown here is a monomer, viewed in the plane of the membrane, with
its three transmembrane H9251-helical segments, seven chlorophyll a
molecules (green), five chlorophyll b molecules (red), and two mole-
cules of the accessory pigment lutein (yellow), which form an internal
cross-brace.
FIGURE 19–43 A phycobilisome. In these highly structured assem-
blies found in cyanobacteria and red algae, phycobilin pigments
bound to specific proteins form complexes called phycoerythrin (PE),
phycocyanin (PC), and allophycocyanin (AP). The energy of photons
absorbed by PE or PC is conveyed through AP (a phycocyanobilin-
binding protein) to chlorophyll a of the reaction center by exciton
transfer, a process discussed in the text.
PE
PE
PC
550–650 nm
PC
AP
AP
Thylakoid
membrane
Exciton
transfer
Chlorophyll a
reaction center
480–570 nm
Light
8885d_c19_690-750 3/1/04 11:32 AM Page 727 mac76 mac76:385_reb:
Accessory Pigments Extend the Range
of Light Absorption
In addition to chlorophylls, thylakoid membranes con-
tain secondary light-absorbing pigments, or accessory
pigments, called carotenoids. Carotenoids may be yel-
low, red, or purple. The most important are H9252-carotene,
which is a red-orange isoprenoid, and the yellow
carotenoid lutein (Fig. 19–40c, d). The carotenoid pig-
ments absorb light at wavelengths not absorbed by the
chlorophylls (Fig. 19–41) and thus are supplementary
light receptors.
Experimental determination of the effectiveness of
light of different colors in promoting photosynthesis
yields an action spectrum (Fig. 19–44), often useful
in identifying the pigment primarily responsible for a bi-
ological effect of light. By capturing light in a region of
the spectrum not used by other organisms, a photosyn-
thetic organism can claim a unique ecological niche. For
example, the phycobilins in red algae and cyanobacte-
ria absorb light in the range 520 to 630 nm (Fig. 19–41),
allowing them to occupy niches where light of lower or
higher wavelength has been filtered out by the pigments
of other organisms living in the water above them, or by
the water itself.
Chlorophyll Funnels the Absorbed Energy to Reaction
Centers by Exciton Transfer
The light-absorbing pigments of thylakoid or bacterial
membranes are arranged in functional arrays called
photosystems. In spinach chloroplasts, for example,
each photosystem contains about 200 chlorophyll and
50 carotenoid molecules. All the pigment molecules in a
photosystem can absorb photons, but only a few chloro-
phyll molecules associated with the photochemical re-
action center are specialized to transduce light into
chemical energy. The other pigment molecules in a
photosystem are called light-harvesting or antenna
molecules. They absorb light energy and transmit it
rapidly and efficiently to the reaction center (Fig. 19–45).
The chlorophyll molecules in light-harvesting com-
plexes have light-absorption properties that are subtly
different from those of free chlorophyll. When isolated
chlorophyll molecules in vitro are excited by light, the
absorbed energy is quickly released as fluorescence and
heat, but when chlorophyll in intact leaves is excited by
visible light (Fig. 19–46, step 1 ), very little fluores-
cence is observed. Instead, the excited antenna chloro-
phyll transfers energy directly to a neighboring chloro-
phyll molecule, which becomes excited as the first
molecule returns to its ground state (step 2 ). This
transfer of energy, exciton transfer, extends to a third,
fourth, or subsequent neighbor, until one of a special
pair of chlorophyll a molecules at the photochemical re-
action center is excited (step 3 ). In this excited chloro-
phyll molecule, an electron is promoted to a higher-
energy orbital. This electron then passes to a nearby
electron acceptor that is part of the electron-transfer
chain, leaving the reaction-center chlorophyll with a
Chapter 19 Oxidative Phosphorylation and Photophosphorylation728
(a)
400
20
Relative rate of photosynthesis
Wavelength (nm)
40
60
80
100
0
500 600 700
(b)
FIGURE 19–44 Two ways to determine the action spectrum for pho-
tosynthesis. (a) Results of a classic experiment performed by T. W. En-
glemann in 1882 to determine the wavelength of light that is most ef-
fective in supporting photosynthesis. Englemann placed cells of a
filamentous photosynthetic alga on a microscope slide and illuminated
them with light from a prism, so that one part of the filament received
mainly blue light, another part yellow, another red. To determine which
algal cells carried out photosynthesis most actively, Englemann also
placed on the microscope slide bacteria known to migrate toward re-
gions of high O
2
concentration. After a period of illumination, the dis-
tribution of bacteria showed highest O
2
levels (produced by photo-
synthesis) in the regions illuminated with violet and red light.
(b) Results of a similar experiment that used modern techniques
(an oxygen electrode) for the measurement of O
2
production. An ac-
tion spectrum (as shown here) describes the relative rate of photo-
synthesis for illumination with a constant number of photons of dif-
ferent wavelengths. An action spectrum is useful because, by
comparison with absorption spectra (such as those in Fig. 19–41), it
suggests which pigments can channel energy into photosynthesis.
8885d_c19_690-750 3/1/04 11:32 AM Page 728 mac76 mac76:385_reb:
19.7 Light Absorption 729
These
molecules
absorb light
energy,
transferring
it between
molecules
until it
reaches the
reaction
center.
Antenna chlorophylls,
bound to protein
Carotenoids, other
accessory pigments
Light
Reaction center
Photochemical reaction here
converts the energy of a photon
into a separation of charge,
initiating electron flow.
FIGURE 19–45 Organization of photosystems in the thylakoid mem-
brane. Photosystems are tightly packed in the thylakoid membrane,
with several hundred antenna chlorophylls and accessory pigments
surrounding a photoreaction center. Absorption of a photon by any of
the antenna chlorophylls leads to excitation of the reaction center by
exciton transfer (black arrow). Also embedded in the thylakoid mem-
brane are the cytochrome b
6
f complex and ATP synthase (see Fig.
19–52).
1
2
3
4
5
The absorption of a photon has caused
separation of charge in the reaction center.
Antenna
molecules
Reaction-center
chlorophyll
*
*
*
Electron
acceptor
Electron
donor
–
+
–
+
Light
The electron hole in
the reaction center is
filled by an electron
from an electron
donor.
The excited reaction-
center chlorophyll
passes an electron to
an electron acceptor.
This energy is
transferred to a
reaction-center
chlorophyll,
exciting it.
The excited antenna
molecule passes
energy to a
neighboring
chlorophyll molecule
(resonance
energy transfer),
exciting it.
Light excites an
antenna molecule
(chlorophyll or
accessory pigment),
raising an electron
to a higher
energy level.
FIGURE 19–46 Exciton and electron transfer. This generalized
scheme shows conversion of the energy of an absorbed photon into
separation of charges at the reaction center. The steps are further de-
scribed in the text. Note that step 1 may repeat between succes-
sive antenna molecules until the exciton reaches a reaction-center
chlorophyll. The asterisk (*) represents the excited state of an antenna
molecule.
missing electron (an “electron hole,” denoted by H11001 in
Fig. 19–46) (step 4 ). The electron acceptor acquires
a negative charge in this transaction. The electron lost
by the reaction-center chlorophyll is replaced by an
electron from a neighboring electron-donor molecule
(step 5 ), which thereby becomes positively charged.
In this way, excitation by light causes electric charge
separation and initiates an oxidation-reduction
chain.
8885d_c19_690-750 3/1/04 11:32 AM Page 729 mac76 mac76:385_reb:
SUMMARY 19.7 Light Absorption
■ Photophosphorylation in the chloroplasts of
green plants and in cyanobacteria involves
electron flow through a series of
membrane-bound carriers.
■ In the light reactions of plants, absorption of a
photon excites chlorophyll molecules and other
(accessory) pigments, which funnel the energy
into reaction centers in the thylakoid
membranes. In the reaction centers, photo-
excitation results in a charge separation that
produces a strong electron donor (reducing
agent) and a strong electron acceptor.
19.8 The Central Photochemical Event:
Light-Driven Electron Flow
Light-driven electron transfer in plant chloroplasts dur-
ing photosynthesis is accomplished by multienzyme sys-
tems in the thylakoid membrane. Our current picture of
photosynthetic mechanisms is a composite, drawn from
studies of plant chloroplasts and a variety of bacteria
and algae. Determination of the molecular structures of
bacterial photosynthetic complexes (by x-ray crystal-
lography) has given us a much improved understanding
of the molecular events in photosynthesis in general.
Bacteria Have One of Two Types of Single
Photochemical Reaction Center
One major insight from studies of photosynthetic bacte-
ria came in 1952 when Louis Duysens found that illumi-
nation of the photosynthetic membranes of the purple
bacterium Rhodospirillum rubrum with a pulse of light
of a specific wavelength (870 nm) caused a temporary
decrease in the absorption of light at that wavelength; a
pigment was “bleached” by 870 nm light. Later studies
by Bessel Kok and Horst Witt showed similar bleaching
of plant chloroplast pigments by light of 680 and 700 nm.
Furthermore, addition of the (nonbiological) electron ac-
ceptor [Fe(CN)
6
]
3H11002
(ferricyanide) caused bleaching at
these wavelengths without illumination. These find-
ings indicated that bleaching of the pigments was due to
the loss of an electron from a photochemical reaction
center. The pigments were named for the wavelength of
maximum bleaching: P870, P680, and P700.
Photosynthetic bacteria have relatively simple pho-
totransduction machinery, with one of two general types
of reaction center. One type (found in purple bacteria)
passes electrons through pheophytin (chlorophyll lack-
ing the central Mg
2H11001
ion) to a quinone. The other (in
green sulfur bacteria) passes electrons through a
quinone to an iron-sulfur center. Cyanobacteria and
plants have two photosystems (PSI, PSII), one of each
type, acting in tandem. Biochemical and biophysical
studies have revealed many of the molecular details of
reaction centers of bacteria, which therefore serve as
prototypes for the more complex phototransduction
systems of plants.
The Pheophytin-Quinone Reaction Center (Type II Reaction Cen-
ter) The photosynthetic machinery in purple bacteria
consists of three basic modules (Fig. 19–47a): a single
reaction center (P870), a cytochrome bc
1
electron-
transfer complex similar to Complex III of the mito-
chondrial electron-transfer chain, and an ATP synthase,
also similar to that of mitochondria. Illumination drives
electrons through pheophytin and a quinone to the cy-
tochrome bc
1
complex; after passing through the com-
plex, electrons flow through cytochrome c
2
back to the
reaction center, restoring its preillumination state. This
light-driven cyclic flow of electrons provides the energy
for proton pumping by the cytochrome bc
1
complex.
Powered by the resulting proton gradient, ATP synthase
produces ATP, exactly as in mitochondria.
The three-dimensional structures of the reaction
centers of purple bacteria (Rhodopseudomonas viridis
and Rhodobacter sphaeroides), deduced from x-ray
crystallography, shed light on how phototransduction
takes place in a pheophytin-quinone reaction center.
The R. viridis reaction center (Fig. 19–48a) is a large
protein complex containing four polypeptide subunits
and 13 cofactors: two pairs of bacterial chlorophylls, a
pair of pheophytins, two quinones, a nonheme iron, and
four hemes in the associated c-type cytochrome.
The extremely rapid sequence of electron transfers
shown in Figure 19–48b has been deduced from physi-
cal studies of the bacterial pheophytin-quinone centers,
using brief flashes of light to trigger phototransduction
and a variety of spectroscopic techniques to follow the
flow of electrons through several carriers. A pair of
bacteriochlorophylls—the “special pair,” designated
(Chl)
2
—is the site of the initial photochemistry in the
bacterial reaction center. Energy from a photon absorbed
by one of the many antenna chlorophyll molecules sur-
rounding the reaction center reaches (Chl)
2
by exciton
transfer. When these two chlorophyll molecules—so
close that their bonding orbitals overlap—absorb an ex-
citon, the redox potential of (Chl)
2
is shifted, by an
amount equivalent to the energy of the photon, con-
verting the special pair to a very strong electron donor.
The (Chl)
2
donates an electron that passes through a
neighboring chlorophyll monomer to pheophytin (Pheo).
This produces two radicals, one positively charged (the
special pair of chlorophylls) and one negatively charged
(the pheophytin):
(Chl)
2
H11001 1 exciton 88n (Chl)
2
*
(excitation)
(Chl)
2
*
H11001 Pheo 88n
H11080
(Chl)
2
H11001
H11001
H11080
Pheo
H11002
(charge separation)
The pheophytin radical now passes its electron to a
tightly bound molecule of quinone (Q
A
), converting it
to a semiquinone radical, which immediately donates its
Chapter 19 Oxidative Phosphorylation and Photophosphorylation730
8885d_c19_690-750 3/1/04 11:32 AM Page 730 mac76 mac76:385_reb:
extra electron to a second, loosely bound quinone (Q
B
).
Two such electron transfers convert Q
B
to its fully re-
duced form, Q
B
H
2
, which is free to diffuse in the mem-
brane bilayer, away from the reaction center:
2
H11080
Pheo
H11002
H11001 2H
H11001
H11001 Q
B
88n 2 Pheo H11001 Q
B
H
2
(quinone reduction)
The hydroquinone (Q
B
H
2
), carrying in its chemical
bonds some of the energy of the photons that originally
excited P870, enters the pool of reduced quinone (QH
2
)
dissolved in the membrane and moves through the lipid
phase of the bilayer to the cytochrome bc
1
complex.
Like the homologous Complex III in mitochondria,
the cytochrome bc
1
complex of purple bacteria carries
electrons from a quinol donor (QH
2
) to an electron ac-
ceptor, using the energy of electron transfer to pump
protons across the membrane, producing a proton-
motive force. The path of electron flow through this
complex is believed to be very similar to that through
mitochondrial Complex III, involving a Q cycle (Fig.
19–12) in which protons are consumed on one side of
the membrane and released on the other. The ultimate
electron acceptor in purple bacteria is the electron-
depleted form of P870,
H11080
(Chl)
2
H11001
(Fig. 19–47a). Electrons
move from the cytochrome bc
1
complex to P870 via a
soluble c-type cytochrome, cytochrome c
2
. The electron-
transfer process completes the cycle, returning the
reaction center to its unbleached state, ready to absorb
another exciton from antenna chlorophyll.
A remarkable feature of this system is that all the
chemistry occurs in the solid state, with reacting species
held close together in the right orientation for reaction.
The result is a very fast and efficient series of reactions.
The Fe-S Reaction Center (Type I Reaction Center) Photo-
synthesis in green sulfur bacteria involves the same
three modules as in purple bacteria, but the process dif-
fers in several respects and involves additional enzy-
matic reactions (Fig. 19–47b). Excitation causes an
electron to move from the reaction center to the cy-
tochrome bc
1
complex via a quinone carrier. Electron
transfer through this complex powers proton transport
and creates the proton-motive force used for ATP syn-
thesis, just as in purple bacteria and in mitochondria.
19.8 The Central Photochemical Event: Light-Driven Electron Flow 731
0.5
–1.0
–0.5
0
Cyt
c
2
e
–
e
–
RC
P870
P870*
Proton
gradient
Purple bacteria
(pheophytin-quinone type)
(a)
Green sulfur bacteria
(Fe-S type)
(b)
Q
Pheo
Proton
gradient
Cyt
bc
1
complex
Excitons
Cyt
c
553
RC
P840
P840*
Q
Fd
e
–
Cyt
bc
1
complex
Fd-NAD
reductase
Excitons
NAD
H11001
NADH
E
H11032H11034
(volts)
FIGURE 19–47 Functional modules of photosynthetic machinery in
purple bacteria and green sulfur bacteria. (a) In purple bacteria, light
energy drives electrons from the reaction center P870 through pheo-
phytin (Pheo), a quinone (Q), and the cytochrome bc
1
complex, then
through cytochrome c
2
back to the reaction center. Electron flow through
the cytochrome bc
1
complex causes proton pumping, creating an elec-
trochemical potential that powers ATP synthesis. (b) Green sulfur bac-
teria have two routes for electrons driven by excitation of P840: a cyclic
route passes through a quinone to the cytochrome bc
1
complex and
back to the reaction center via cytochrome c, and a noncyclic route
from the reaction center through the iron-sulfur protein ferredoxin (Fd),
then to NAD
H11001
in a reaction catalyzed by ferredoxin:NAD reductase.
8885d_c19_690-750 3/1/04 11:32 AM Page 731 mac76 mac76:385_reb:
(a)
Bacteriochlorophyll (2)
((Chl)
2
, the special pair)
(270 ns)
(3 ps)
(200 ps)
Hemes of
c-type
cytochrome
Bacteriochlorophyll (2)
(accessory pigments)
Bacteriopheophytin (2)
Q
A
(quinone)
Q
B
(quinone)
Fe
Light
3
4
1
2
(b) (6 s)H9262
P side
N side
FIGURE 19–48 Photoreaction center of the purple bacterium
Rhodopseudomonas viridis. (PDB ID 1PRC) (a) The system has four
components: three subunits, H, M, and L (brown, blue, and gray, re-
spectively), with a total of 11 transmembrane helical segments, and a
fourth protein, cytochrome c (yellow), associated with the complex at
the membrane surface. Subunits L and M are paired transmembrane
proteins that together form a cylindrical structure with roughly bilat-
eral symmetry about its long axis. Shown as space-filling models (and
in (b) as ball-and-stick structures) are the prosthetic groups that par-
ticipate in the photochemical events. Bound to the L and M chains
are two pairs of bacteriochlorophyll molecules (green); one of the pairs
(the “special pair,” (Chl)
2
) is the site of the first photochemical changes
after light absorption. Also incorporated in the system are a pair of
pheophytin a (Pheo a) molecules (blue); two quinones, menaquinone
(Q
A
) and ubiquinone (Q
B
) (orange and yellow), also arranged with bi-
lateral symmetry; and a single nonheme Fe (red) located approximately
on the axis of symmetry between the quinones. Shown at the top of
the figure are four heme groups (red) associated with the c-type cy-
tochrome of the reaction center. The reaction center of another pur-
ple bacterium, Rhodobacter sphaeroides, is very similar, except that
cytochrome c is not part of the crystalline complex.
(b) Sequence of events following excitation of the special pair of
bacteriochlorophylls (all components colored as in (a)), with the time
scale of the electron transfers in parentheses. 1 The excited special
pair passes an electron to pheophytin, 2 from which the electron
moves rapidly to the tightly bound menaquinone, Q
A
. 3 This quinone
passes electrons much more slowly to the diffusible ubiquinone, Q
B
,
through the nonheme Fe. Meanwhile, 4 the “electron hole” in the
special pair is filled by an electron from a heme of cytochrome c.
However, in contrast to the cyclic flow of electrons in
purple bacteria, some electrons flow from the reaction
center to an iron-sulfur protein, ferredoxin, which then
passes electrons via ferredoxin:NAD reductase to
NAD
H11001
, producing NADH. The electrons taken from the
reaction center to reduce NAD
H11001
are replaced by the ox-
idation of H
2
S to elemental S, then to SO
4
2H11002
, in the re-
action that defines the green sulfur bacteria. This oxi-
dation of H
2
S by bacteria is chemically analogous to the
oxidation of H
2
O by oxygenic plants.
Kinetic and Thermodynamic Factors Prevent the
Dissipation of Energy by Internal Conversion
The complex construction of reaction centers is the
product of evolutionary selection for efficiency in the
photosynthetic process. The excited state (Chl)
2
*
could
in principle decay to its ground state by internal con-
version, a very rapid process (10 picoseconds; 1 ps H11005
10
H1100212
s) in which the energy of the absorbed photon is
converted to heat (molecular motion). Reaction centers
are constructed to prevent the inefficiency that would
result from internal conversion. The proteins of the re-
action center hold the bacteriochlorophylls, bacterio-
pheophytins, and quinones in a fixed orientation rela-
tive to each other, allowing the photochemical reactions
to take place in a virtually solid state. This accounts for
the high efficiency and rapidity of the reactions; noth-
ing is left to chance collision or random diffusion. Exci-
ton transfer from antenna chlorophyll to the special pair
of the reaction center is accomplished in less than 100
ps with H1102290% efficiency. Within 3 ps of the excitation
of P870, pheophytin has received an electron and be-
come a negatively charged radical; less than 200 ps later,
the electron has reached the quinone Q
B
(Fig. 19–48b).
The electron-transfer reactions not only are fast but are
thermodynamically “downhill”; the excited special pair
(Chl)
2
*
is a very good electron donor (EH11032H11034 H110021 V), and
each successive electron transfer is to an acceptor of
substantially less negative EH11032H11034. The standard free-energy
732
8885d_c19_690-750 3/1/04 11:32 AM Page 732 mac76 mac76:385_reb:
change for the process is therefore negative and large;
recall from Chapter 13 that H9004GH11032H11034 H11005 H11002n H9004EH11032H11034; here,
H9004EH11032H11034 is the difference between the standard reduction
potentials of the two half-reactions
(1) (Chl)
2
*
88n
H11080
(Chl)
2
H11001
H11001 e
H11002
EH11032H11034 ≈ H110021.0 V
(2) Q H11001 2H
H11001
H11001 2e
H11002
88n QH
2
EH11032H11034 H11005 H110020.045 V
Thus
H9004EH11032H11034 H11005 H110020.045 V H11002 (H110021.0 V) ≈ 0.95 V
and
H9004GH11032H11034 H11005 H110022(96.5 kJ/V H11554 mol)(0.95 V) H11005H11002180 kJ/mol
The combination of fast kinetics and favorable thermo-
dynamics makes the process virtually irreversible and
highly efficient. The overall energy yield (the percent-
age of the photon’s energy conserved in QH
2
) is H1102230%,
with the remainder of the energy dissipated as heat.
In Plants, Two Reaction Centers Act in Tandem
The photosynthetic apparatus of modern cyanobacteria,
algae, and vascular plants is more complex than the one-
center bacterial systems, and it appears to have evolved
through the combination of two simpler bacterial pho-
tocenters. The thylakoid membranes of chloroplasts
have two different kinds of photosystems, each with its
own type of photochemical reaction center and set of
antenna molecules. The two systems have distinct and
complementary functions (Fig. 19–49). Photosystem
II (PSII) is a pheophytin-quinone type of system (like
the single photosystem of purple bacteria) containing
roughly equal amounts of chlorophylls a and b. Excita-
tion of its reaction center P680 drives electrons through
the cytochrome b
6
f complex with concomitant move-
ment of protons across the thylakoid membrane. Pho-
tosystem I (PSI) is structurally and functionally
related to the type I reaction center of green sulfur bac-
teria. It has a reaction center designated P700 and a
high ratio of chlorophyll a to chlorophyll b. Excited P700
passes electrons to the Fe-S protein ferredoxin, then to
NADP
H11001
, producing NADPH. The thylakoid membranes
of a single spinach chloroplast have many hundreds of
each kind of photosystem.
These two reaction centers in plants act in tandem
to catalyze the light-driven movement of electrons from
H
2
O to NADP
H11001
(Fig. 19–49). Electrons are carried be-
tween the two photosystems by the soluble protein
plastocyanin, a one-electron carrier functionally simi-
lar to cytochrome c of mitochondria. To replace the elec-
trons that move from PSII through PSI to NADP
H11001
,
cyanobacteria and plants oxidize H
2
O (as green sulfur
19.8 The Central Photochemical Event: Light-Driven Electron Flow 733
1.0
0
E
H11032H11034
(volts)
–1.0
P680*
P680
P700*
P700
Photosystem II
Photosystem I
NADP
+
Light
A
0
A
1
Fe-S
Fd
Fd:NADP
+
oxidoreductase
Plastocyanin
Proton
gradient
PQ
B
PQ
A
Pheo
Cyt
b
6
f
complex
e
–
H
2
O
O
2
Light
PQ
A
= plastoquinone
PQ
B
= second quinone
A
0
= electron acceptor chlorophyll
A
1
= phylloquinone
NADPH
O
2
evolving
complex
1
2
FIGURE 19–49 Integration of photosys-
tems I and II in chloroplasts. This “Z
scheme” shows the pathway of electron
transfer from H
2
O (lower left) to NADP
H11001
(far right) in noncyclic photosynthesis. The
position on the vertical scale of each
electron carrier reflects its standard reduc-
tion potential. To raise the energy of
electrons derived from H
2
O to the energy
level required to reduce NADP
H11001
to
NADPH, each electron must be “lifted”
twice (heavy arrows) by photons absorbed
in PSII and PSI. One photon is required
per electron in each photosystem. After
excitation, the high-energy electrons flow
“downhill” through the carrier chains
shown. Protons move across the thylakoid
membrane during the water-splitting
reaction and during electron transfer
through the cytochrome b
6
f complex,
producing the proton gradient that is
central to ATP formation. The dashed
arrow is the path of cyclic electron
transfer (discussed later in the text), which
involves only PSI; electrons return via the
cyclic pathway to PSI, instead of reducing
NADP
H11001
to NADPH.
8885d_c19_690-750 3/1/04 11:32 AM Page 733 mac76 mac76:385_reb:
D2
Lumen
(P side)
Stroma
(N side)
Pheo
Pheo
P680
2H
2
O4H
+
+ O
2
D1
(Chl)
2
D
2
(Chl)
2
D
1
Tyr
D
PQ
B
Tyr
Z
Mn
4
PQ
A
F
e
3
1
2
4
5
FIGURE 19–50 Photosystem II of the cyanobacterium Synechococcus
elongates. The monomeric form of the complex shown here has two
major transmembrane proteins, D1 and D2, each with its set of co-
factors. Although the two subunits are nearly symmetric, electron flow
occurs through only one of the two branches of cofactors, that on the
right (on D1). The arrows show the path of electron flow from the Mn
ion cluster (Mn
4
, purple) of the water-splitting enzyme to the quinone
PQ
B
(orange). The photochemical events occur in the sequence indi-
cated by the step numbers. Notice the close similarity between the
positions of cofactors here and the positions in the bacterial photore-
action center shown in Figure 19–48. The role of the Tyr residues is
discussed later in the text.
bacteria oxidize H
2
S), producing O
2
(Fig. 19–49, bottom
left). This process is called oxygenic photosynthesis
to distinguish it from the anoxygenic photosynthesis of
purple and green sulfur bacteria. All O
2
-evolving
photosynthetic cells—those of plants, algae, and
cyanobacteria—contain both PSI and PSII; organisms
with only one photosystem do not evolve O
2
. The dia-
gram in Figure 19–49, often called the Z scheme be-
cause of its overall form, outlines the pathway of elec-
tron flow between the two photosystems and the energy
relationships in the light reactions. The Z scheme thus
describes the complete route by which electrons flow
from H
2
O to NADP
H11001
, according to the equation
2H
2
O H11001 2NADP
H11001
H11001 8 photons On O
2
H11001 2NADPH H11001 2H
H11001
For every two photons absorbed (one by each photo-
system), one electron is transferred from H
2
O to
NADP
H11001
. To form one molecule of O
2
, which requires
transfer of four electrons from two H
2
O to two NADP
H11001
,
a total of eight photons must be absorbed, four by each
photosystem.
The mechanistic details of the photochemical reac-
tions in PSII and PSI are essentially similar to those of
the two bacterial photosystems, with several important
additions. In PSII, two very similar proteins, D1 and D2,
form an almost symmetrical dimer, to which all the
electron-carrying cofactors are bound (Fig. 19–50). Ex-
citation of P680 in PSII produces P680*, an excellent
electron donor that, within picoseconds, transfers an
electron to pheophytin, giving it a negative charge
(
H11080
Pheo
H11002
). With the loss of its electron, P680* is trans-
formed into a radical cation, P680
H11001
.
H11080
Pheo
H11002
very rap-
idly passes its extra electron to a protein-bound plas-
toquinone, PQ
A
(or Q
A
), which in turn passes its
electron to another, more loosely bound plastoquinone,
PQ
B
(or Q
B
). When PQ
B
has acquired two electrons in
two such transfers from PQ
A
and two protons from the
solvent water, it is in its fully reduced quinol form,
PQ
B
H
2
. The overall reaction initiated by light in PSII is
4P680 H11001 4H
H11001
H11001 2PQ
B
H11001 4 photons On
4P680
H11001
H11001 2PQ
B
H
2
(19–12)
Eventually, the electrons in PQ
B
H
2
pass through the
cytochrome b
6
f complex (Fig. 19–49). The electron ini-
tially removed from P680 is replaced with an electron
obtained from the oxidation of water, as described
below. The binding site for plastoquinone is the point of
action of many commercial herbicides that kill plants by
blocking electron transfer through the cytochrome b
6
f
complex and preventing photosynthetic ATP production.
The photochemical events that follow excitation of
PSI at the reaction center P700 are formally similar to
those in PSII. The excited reaction center P700* loses
an electron to an acceptor, A
0
(believed to be a special
form of chlorophyll, functionally homologous to the
pheophytin of PSII), creating A
0
H11002
and P700
H11001
(Fig. 19–49,
right side); again, excitation results in charge separation
at the photochemical reaction center. P700
H11001
is a strong
oxidizing agent, which quickly acquires an electron from
plastocyanin, a soluble Cu-containing electron-transfer
protein. A
0
H11002
is an exceptionally strong reducing agent
that passes its electron through a chain of carriers that
leads to NADP
H11001
. First, phylloquinone (A
1
) accepts an
electron and passes it to an iron-sulfur protein (through
three Fe-S centers in PSI). From here, the electron
moves to ferredoxin (Fd), another iron-sulfur protein
loosely associated with the thylakoid membrane. Spinach
ferredoxin (M
r
10,700) contains a 2Fe-2S center (Fig.
19–5) that undergoes one-electron oxidation and reduc-
tion reactions. The fourth electron carrier in the chain
is the flavoprotein ferredoxin : NADP
H11001
oxidoreduc-
tase, which transfers electrons from reduced ferredoxin
(Fd
red
) to NADP
H11001
:
2Fd
red
H11001 2H
H11001
H11001 NADP
H11001
On 2Fdox H11001 NADPH H11001 H
H11001
This enzyme is homologous to the ferredoxin:NAD re-
ductase of green sulfur bacteria (Fig. 19–47b).
Antenna Chlorophylls Are Tightly Integrated
with Electron Carriers
The electron-carrying cofactors of PSI and the light-
harvesting complexes are part of a supramolecular com-
plex (Fig. 19–51a), the structure of which has been
solved crystallographically. The protein consists of three
identical complexes, each composed of 11 different pro-
teins (Fig. 19–51b). In this remarkable structure the
many antenna chlorophyll and carotenoid molecules are
Chapter 19 Oxidative Phosphorylation and Photophosphorylation734
8885d_c19_690-750 3/1/04 11:32 AM Page 734 mac76 mac76:385_reb:
19.8 The Central Photochemical Event: Light-Driven Electron Flow 735
FIGURE 19–51 The supramolecular complex of PSI and its associated
antenna chlorophylls. (a) Schematic drawing of the essential proteins
and cofactors in a single unit of PSI. A large number of antenna chloro-
phylls surround the reaction center and convey to it (red arrows) the
energy of photons they have absorbed. The result is excitation of the
pair of chlorophyll molecules that constitute P700. Excitation of P700
greatly decreases its reduction potential, and it passes an electron
through two nearby chlorophylls to phylloquinone (Q
K
; also called
A
1
). Reduced phylloquinone is reoxidized as it passes two electrons,
one at a time, to an Fe-S protein (F
X
) near the N side of the mem-
brane. From F
X
, electrons move through two more Fe-S centers (F
A
and F
B
), then to the Fe-S protein ferredoxin in the stroma. Ferredoxin
then donates electrons to NADP
H11001
(not shown), reducing it to NADPH,
one of the forms in which the energy of photons is trapped in chloro-
plasts. (b) The trimeric structure (derived from PDB ID 1JBO), viewed
from the thylakoid lumen perpendicular to the membrane, showing
all protein subunits (gray) and cofactors (described in (c)). (c) A
monomer of PSI with all the proteins omitted, revealing the antenna
and reaction center chlorophylls (green with dark green Mg
2H11001
ions
in the center), carotenoids (yellow), and Fe-S centers of the reaction
center (space-filling red and orange structures). The proteins in the
complex hold the components rigidly in orientations that maximize
efficient exciton transfers between excited antenna molecules and the
reaction center.
Light
Subunit B
PSI
Stroma
(N side)
Lumen
(P side)
Chl
F
X
F
A
Subunit
C
F
B
(Chl)
A
0
Chl
(Chl)
A
0
Q
K
Q
K
e
–
e
–
e
–
e
–
Subunit A
(a)
Ferredoxin
Plastocyanin
(Chl)
2
P
700
Exciton
transfer
(b) (c)
8885d_c19_690-750 3/1/04 11:32 AM Page 735 mac76 mac76:385_reb:
Chapter 19 Oxidative Phosphorylation and Photophosphorylation736
Photosystem I
Stacked membranes
(granal lamellae)
Photosystem II
ATP synthase
Cytochrome
b
6
f complex
Light-harvesting
complex II
Unstacked
membranes
(stromal
lamellae)
Stroma
Lumen
FIGURE 19–52 Localization of PSI and
PSII in thylakoid membranes. Light-
harvesting complex LHCII and ATP
synthase are located in regions of the
thylakoid membrane that are appressed
(granal lamellae, in which several
membranes are in contact) and in
regions that are not appressed (stromal
lamellae) and have ready access to ADP
and NADP
H11001
in the stroma. Photosystem
II is present almost exclusively in the
appressed regions, and photosystem I
almost exclusively in nonappressed
regions exposed to the stroma. LHCII is
the “adhesive” that holds appressed
lamellae together (see Fig. 19–53).
LHCII
–Thr–OH
–Thr–
Thylakoid
membrane
Appressed Nonappressed
LHCII
ATP
P
i
ADP
protein
phosphatase
protein
kinase
P
FIGURE 19–53 Equalization of electron flow in PSI and PSII by mod-
ulation of granal stacking. A hydrophobic domain of LHCII in one
thylakoid lamella inserts into the neighboring lamella and closely ap-
presses the two membranes. Accumulation of plastoquinol (not shown)
stimulates a protein kinase that phosphorylates a Thr residue in the
hydrophobic domain of LHCII, which reduces its affinity for the neigh-
boring thylakoid membrane and converts appressed regions (granal
lamellae) to nonappressed regions (stromal lamellae). A specific pro-
tein phosphatase reverses this regulatory phosphorylation when the
[PQ]/[PQH
2
] ratio increases.
precisely arrayed around the reaction center (Fig.
19–51c). The reaction center’s electron-carrying cofac-
tors are therefore tightly integrated with antenna
chlorophylls. This arrangement allows very rapid and ef-
ficient exciton transfer from antenna chlorophylls to the
reaction center. In contrast to the single path of electrons
in PSII, the electron flow initiated by absorption of a
photon is believed to occur through both branches of
carriers in PSI.
Spatial Separation of Photosystems I and II Prevents
Exciton Larceny
The energy required to excite PSI (P700) is less than
that needed to excite PSII (P680) (shorter wavelength
corresponds to higher energy). If PSI and PSII were
physically contiguous, excitons originating in the an-
tenna system of PSII would migrate to the reaction cen-
ter of PSI, leaving PSII chronically underexcited and in-
terfering with the operation of the two-center system.
This “exciton larceny” is prevented by separation of PSI
and PSII in the thylakoid membrane (Fig. 19–52). PSII
is located almost exclusively in the tightly appressed
membrane stacks of thylakoid grana (granal lamellae);
its associated light-harvesting complex (LHCII) medi-
ates the tight association of adjacent membranes in the
grana. PSI and the ATP synthase complex are located
almost exclusively in the thylakoid membranes that are
not appressed (the stromal lamellae), where both have
access to the contents of the stroma, including ADP and
NADP
H11001
. The cytochrome b
6
f complex is present
throughout the thylakoid membrane.
The association of LHCII with PSII is regulated by
light intensity and wavelength. In bright sunlight (with
a large component of blue light), PSII absorbs more light
than PSI and produces reduced plastoquinone (plasto-
quinol, PQH
2
) faster than PSI can oxidize it. The re-
sulting accumulation of PQH
2
activates a protein kinase
that phosphorylates a Thr residue on LHCII (Fig. 19–53).
Phosphorylation weakens the interaction of LHCII with
PSII, and some LHCII dissociates and moves to the stro-
mal lamellae; here it captures photons for PSI, speed-
ing the oxidation of PQH
2
and reversing the imbalance
between electron flow in PSI and PSII. In less intense
light (in the shade, with more red light), PSI oxidizes
PQH
2
faster than PSII can make it, and the resulting in-
crease in PQ concentration triggers dephosphorylation
of LHCII, reversing the effect of phosphorylation.
8885d_c19_690-750 3/1/04 11:32 AM Page 736 mac76 mac76:385_reb:
Subunit IV
Cyt b
6
b
L
b
H
Cyt f
Rieske iron-
sulfur protein
Thylakoid lumen
(P side)
Q cycle
e
– e
–
Stroma (N side)
Cu
Plasto-
cyanin
PQ PQH
2
4H
+
2 × 2H
+
(c)
FIGURE 19–54 Electron and proton flow through the
cytochrome b
6
f complex. (a) The crystal structure of the
complex (PDB ID 1UM3) reveals the positions of the cofac-
tors involved in electron transfers. In addition to the hemes
of cytochrome b (heme b
H
and b
L
; also called heme bN
and b
P
, respectively, because of their proximity to the N and
P sides of the bilayer) and that of cytochrome f (heme f ),
there is a fourth (heme x) near heme b
H
, and there is a
H9252-carotene of unknown function. Two sites bind plasto-
quinone: the PQH
2
site near the P side of the bilayer, and
the PQ site near the N side. The Fe-S center of the Rieske
protein lies just outside the bilayer on the P side, and the
heme f site is on a protein domain that extends well into
the thylakoid lumen. (b) The complex is a homodimer
arranged to create a cavern connecting the PQH
2
and PQ
sites (compare with the structure of mitochondrial Complex
III in Fig. 19–12). This cavern allows plastoquinone
movement between the sites of its oxidation and reduction.
(c) Plastoquinol (PQH
2
) formed in PSII is oxidized by
the cytochrome b
6
f complex in a series of steps like those
of the Q cycle in the cytochrome bc
1
complex (Complex
III) of mitochondria (see Fig. 19–11). One electron from
PQH
2
passes to the Fe-S center of the Rieske protein
(purple), the other to heme b
L
of cytochrome b
6
(green).
The net effect is passage of electrons from PQH
2
to the
soluble protein plastocyanin, which carries them to PSI.
The Cytochrome b
6
f Complex Links
Photosystems II and I
Electrons temporarily stored in plastoquinol as a result
of the excitation of P680 in PSII are carried to P700 of
PSI via the cytochrome b
6
f complex and the soluble pro-
tein plastocyanin (Fig. 19–49, center). Like Complex III
of mitochondria, the cytochrome b
6
f complex (Fig.
19–54) contains a b-type cytochrome with two heme
groups (designated b
H
and b
L
), a Rieske iron-sulfur pro-
tein (M
r
20,000), and cytochrome f (for the Latin frons,
“leaf”). Electrons flow through the cytochrome b
6
f com-
plex from PQ
B
H
2
to cytochrome f, then to plastocyanin,
and finally to P700, thereby reducing it.
19.8 The Central Photochemical Event: Light-Driven Electron Flow 737
Stroma
(N side)
Lumen
(P side)
Heme ?
Fe-S center
Plastocyanin
H
+
2H
+
Heme x
PQ
PQH
2
Heme b
H
Heme b
L
e
–
e
–
e
–
e
–
e
–
(a) (b)
b-carotene
8885d_c19_737 3/1/04 2:10 PM Page 737 mac76 mac76:385_reb:
what is available in a particular ecological niche. About
3 billion years ago, evolution of primitive photosynthetic
bacteria (the progenitors of the modern cyanobacteria)
produced a photosystem capable of taking electrons
from a donor that is always available—water. Two water
molecules are split, yielding four electrons, four protons,
and molecular oxygen:
2H
2
O 88n 4H
H11001
H11001 4e
H11002
H11001 O
2
A single photon of visible light does not have enough
energy to break the bonds in water; four photons are
required in this photolytic cleavage reaction.
The four electrons abstracted from water do not
pass directly to P680
H11001
, which can accept only one elec-
tron at a time. Instead, a remarkable molecular device,
the oxygen-evolving complex (also called the water-
splitting complex), passes four electrons one at a
Chapter 19 Oxidative Phosphorylation and Photophosphorylation738
Like Complex III of mitochondria, cytochrome b
6
f
conveys electrons from a reduced quinone—a mobile,
lipid-soluble carrier of two electrons (Q in mitochondria,
PQ
B
in chloroplasts)—to a water-soluble protein that
carries one electron (cytochrome c in mitochondria,
plastocyanin in chloroplasts). As in mitochondria, the
function of this complex involves a Q cycle (Fig. 19–12)
in which electrons pass, one at a time, from PQ
B
H
2
to
cytochrome b
6
. This cycle results in the pumping of pro-
tons across the membrane; in chloroplasts, the direction
of proton movement is from the stromal compartment
to the thylakoid lumen, up to four protons moving for
each pair of electrons. The result is production of a pro-
ton gradient across the thylakoid membrane as electrons
pass from PSII to PSI. Because the volume of the flat-
tened thylakoid lumen is small, the influx of a small num-
ber of protons has a relatively large effect on lumenal
pH. The measured difference in pH between the stroma
(pH 8) and the thylakoid lumen (pH 5) represents a
1,000-fold difference in proton concentration—a pow-
erful driving force for ATP synthesis.
Cyanobacteria Use the Cytochrome b
6
f Complex and
Cytochrome c
6
in Both Oxidative Phosphorylation and
Photophosphorylation
Cyanobacteria can synthesize ATP by oxidative phos-
phorylation or by photophosphorylation, although they
have neither mitochondria nor chloroplasts. The enzy-
matic machinery for both processes is in a highly con-
voluted plasma membrane (see Fig. 1–6). Two protein
components function in both processes (Fig. 19–55).
The proton-pumping cytochrome b
6
f complex carries
electrons from plastoquinone to cytochrome c
6
in pho-
tosynthesis, and also carries electrons from ubiquinone
to cytochrome c
6
in oxidative phosphorylation—the role
played by cytochrome bc
1
in mitochondria. Cytochrome
c
6
, homologous to mitochondrial cytochrome c, carries
electrons from Complex III to Complex IV in cyanobac-
teria; it can also carry electrons from the cytochrome
b
6
f complex to PSI—a role performed in plants by plas-
tocyanin. We therefore see the functional homology be-
tween the cyanobacterial cytochrome b
6
f complex and
the mitochondrial cytochrome bc
1
complex, and between
cyanobacterial cytochrome c
6
and plant plastocyanin.
Water Is Split by the Oxygen-Evolving Complex
The ultimate source of the electrons passed to NADPH
in plant (oxygenic) photosynthesis is water. Having
given up an electron to pheophytin, P680
H11001
(of PSII)
must acquire an electron to return to its ground state
in preparation for capture of another photon. In princi-
ple, the required electron might come from any number
of organic or inorganic compounds. Photosynthetic bac-
teria use a variety of electron donors for this purpose—
acetate, succinate, malate, or sulfide—depending on
H
2
O
2 Fd
red
2 Fd
ox
NADPH + H
+
NADP
+
PSII
PSI
Cyt b
6
f
Complex
III
NAD
+
NADH + H
+
H
2
OO
2
NADH
dehydrogenase
Complex
I
e
H11002
e
H11002
Cyt a H11001 a
3
Complex
IV
2
1
O
2
2
1
Plasto-
quinone
(a) (b)
Photophosphorylation
Oxidative
phosphorylation
Cyt c
6
Light
Light
FIGURE 19–55 Dual roles of cytochrome b
6
f and cytochrome c
6
in
cyanobacteria. Cyanobacteria use cytochrome b
6
f, cytochrome c
6
,
and plastoquinone for both oxidative phosphorylation and pho-
tophosphorylation. (a) In photophosphorylation, electrons flow (top to
bottom) from water to NADP
H11001
. (b) In oxidative phosphorylation, elec-
trons flow from NADH to O
2
. Both processes are accompanied by
proton movement across the membrane, accomplished by a Q cycle.
8885d_c19_690-750 3/1/04 11:32 AM Page 738 mac76 mac76:385_reb:
19.8 The Central Photochemical Event: Light-Driven Electron Flow 739
time to P680
H11001
(Fig. 19–56). The immediate electron
donor to P680
H11001
is a Tyr residue (often designated Z or
Tyr
z
) in protein subunit D1 of the PSII reaction center.
The Tyr residue loses both a proton and an electron,
generating the electrically neutral Tyr free radical,
H11080
Tyr:
4P680
H11001
H11001 4 Tyr 88n 4P680 H11001 4
H11080
Tyr (19–13)
The Tyr radical regains its missing electron and proton
by oxidizing a cluster of four manganese ions in the
water-splitting complex. With each single-electron
transfer, the Mn cluster becomes more oxidized; four
single-electron transfers, each corresponding to the ab-
sorption of one photon, produce a charge of H110014 on the
Mn complex (Fig. 19–56):
4
H11080
Tyr H11001 [Mn complex]
0
88n
4 Tyr H11001 [Mn complex]
4H11001
(19–14)
In this state, the Mn complex can take four electrons
from a pair of water molecules, releasing 4 H
H11001
and O
2
:
[Mn complex]
4H11001
H11001 2H
2
O 88n
[Mn complex]
0
H11001 4H
H11001
H11001 O
2
(19–15)
Because the four protons produced in this reaction are
released into the thylakoid lumen, the oxygen-evolving
complex acts as a proton pump, driven by electron
transfer. The sum of Equations 19–12 through 19–15 is
2H
2
O H11001 2PQ
B
H11001 4 photons On O
2
H11001 2PQ
B
H
2
(19–16)
The water-splitting activity associated with the PSII
reaction center has proved exceptionally difficult to pu-
rify. A peripheral membrane protein (M
r
33,000) on the
lumenal side of the thylakoid membrane is believed to
stabilize the Mn complex. In the crystal structure (PDB
ID 1FE1; see Fig. 19–50), four Mn ions are clustered
with precise geometry near a Tyr residue on the D1 sub-
unit, presumably the one involved in water oxidation.
Manganese can exist in stable oxidation states from H110012
to H110017, so a cluster of four Mn ions can certainly donate
or accept four electrons. The mechanism shown in Fig-
ure 19–56 is consistent with the experimental facts, but
until the exact chemical structures of all the interme-
diates of the Mn cluster are known, the detailed mech-
anism remains elusive.
SUMMARY 19.8 The Central Photochemical Event:
Light-Driven Electron Flow
■ Bacteria have a single reaction center; in purple
bacteria, it is of the pheophytin-quinone type,
and in green sulfur bacteria, the Fe-S type.
■ Structural studies of the reaction center of a
purple bacterium have provided information
about light-driven electron flow from an
excited special pair of chlorophyll molecules,
through pheophytin, to quinones. Electrons
then pass from quinones through the
cytochrome bc
1
complex, and back to the
photoreaction center.
■ An alternative path, in green sulfur bacteria,
sends electrons from reduced quinones to
NAD
H11001
.
FIGURE 19–56 Water-splitting activity of the oxygen-evolving com-
plex. Shown here is the process that produces a four-electron oxidiz-
ing agent—believed to be a multinuclear center with several Mn ions—
in the water-splitting complex of PSII. The sequential absorption of
four photons (excitons), each absorption causing the loss of one elec-
tron from the Mn center, produces an oxidizing agent that can remove
four electrons from two molecules of water, producing O
2
. The elec-
trons lost from the Mn center pass one at a time to an oxidized Tyr
residue in a PSII protein, then to P680
H11001
.
Tyr
Exciton Exciton Exciton Exciton
e
H11002
e
H11002
H
H11001
Lumen
e
H11002
e
H11002
Mn
complex
0
Mn
complex
1H11001
Mn
complex
2H11001
Mn
complex
3H11001
Mn
complex
O
2
4H11001
H
H11001
H
H11001
H
H11001
2H
2
O
P680
8885d_c19_690-750 3/1/04 11:32 AM Page 739 mac76 mac76:385_reb:
■ Cyanobacteria and plants have two different
photoreaction centers, arranged in tandem.
■ Plant photosystem I passes electrons from its
excited reaction center, P700, through a series
of carriers to ferredoxin, which then reduces
NADP
H11001
to NADPH.
■ The reaction center of plant photosystem II,
P680, passes electrons to plastoquinone, and
the electrons lost from P680 are replaced by
electrons from H
2
O (electron donors other than
H
2
O are used in other organisms).
■ The light-driven splitting of H
2
O is catalyzed by
a Mn-containing protein complex; O
2
is
produced. The reduced plastoquinone carries
electrons to the cytochrome b
6
f complex; from
here they pass to plastocyanin, and then to
P700 to replace those lost during its
photoexcitation.
■ Electron flow through the cytochrome b
6
f
complex drives protons across the plasma
membrane, creating a proton-motive force that
provides the energy for ATP synthesis by an
ATP synthase.
19.9 ATP Synthesis
by Photophosphorylation
The combined activities of the two plant photosystems
move electrons from water to NADP
H11001
, conserving some
of the energy of absorbed light as NADPH (Fig. 19–49).
Simultaneously, protons are pumped across the thy-
lakoid membrane and energy is conserved as an elec-
trochemical potential. We turn now to the process by
which this proton gradient drives the synthesis of ATP,
the other energy-conserving product of the light-
dependent reactions.
In 1954 Daniel Arnon and
his colleagues discovered that
ATP is generated from ADP
and P
i
during photosynthetic
electron transfer in illuminated
spinach chloroplasts. Support
for these findings came from
the work of Albert Frenkel,
who detected light-dependent
ATP production in pigment-
containing membranous struc-
tures called chromatophores,
derived from photosynthetic
bacteria. Investigators concluded that some of the light
energy captured by the photosynthetic systems of these
organisms is transformed into the phosphate bond en-
ergy of ATP. This process is called photophosphory-
lation, to distinguish it from oxidative phosphorylation
in respiring mitochondria.
A Proton Gradient Couples Electron Flow
and Phosphorylation
Several properties of photosynthetic electron transfer
and photophosphorylation in chloroplasts indicate that
a proton gradient plays the same role as in mitochon-
drial oxidative phosphorylation. (1) The reaction
centers, electron carriers, and ATP-forming enzymes are
located in a proton-impermeable membrane—the thy-
lakoid membrane—which must be intact to support pho-
tophosphorylation. (2) Photophosphorylation can be
uncoupled from electron flow by reagents that promote
the passage of protons through the thylakoid membrane.
(3) Photophosphorylation can be blocked by venturi-
cidin and similar agents that inhibit the formation of ATP
from ADP and P
i
by the mitochondrial ATP synthase
(Table 19–4). (4) ATP synthesis is catalyzed by F
o
F
1
complexes, located on the outer surface of the thylakoid
membranes, that are very similar in structure and func-
tion to the F
o
F
1
complexes of mitochondria.
Electron-transferring molecules in the chain of
carriers connecting PSII and PSI are oriented asym-
metrically in the thylakoid membrane, so photoinduced
electron flow results in the net movement of protons
across the membrane, from
the stromal side to the thy-
lakoid lumen (Fig. 19–57). In
1966 André Jagendorf showed
that a pH gradient across the
thylakoid membrane (alkaline
outside) could furnish the
driving force to generate ATP.
His early observations pro-
vided some of the most im-
portant experimental evi-
dence in support of Mitchell’s
chemiosmotic hypothesis.
Jagendorf incubated chloroplasts, in the dark, in a
pH 4 buffer; the buffer slowly penetrated into the inner
compartment of the thylakoids, lowering their internal
pH. He added ADP and P
i
to the dark suspension of
chloroplasts and then suddenly raised the pH of the
outer medium to 8, momentarily creating a large pH gra-
dient across the membrane. As protons moved out of
the thylakoids into the medium, ATP was generated
from ADP and P
i
. Because the formation of ATP oc-
curred in the dark (with no input of energy from light),
this experiment showed that a pH gradient across the
membrane is a high-energy state that, as in mitochon-
drial oxidative phosphorylation, can mediate the trans-
duction of energy from electron transfer into the chem-
ical energy of ATP.
Chapter 19 Oxidative Phosphorylation and Photophosphorylation740
Daniel Arnon, 1910–1994
André Jagendorf
8885d_c19_690-750 3/1/04 11:32 AM Page 740 mac76 mac76:385_reb:
19.9 ATP Synthesis by Photophosphorylation 741
Stroma
(N side)
PSII
PSI
Lumen
(P side)
Thylakoid
membrane
Light
2H
H11001
NADP
H11001
H11001H
H11001
O
2
H11001 4H
H11001
PQH
2
2H
2
O
ADP H11001 P
i
ATP
2H
H11001
NADPH
Plasto-
cyanin
PQ
CF
O
CF
1
Fd
Cyt b
6
f
complex
Mn
Light
FIGURE 19–57 Proton and electron circuits in thylakoids. Electrons
(blue arrows) move from H
2
O through PSII, through the intermediate
chain of carriers, through PSI, and finally to NADP
H11001
. Protons (red ar-
rows) are pumped into the thylakoid lumen by the flow of electrons
through the carriers linking PSII and PSI, and reenter the stroma
through proton channels formed by the F
o
(designated CF
o
) of ATP
synthase. The F
1
subunit (CF
1
) catalyzes synthesis of ATP.
The Approximate Stoichiometry
of Photophosphorylation Has Been Established
As electrons move from water to NADP
H11001
in plant chloro-
plasts, about 12 H
H11001
move from the stroma to the thy-
lakoid lumen per four electrons passed (that is, per O
2
formed). Four of these protons are moved by the
oxygen-evolving complex, and up to eight by the cy-
tochrome b
6
f complex. The measurable result is a
1,000-fold difference in proton concentration across the
thylakoid membrane (H9004pH H11005 3). Recall that the energy
stored in a proton gradient (the electrochemical poten-
tial) has two components: a proton concentration dif-
ference (H9004pH) and an electrical potential (H9004H9274) due to
charge separation. In chloroplasts, H9004pH is the dominant
component; counterion movement apparently dissipates
most of the electrical potential. In illuminated chloro-
plasts, the energy stored in the proton gradient per mole
of protons is
H9004G H11005 2.3RT H9004pH H11001 Z H9004H9274 H11005H1100217 kJ/mol
so the movement of 12 mol of protons across the
thylakoid membrane represents conservation of about
200 kJ of energy—enough energy to drive the synthe-
sis of several moles of ATP (H9004GH11032H11034 H11005 30.5 kJ/mol). Ex-
perimental measurements yield values of about 3 ATP
per O
2
produced.
At least eight photons must be absorbed to drive
four electrons from H
2
O to NADPH (one photon per
electron at each reaction center). The energy in eight
photons of visible light is more than enough for the syn-
thesis of three molecules of ATP.
ATP synthesis is not the only energy-conserving re-
action of photosynthesis in plants; the NADPH formed
in the final electron transfer is (like its close analog
NADH) also energetically rich. The overall equation for
noncyclic photophosphorylation (a term explained be-
low) is
2H
2
O H11001 8 photons H11001 2NADP
H11001
H11001 ~3ADP H11001 ~3P
i O
n
O
2
H11001 ~3ATP H11001 2NADPH (19–17)
Cyclic Electron Flow Produces ATP but
Not NADPH or O
2
Using an alternative path of light-induced electron flow,
plants can vary the ratio of NADPH to ATP formed in
the light; this path is called cyclic electron flow to dif-
ferentiate it from the normally unidirectional or non-
cyclic electron flow from H
2
O to NADP
H11001
, as discussed
thus far. Cyclic electron flow (Fig. 19–49) involves
only PSI. Electrons passing from P700 to ferredoxin do
not continue to NADP
H11001
, but move back through the
cytochrome b
6
f complex to plastocyanin. The path of
8885d_c19_690-750 3/1/04 11:32 AM Page 741 mac76 mac76:385_reb:
electrons matches that in green sulfur bacteria (Fig.
19–47b). Plastocyanin donates electrons to P700, which
transfers them to ferredoxin when the plant is illumi-
nated. Thus, in the light, PSI can cause electrons to cycle
continuously out of and back into the reaction center of
PSI, each electron propelled around the cycle by the en-
ergy yielded by the absorption of one photon. Cyclic
electron flow is not accompanied by net formation of
NADPH or evolution of O
2
. However, it is accompanied
by proton pumping by the cytochrome b
6
f complex and
by phosphorylation of ADP to ATP, referred to as cyclic
photophosphorylation. The overall equation for cyclic
electron flow and photophosphorylation is simply
light
ADP H11001 P
i
888n ATP H11001 H
2
O
By regulating the partitioning of electrons between
NADP
H11001
reduction and cyclic photophosphorylation, a
plant adjusts the ratio of ATP to NADPH produced in
the light-dependent reactions to match its needs for
these products in the carbon-assimilation reactions and
other biosynthetic processes. As we shall see in Chap-
ter 20, the carbon-assimilation reactions require ATP
and NADPH in the ratio 3:2.
The ATP Synthase of Chloroplasts Is Like That
of Mitochondria
The enzyme responsible for ATP synthesis in chloro-
plasts is a large complex with two functional compo-
nents, CF
o
and CF
1
(C denoting its location in
chloroplasts). CF
o
is a transmembrane proton pore
composed of several integral membrane proteins and is
homologous to mitochondrial F
o
. CF
1
is a peripheral
membrane protein complex very similar in subunit com-
position, structure, and function to mitochondrial F
1
.
Electron microscopy of sectioned chloroplasts
shows ATP synthase complexes as knoblike projections
on the outside (stromal or N) surface of thylakoid mem-
branes; these complexes correspond to the ATP syn-
thase complexes seen to project on the inside (matrix
or N) surface of the inner mitochondrial membrane.
Thus the relationship between the orientation of the
ATP synthase and the direction of proton pumping is
the same in chloroplasts and mitochondria. In both
cases, the F
1
portion of ATP synthase is located on the
more alkaline (N) side of the membrane through which
protons flow down their concentration gradient; the di-
rection of proton flow relative to F
1
is the same in both
cases: P to N (Fig. 19–58).
The mechanism of chloroplast ATP synthase is also
believed to be essentially identical to that of its mito-
chondrial analog; ADP and P
i
readily condense to form
ATP on the enzyme surface, and the release of this
enzyme-bound ATP requires a proton-motive force. Ro-
tational catalysis sequentially engages each of the three
H9252 subunits of the ATP synthase in ATP synthesis, ATP
release, and ADP H11001 P
i
binding (Figs 19–24, 19–25).
Chloroplasts Evolved from Endosymbiotic Bacteria
Like mitochondria, chloroplasts contain their own DNA
and protein-synthesizing machinery. Some of the
polypeptides of chloroplast proteins are encoded by
chloroplast genes and synthesized in the chloroplast;
others are encoded by nuclear genes, synthesized out-
side the chloroplast, and imported (Chapter 27). When
plant cells grow and divide, chloroplasts give rise to new
Chapter 19 Oxidative Phosphorylation and Photophosphorylation742
Mitochondrion
Matrix (N side)
Intermembrane
space (P side)
Bacterium (E. coli)Chloroplast
H
H11001
Thylakoid
lumen (P side)
Stroma (N side)
ATP
Cytosol (N side)
Intermembrane
space (P side)
ATP
ATP
H
H11001
H
H11001
FIGURE 19–58 Comparison of the topology of proton movement and ATP synthase orientation
in the membranes of mitochondria, chloroplasts, and the bacterium E. coli. In each case, orien-
tation of the proton gradient relative to ATP synthase activity is the same.
8885d_c19_690-750 3/1/04 11:32 AM Page 742 mac76 mac76:385_reb:
chloroplasts by division, during which their DNA is repli-
cated and divided between daughter chloroplasts. The
machinery and mechanism for light capture, electron
flow, and ATP synthesis in photosynthetic bacteria are
similar in many respects to those in the chloroplasts of
plants. These observations led to the now widely ac-
cepted hypothesis that the evolutionary progenitors of
modern plant cells were primitive eukaryotes that en-
gulfed photosynthetic bacteria and established stable
endosymbiotic relationships with them (see Fig. 1–36).
Diverse Photosynthetic Organisms Use Hydrogen
Donors Other Than Water
At least half of the photosynthetic activity on Earth
occurs in microorganisms—algae, other photosynthetic
eukaryotes, and photosynthetic bacteria. Cyanobacteria
have PSII and PSI in tandem, and the PSII has an asso-
ciated water-splitting activity resembling that of plants.
However, the other groups of photosynthetic bacteria
have single reaction centers and do not split H
2
O or
produce O
2
. Many are obligate anaerobes and cannot
tolerate O
2
; they must use some compound other than
H
2
O as electron donor. Some photosynthetic bacteria
use inorganic compounds as electron (and hydrogen)
donors. For example, green sulfur bacteria use hydrogen
sulfide:
light
2H
2
S H11001 CO
2
888n (CH
2
O) H11001 H
2
O H11001 2S
These bacteria, instead of producing molecular O
2
, form
elemental sulfur as the oxidation product of H
2
S. (They
further oxidize the S to SO
4
2H11002
.) Other photosynthetic
bacteria use organic compounds such as lactate as elec-
tron donors:
light
2 Lactate H11001 CO
2
888n (CH
2
O) H11001 H
2
O H11001 2 pyruvate
The fundamental similarity of photosynthesis in plants
and bacteria, despite the differences in the electron
donors they employ, becomes more obvious when the
equation of photosynthesis is written in the more gen-
eral form
light
2H
2
D H11001 CO
2
888n (CH
2
O) H11001 H
2
O H11001 2D
in which H
2
D is an electron (and hydrogen) donor and
D is its oxidized form. H
2
D may be water, hydrogen sul-
fide, lactate, or some other organic compound, de-
pending on the species. Most likely, the bacteria that
first developed photosynthetic ability used H
2
S as their
electron source, and only after the later development of
oxygenic photosynthesis (about 2.3 billion years ago)
did oxygen become a significant proportion of the
earth’s atmosphere. With that development, the evolu-
tion of electron-transfer systems that used O
2
as their
ultimate electron acceptor became possible, leading to
the highly efficient energy extraction of oxidative phos-
phorylation.
In Halophilic Bacteria, a Single Protein Absorbs Light
and Pumps Protons to Drive ATP Synthesis
The halophilic (“salt-loving”) bacterium Halobacterium
salinarum, an archaebacterium derived from very an-
cient evolutionary progenitors, traps the energy of sun-
light in a process very different from the photosynthetic
mechanisms we have described so far. This bacterium
lives only in brine ponds and salt lakes (Great Salt Lake
and the Dead Sea, for example), where the high salt
concentration—which can exceed 4 M—results from
water loss by evaporation; indeed, halobacteria cannot
live in NaCl concentrations lower than 3 M. These or-
ganisms are aerobes and normally use O
2
to oxidize
organic fuel molecules. However, the solubility of O
2
is
so low in brine ponds that sometimes oxidative metab-
olism must be supplemented by sunlight as an alterna-
tive source of energy.
The plasma membrane of H. salinarum contains
patches of the light-absorbing pigment bacteriorho-
dopsin, which contains retinal (the aldehyde derivative
of vitamin A; see Fig. 10–21) as a prosthetic group.
When the cells are illuminated, all-trans-retinal bound
to the bacteriorhodopsin absorbs a photon and under-
goes photoisomerization to 13-cis-retinal. The restora-
tion of all-trans-retinal is accompanied by the outward
movement of protons through the plasma membrane.
Bacteriorhodopsin, with only 247 amino acid residues,
is the simplest light-driven proton pump known. The dif-
ference in the three-dimensional structure of bacteri-
orhodopsin in the dark and after illumination (Fig.
19–59a) suggests a pathway by which a concerted se-
ries of proton “hops” could effectively move a proton
across the membrane. The chromophore retinal is bound
through a Schiff-base linkage to the H9255-amino group of a
Lys residue. In the dark, the N of this Schiff base is pro-
tonated; in the light, photoisomerization of retinal low-
ers the pK
a
of this group and it releases its proton to a
nearby Asp residue, triggering a series of proton hops
that ultimately result in the release of a proton at the
outer surface of the membrane (Fig. 19–59b).
The electrochemical potential across the membrane
drives protons back into the cell through a membrane
ATP synthase complex very similar to that of mitochon-
dria and chloroplasts. Thus, when O
2
is limited, halobac-
teria can use light to supplement the ATP synthesized
by oxidative phosphorylation. Halobacteria do not evolve
O
2
, nor do they carry out photoreduction of NADP
H11001
;
their phototransducing machinery is therefore much
simpler than that of cyanobacteria or plants. Neverthe-
less, the proton-pumping mechanism used by this simple
protein may prove to be prototypical for the many other,
more complex, ion pumps. Bacteriorhodopsin ■
19.9 ATP Synthesis by Photophosphorylation 743
8885d_c19_690-750 3/1/04 11:32 AM Page 743 mac76 mac76:385_reb:
FIGURE 19–59 Light-driven proton pumping by bacteriorhodopsin.
(a) Bacteriorhodopsin (M
r
26,000) has seven membrane-spanning H9251
helices (PDB ID 1C8R). The chromophore all-trans-retinal (purple) is
covalently attached via a Schiff base to the H9255-amino group of a Lys
residue deep in the membrane interior. Running through the protein
are a series of Asp and Glu residues and a series of closely associated
water molecules that together provide the transmembrane path for pro-
tons (red arrows). Steps 1 through 5 indicate proton movements,
described below.
(b) In the dark (left panel), the Schiff base is protonated. Illumina-
tion (right panel) photoisomerizes the retinal, forcing subtle confor-
mational changes in the protein that alter the distance between the
Schiff base and its neighboring amino acid residues. Interaction with
these neighbors lowers the pK
a
of the protonated Schiff base, and the
base gives up its proton to a nearby carboxyl group on Asp
85
(step 1
in (a)). This initiates a series of concerted proton hops between water
molecules (see Fig. 2–14) in the interior of the protein, which ends with
2 the release of a proton that was shared by Glu
194
and Glu
204
near
the extracellular surface. 3 The Schiff base reacquires a proton from
Asp
96
, which 4 takes up a proton from the cytosol. 5 Finally, Asp
85
gives up its proton, leading to reprotonation of the Glu
204
-Glu
194
pair.
The system is now ready for another round of proton pumping.
(a)
Cytosol
Asp
96
Asp
85
Arg
82
Glu
204
Retinal
Glu
194
H
+
H+
External
medium
4
3
2
1
5
Thr
89
OH
Asp
85
C
C
O
OO
Protonated
Schiff base
(high pK
a
)
Proton-release complex
(protonated; high pK
a
)
Low pK
a
Retinal
Dark
O
N
H
H11001
Lys
216
Arg
82
Glu
194
C
HO O
Glu
204
CN
H
NH
2
NH
2
H11001 H11001
N
Lys
216
Leu
93
Val
49
Asp
85
C
C
O
OO
Conformational
change lowers pK
a
of Schiff base
pK
a
of proton-
release complex
lowered
Higher pK
a
Light
OH
Arg
82
Tyr
83
Glu
204
C
OO
Glu
194
Proton
release
Proton
transfer
CNH
NH
2
OH
NH
2
H11002
H11002
H11002
H11002
Chapter 19 Oxidative Phosphorylation and Photophosphorylation744
(b)
8885d_c19_690-750 3/1/04 11:32 AM Page 744 mac76 mac76:385_reb:
Chapter 19 Further Reading 745
SUMMARY 19.9 ATP Synthesis
by Photophosphorylation
■ In plants, both the water-splitting reaction and
electron flow through the cytochrome b
6
f
complex are accompanied by proton pumping
across the thylakoid membrane. The
proton-motive force thus created drives ATP
synthesis by a CF
o
CF
1
complex similar to the
mitochondrial F
o
F
1
complex.
■ Flow of electrons through the photosystems
produces NADPH and ATP, in the ratio of about
2:3. A second type of electron flow (cyclic flow)
produces ATP only and allows variability in the
proportions of NADPH and ATP formed.
■ The localization of PSI and PSII between the
granal and stromal lamellae can change and is
indirectly controlled by light intensity,
optimizing the distribution of excitons between
PSI and PSII for efficient energy capture.
■ Chloroplasts, like mitochondria, evolved from
bacteria living endosymbiotically within early
eukaryotic cells. The ATP synthases of
eubacteria, cyanobacteria, mitochondria, and
chloroplasts share a common evolutionary
precursor and a common enzymatic
mechanism.
■ Many photosynthetic microorganisms obtain
electrons for photosynthesis not from water but
from donors such as H
2
S.
Key Terms
chemiosmotic theory 690
nicotinamide nucleotide–linked
dehydrogenases 692
flavoprotein 692
reducing equivalent 693
ubiquinone (coenzyme Q, Q) 693
cytochromes 693
iron-sulfur protein 693
Complex I 696
vectorial metabolism 697
Complex II 698
Complex III 699
cytochrome bc
1
complex 699
Q cycle 700
Complex IV 700
cytochrome oxidase 700
proton-motive force 703
ATP synthase 704
F
1
ATPase 708
rotational catalysis 711
P/O ratio 712
P/2e
H11002
ratio 712
acceptor control 716
mass-action ratio 716
light-dependent reactions 723
light reactions 723
carbon-assimilation
reactions 723
carbon-fixation reaction 723
thylakoid 724
stroma 724
exciton transfer 725
chlorophylls 725
light-harvesting complexes
(LHCs) 725
accessory pigments 728
photosystem 728
photochemical reaction
center 728
light-harvesting (antenna)
molecules 728
photosystem II (PSII) 733
photosystem I (PSI) 733
oxygenic photosynthesis 734
oxygen-evolving complex (water-
splitting complex) 738
photophosphorylation 740
cyclic electron flow 741
noncyclic electron flow 741
cyclic photophosphorylation 742
Terms in bold are defined in the glossary.
Further Reading
History and General Background
Arnon, D.I. (1984) The discovery of photosynthetic phosphoryla-
tion. Trends Biochem. Sci. 9, 258–262.
Beinert, H. (1995) These are the moments when we live! From
Thunberg tubes and manometry to phone, fax and FedEx. In
Selected Topics in the History of Biochemistry: Personal
Recollections, Comprehensive Biochemistry, Vol. 38, Elsevier
Science Publishing Co., Inc., New York.
An engaging personal account of the exciting period when the
biochemistry of respiratory electron transfer was worked out.
Blankenship, R.E. (2002) Molecular Mechanisms of Photosyn-
thesis, Blackwell Science Inc., London.
An intermediate-level discussion of all aspects of photosynthesis.
Gray, M.W., Berger, G., & Lang, B.F. (1999) Mitochondrial evo-
lution. Science 283, 1476–1481.
Compact review of the endosymbiotic origin hypothesis and the
evidence for and against it.
Harold, F.M. (1986) The Vital Force: A Study in Bioenergetics,
W. H. Freeman and Company, New York.
A very readable synthesis of the principles of bioenergetics and
their application to energy transductions.
Heldt, H.-W. (1997) Plant Biochemistry and Molecular Biology,
Oxford University Press, Oxford.
A textbook of plant biochemistry with excellent discussions of
photophosphorylation.
Keilin, D. (1966) The History of Cell Respiration and Cy-
tochrome, Cambridge University Press, London.
An authoritative and absorbing account of the discovery of cy-
tochromes and their roles in respiration, written by the man
who discovered cytochromes.
8885d_c19_690-750 3/1/04 11:32 AM Page 745 mac76 mac76:385_reb:
Chapter 19 Oxidative Phosphorylation and Photophosphorylation746
Mitchell, P. (1979) Keilin’s respiratory chain concept and its
chemiosmotic consequences. Science 206, 1148–1159.
Mitchell’s Nobel lecture, outlining the evolution of the chemios-
motic hypothesis.
Nicholls, D.G. & Ferguson, S.J. (2002) Bioenergetics 3, Acade-
mic Press, Amsterdam.
Up-to-date, comprehensive, well-illustrated treatment of all as-
pects of mitochondrial and chloroplast energy transductions.
A comprehensive, advanced treatise.
Scheffler, I.E. (1999) Mitochondria, Wiley-Liss, New York.
An excellent survey of mitochondrial structure and function.
Slater, E.C. (1987) The mechanism of the conservation of energy
of biological oxidations. Eur. J. Biochem. 166, 489–504.
A clear and critical account of the evolution of the
chemiosmotic model.
OXIDATIVE PHOSPHORYLATION
Respiratory Electron Flow
Babcock, G.T. & Wickstr?m, M. (1992) Oxygen activation and
the conservation of energy in cell respiration. Nature 356, 301–309.
Advanced discussion of the reduction of water and pumping of
protons by cytochrome oxidase.
Brandt, U. (1997) Proton-translocation by membrane-bound
NADH:ubiquinone-oxidoreductase (complex I) through redox-
gated ligand conduction. Biochim. Biophys. Acta 1318, 79–91.
Advanced discussion of models for electron movement through
Complex I.
Brandt, U. & Trumpower, B. (1994) The protonmotive Q cycle
in mitochondria and bacteria. Crit. Rev. Biochem. Mol. Biol. 29,
165–197.
Crofts, A.R. & Berry, E.A. (1998) Structure and function of the
cytochrome bc
1
complex of mitochondria and photosynthetic bac-
teria. Curr. Opin. Struct. Biol. 8, 501–509.
Michel, H., Behr, J., Harrenga, A., & Kannt, A. (1998) Cy-
tochrome c oxidase: structure and spectroscopy. Annu. Rev. Bio-
phys. Biomol. Struct. 27, 329–356.
Advanced review of Complex IV structure and function.
Rottenberg, H. (1998) The generation of proton electrochemical
potential gradient by cytochrome c oxidase. Biochim. Biophys.
Acta 1364, 1–16.
Tielens, A.G.M., Rotte, C., van Hellemond, J.J., & Martin, W.
(2002) Mitochondria as we don’t know them. Trends Biochem.
Sci. 27, 564–572.
Intermediate-level discussion of the many organisms in which
mitochondria do not depend on oxygen as the final electron
donor.
Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T.,
Yamaguchi, H., Shinzawa-Itoh, K., Nakashima, R., Yaono, R.,
& Yoshikawa, S. (1996) The whole structure of the 13-subunit
oxidized cytochrome c oxidase at 2.8 ?. Science 272, 1136–1144.
The solution by x-ray crystallography of the structure of this
huge membrane protein.
Xia, D., Yu, C.-A., Kim, H., Xia, J.-Z., Kachurin, A.M., Zhang,
L., Yu, L., & Deisenhofer, J. (1997) Crystal structure of the cy-
tochrome bc
1
complex from bovine heart mitochondria. Science
277, 60–66.
Report revealing the crystallographic structure of Complex III.
Yankovskaya, V., Horsefield, R., T?rnroth, S., Luna-Chavez,
C., Myoshi, H., Léger, C., Byrne, B., Cecchini, G., & Iwata, S.
(2003) Architecture of succinate dehydrogenase and reactive oxy-
gen species generation. Science 299, 700–704.
Coupling ATP Synthesis to Respiratory Electron Flow
Abrahams, J.P., Leslie, A.G.W., Lutter, R., & Walker, J.E.
(1994) The structure of F
1
-ATPase from bovine heart mitochondria
determined at 2.8 ? resolution. Nature 370, 621–628.
Bianchet, M.A., Hullihen, J., Pedersen, P.L., & Amzel, L.M.
(1998) The 2.80 ? structure of rat liver F
1
-ATPase: configuration
of a critical intermediate in ATP synthesis-hydrolysis. Proc. Natl.
Acad. Sci. USA 95, 11,065–11,070.
Research paper that provided important structural detail in
support of the catalytic mechanism.
Boyer, P.D. (1997) The ATP synthase—a splendid molecular ma-
chine. Annu. Rev. Biochem. 66, 717–749.
An account of the historical development and current state of
the binding-change model, written by its principal architect.
Cabezón, E., Montgomery, M.G., Leslie, A.G.W., & Walker,
J.E. (2003) The structure of bovine F
1
-ATPase in complex with its
regulatory protein IF
1
Nat. Struct. Biol. 10, 744–750.
Hinkle, P.C., Kumar, M.A., Resetar, A., & Harris, D.L. (1991)
Mechanistic stoichiometry of mitochondrial oxidative phosphoryla-
tion. Biochemistry 30, 3576–3582.
A careful analysis of experimental results and theoretical con-
siderations on the question of nonintegral P/O ratios.
Khan, S. (1997) Rotary chemiosmotic machines. Biochim. Bio-
phys. Acta 1322, 86–105.
Detailed review of the structures that underlie proton-driven
rotary motion of ATP synthase and bacterial flagella.
Sambongi, Y., Iko, Y., Tanabe, M., Omote, H., Iwamoto-
Kihara, A., Ueda, I., Yanagida, T., Wada, Y., & Futai, M.
(1999) Mechanical rotation of the c subunit oligomer in ATP
synthase (F
o
F
1
): direct observation. Science 286, 1722–1724.
The experimental evidence for rotation of the entire cylinder of
c subunits in F
o
F
1
.
Stock, D., Leslie, A.G.W., & Walker, J.E. (1999) Molecular
architecture of the rotary motor in ATP synthase. Science 286,
1700–1705.
The first crystallographic view of the F
o
subunit, in the yeast
F
o
F
1
. See also R. H. Fillingame’s editorial comment in the same
issue of Science.
Weber, J. & Senior, A.E. (1997) Catalytic mechanism of F
1
-
ATPase. Biochim. Biophys. Acta 1319, 19–58.
An advanced review of kinetic, structural, and biochemical evi-
dence for the ATP synthase mechanism.
Yasuda, R., Noji, H., Kinosita, K., Jr., & Yoshida, M. (1998)
F
1
-ATPase is a highly efficient molecular motor that rotates with
discrete 120H11034 steps. Cell 93, 1117–1124.
Graphical demonstration of the rotation of ATP synthase.
Regulation of Mitochondrial Oxidative Phosphorylation
Brand, M.D. & Murphy, M.P. (1987) Control of electron flux
through the respiratory chain in mitochondria and cells. Biol. Rev.
Camb. Philos. Soc. 62, 141–193.
An advanced description of respiratory control.
8885d_c19_690-750 3/1/04 11:32 AM Page 746 mac76 mac76:385_reb:
Chapter 19 Further Reading 747
Harris, D.A. & Das, A.M. (1991) Control of mitochondrial ATP
synthesis in the heart. Biochem. J. 280, 561–573.
Advanced discussion of the regulation of ATP synthase by Ca
2H11001
and other factors.
Klingenberg, M. & Huang, S.-G. (1999) Structure and function
of the uncoupling protein from brown adipose tissue. Biochim.
Biophys. Acta 1415, 271–296.
Apoptosis and Mitochondrial Diseases
Kroemer, G. (2003) Mitochondrial control of apoptosis: an intro-
duction. Biochem. Biophys. Res. Commun. 304, 433–435.
A short introduction to a collection of excellent papers in this
journal issue.
McCord, J.M. (2002) Superoxide dismutase in aging and disease:
an overview. Meth. Enzymol. 349, 331–341.
Newmeyer, D.D. & Ferguson-Miller, S. (2003) Mitochondria: re-
leasing power for life and unleashing the machineries of death. Cell
112, 481–490.
Intermediate-level review of the possible roles of mitochondria
in apoptosis.
Schapira, A.H.V. (2002) Primary and secondary defects of the mi-
tochondrial respiratory chain. J. Inher. Metab. Dis. 25, 207–214.
Wallace, D.C. (1999) Mitochondrial disease in man and mouse.
Science 283, 1482–1487.
PHOTOSYNTHESIS
Harvesting Light Energy
Cogdell, R.J., Isaacs, N.W., Howard, T.D., McLuskey, K.,
Fraser, N.J., & Prince, S.M. (1999) How photosynthetic bacteria
harvest solar energy. J. Bacteriol. 181, 3869–3879.
A short, intermediate-level review of the structure and function
of the light-harvesting complex of the purple bacteria and exci-
ton flow to the reaction center.
Green, B.R., Pichersky, E., & Kloppstech, K. (1991) Chloro-
phyll a/b-binding proteins: an extended family. Trends Biochem.
Sci. 16, 181–186.
An intermediate-level description of the proteins that orient
chlorophyll molecules in chloroplasts.
Kargul, J., Nield, J., & Barber, J. (2003) Three-dimensional re-
construction of a light-harvesting Complex I–Photosystem I (LHCI-
PSI) supercomplex from the green alga Chlamydomonas rein-
hardtii. J. Biol. Chem. 278, 16,135–16,141.
Zuber, H. (1986) Structure of light-harvesting antenna complexes
of photosynthetic bacteria, cyanobacteria and red algae. Trends
Biochem. Sci. 11, 414–419.
Light-Driven Electron Flow
Barber, J. (2002) Photosystem II: a multisubunit membrane pro-
tein that oxidizes water. Curr. Opin. Struct. Biol. 12, 523–530.
A short, intermediate-level summary of the structure of PSII.
Barber, J. & Anderson, J.M. (eds) (2002) Photosystem II: mo-
lecular structure and function. Philos. Trans. R. Soc. (Biol. Sci.)
357, 1321–1512.
A collection of 16 papers on photosystem II.
Chitnis, P.R. (2001) Photosystem I: function and physiology.
Annu. Rev. Plant Physiol. Plant Mol. Biol. 52, 593–626.
An advanced and lengthy review.
Deisenhofer, J. & Michel, H. (1991) Structures of bacterial pho-
tosynthetic reaction centers. Annu. Rev. Cell Biol. 7, 1–23.
Description of the structure of the reaction center of purple
bacteria and implications for the function of bacterial and plant
reaction centers.
Heathcote, P., Fyfe, P.K., & Jones, M.R. (2002) Reaction cen-
tres: the structure and evolution of biological solar power. Trends.
Biochem Sci. 27, 79–87.
Intermediate-level review of photosystems I and II.
Huber, R. (1990) A structural basis of light energy and electron
transfer in biology. Eur. J. Biochem. 187, 283–305.
Huber’s Nobel lecture, describing the physics and chemistry of
phototransductions; an exceptionally clear and well-illustrated
discussion, based on crystallographic studies of reaction centers.
Jagendorf, A.T. (1967) Acid-base transitions and phosphorylation
by chloroplasts. Fed. Proc. 26, 1361–1369.
Report of the classic experiment establishing the ability of a
proton gradient to drive ATP synthesis in the dark.
Jordan, P., Fromme, P., Witt, H.T., Klukas, O., Saenger, W., &
Krauss, N. (2001) Three-dimensional structure of cyanobacterial
photosystem I at 2.5 ?. Nature 411, 909–917.
Kamiya, N. & Shen, J.-R. (2003) Crystal structure of oxygen-
evolving photosystem II from Thermosynechococcus vulcanus at
3.7 ? resolution. Proc. Natl. Acad. Sci. USA 100, 98–103.
Kargul, J., Nield, J., & Barber, J. (2003) Three-dimensional re-
construction of a light-harvesting complex I–photosystem I
(LHCI-PSI) supercomplex from the green alga Chlamydomonas
reinhardtii: insights into light harvesting for PSI. J. Biol. Chem.
278, 16,135–16,141.
Leister, D. (2003) Chloroplast research in the genomic age.
Trends Genet. 19, 47–56.
Zouni, A., Witt, H.-T., Kern, J., Fromme, P., Krauss, N.,
Saenger, W., & Orth, P. (2001) Crystal structure of photosystem
II from Synechococcus elongates at 3.8 ? resolution. Nature 409,
739–743.
Water-Splitting Complex
R?gner, M., Boekema, E.J., & Barber, J. (1996) How does pho-
tosystem 2 split water? The structural basis of efficient energy
conversion. Trends Biochem. Sci. 21, 44–49.
Szalai, V.A. & Brudvig, G.W. (1998) How plants produce dioxy-
gen. Am. Sci. 86, 542–551.
A well-illustrated introduction to the oxygen-evolving complex
of plants.
Bacteriorhodopsin
Luecke, H. (2000) Atomic resolution structures of bacteriorhodopsin
photocycle intermediates: the role of discrete water molecules in
the function of this light-driven ion pump. Biochim. Biophys. Acta
1460, 133–156.
Advanced review of a proton pump that employs an internal
chain of water molecules.
Luecke, H., Schobert, B., Richter, H.-T., Cartailler, J.-P., &
Lanyi, J.K. (1999) Structural changes in bacteriorhodopsin during
ion transport at 2 angstrom resolution. Science 286, 255–264.
This paper, accompanied by an editorial comment in the same
Science issue, describes the model for H
H11001
translocation by pro-
ton hopping.
8885d_c19_690-750 3/1/04 11:32 AM Page 747 mac76 mac76:385_reb:
Chapter 19 Oxidative Phosphorylation and Photophosphorylation748
Problems
1. Oxidation-Reduction Reactions The NADH dehy-
drogenase complex of the mitochondrial respiratory chain pro-
motes the following series of oxidation-reduction reactions, in
which Fe
3H11001
and Fe
2H11001
represent the iron in iron-sulfur cen-
ters, Q is ubiquinone, QH
2
is ubiquinol, and E is the enzyme:
(1) NADH H11001 H
H11001
H11001 E-FMN 88n NAD
H11001
H11001 E-FMNH
2
(2) E-FMNH
2
H11001 2Fe
3H11001
88n E-FMN H11001 2Fe
2H11001
H11001 2H
H11001
(3) 2Fe
2H11001
H11001 2H
H11001
H11001 Q 88n 2Fe
3H11001
H11001 QH
2
Sum: NADH H11001 H
H11001
H11001 Q 88n NAD
H11001
H11001 QH
2
For each of the three reactions catalyzed by the NADH de-
hydrogenase complex, identify (a) the electron donor, (b) the
electron acceptor, (c) the conjugate redox pair, (d) the re-
ducing agent, and (e) the oxidizing agent.
2. All Parts of Ubiquinone Have a Function In elec-
tron transfer, only the quinone portion of ubiquinone under-
goes oxidation-reduction; the isoprenoid side chain remains
unchanged. What is the function of this chain?
3. Use of FAD Rather Than NAD
H11545
in Succinate Oxi-
dation All the dehydrogenases of glycolysis and the citric
acid cycle use NAD
H11001
(EH11032H11034 for NAD
H11001
/NADH is H110020.32 V) as
electron acceptor except succinate dehydrogenase, which
uses covalently bound FAD (EH11032H11034 for FAD/FADH
2
in this en-
zyme is 0.050 V). Suggest why FAD is a more appropriate elec-
tron acceptor than NAD
H11001
in the dehydrogenation of succinate,
based on the EH11032H11034 values of fumarate/succinate (EH11032H11034 H11005 0.031),
NAD
H11001
/NADH, and the succinate dehydrogenase FAD/FADH
2
.
4. Degree of Reduction of Electron Carriers in the
Respiratory Chain The degree of reduction of each carrier
in the respiratory chain is determined by conditions in the mi-
tochondrion. For example, when NADH and O
2
are abundant,
the steady-state degree of reduction of the carriers decreases
as electrons pass from the substrate to O
2
. When electron
transfer is blocked, the carriers before the block become more
reduced and those beyond the block become more oxidized
(see Fig. 19–6). For each of the conditions below, predict the
state of oxidation of ubiquinone and cytochromes b, c
1
, c, and
a H11001 a
3
.
(a) Abundant NADH and O
2
, but cyanide added
(b) Abundant NADH, but O
2
exhausted
(c) Abundant O
2
, but NADH exhausted
(d) Abundant NADH and O
2
5. Effect of Rotenone and Antimycin A on Electron
Transfer Rotenone, a toxic natural product from plants,
strongly inhibits NADH dehydrogenase of insect and fish mi-
tochondria. Antimycin A, a toxic antibiotic, strongly inhibits
the oxidation of ubiquinol.
(a) Explain why rotenone ingestion is lethal to some in-
sect and fish species.
(b) Explain why antimycin A is a poison.
(c) Given that rotenone and antimycin A are equally
effective in blocking their respective sites in the electron-
transfer chain, which would be a more potent poison? Explain.
6. Uncouplers of Oxidative Phosphorylation In normal
mitochondria the rate of electron transfer is tightly coupled
to the demand for ATP. When the rate of use of ATP is rela-
tively low, the rate of electron transfer is low; when demand
for ATP increases, electron-transfer rate increases. Under
these conditions of tight coupling, the number of ATP mole-
cules produced per atom of oxygen consumed when NADH
is the electron donor—the P/O ratio—is about 2.5.
(a) Predict the effect of a relatively low and a relatively
high concentration of uncoupling agent on the rate of elec-
tron transfer and the P/O ratio.
(b) Ingestion of uncouplers causes profuse sweating and
an increase in body temperature. Explain this phenomenon
in molecular terms. What happens to the P/O ratio in the pres-
ence of uncouplers?
(c) The uncoupler 2,4-dinitrophenol was once prescribed
as a weight-reducing drug. How could this agent, in principle,
serve as a weight-reducing aid? Uncoupling agents are no
longer prescribed, because some deaths occurred following
their use. Why might the ingestion of uncouplers lead to death?
7. Effects of Valinomycin on Oxidative Phosphoryla-
tion When the antibiotic valinomycin is added to actively
respiring mitochondria, several things happen: the yield of
ATP decreases, the rate of O
2
consumption increases, heat is
released, and the pH gradient across the inner mitochon-
drial membrane increases. Does valinomycin act as an uncou-
pler or an inhibitor of oxidative phosphorylation? Explain the
experimental observations in terms of the antibiotic’s ability to
transfer K
H11001
ions across the inner mitochondrial membrane.
8. Mode of Action of Dicyclohexylcarbodiimide
(DCCD) When DCCD is added to a suspension of tightly
coupled, actively respiring mitochondria, the rate of electron
transfer (measured by O
2
consumption) and the rate of ATP
production dramatically decrease. If a solution of 2,4-dinitro-
phenol is now added to the preparation, O
2
consumption re-
turns to normal but ATP production remains inhibited.
(a) What process in electron transfer or oxidative phos-
phorylation is affected by DCCD?
(b) Why does DCCD affect the O
2
consumption of mi-
tochondria? Explain the effect of 2,4-dinitrophenol on the in-
hibited mitochondrial preparation.
(c) Which of the following inhibitors does DCCD most
resemble in its action: antimycin A, rotenone, or oligomycin?
9. Compartmentalization of Citric Acid Cycle Compo-
nents Isocitrate dehydrogenase is found only in the mito-
chondrion, but malate dehydrogenase is found in both the
cytosol and mitochondrion. What is the role of cytosolic
malate dehydrogenase?
10. The Malate–H9251-Ketoglutarate Transport System
The transport system that conveys malate and H9251-ketoglu-
tarate across the inner mitochondrial membrane (see Fig.
19–27) is inhibited by n-butylmalonate. Suppose n-butyl-
malonate is added to an aerobic suspension of kidney cells
using glucose exclusively as fuel. Predict the effect of this in-
hibitor on (a) glycolysis, (b) oxygen consumption, (c) lactate
formation, and (d) ATP synthesis.
11. Cellular ADP Concentration Controls ATP Forma-
tion Although both ADP and P
i
are required for the syn-
8885d_c19_690-750 3/1/04 11:32 AM Page 748 mac76 mac76:385_reb:
749Chapter 19 Problems
thesis of ATP, the rate of synthesis depends mainly on the
concentration of ADP, not P
i
. Why?
12. The Pasteur Effect When O
2
is added to an anaero-
bic suspension of cells consuming glucose at a high rate, the
rate of glucose consumption declines greatly as the O
2
is used
up, and accumulation of lactate ceases. This effect, first ob-
served by Louis Pasteur in the 1860s, is characteristic of most
cells capable of both aerobic and anaerobic glucose catabolism.
(a) Why does the accumulation of lactate cease after O
2
is added?
(b) Why does the presence of O
2
decrease the rate of
glucose consumption?
(c) How does the onset of O
2
consumption slow down
the rate of glucose consumption? Explain in terms of specific
enzymes.
13. Respiration-Deficient Yeast Mutants and Ethanol
Production Respiration-deficient yeast mutants (p
H11002
; “pe-
tites”) can be produced from wild-type parents by treatment
with mutagenic agents. The mutants lack cytochrome oxi-
dase, a deficit that markedly affects their metabolic behavior.
One striking effect is that fermentation is not suppressed by
O
2
—that is, the mutants lack the Pasteur effect (see Prob-
lem 12). Some companies are very interested in using these
mutants to ferment wood chips to ethanol for energy use. Ex-
plain the advantages of using these mutants rather than wild-
type yeast for large-scale ethanol production. Why does the
absence of cytochrome oxidase eliminate the Pasteur effect?
14. How Many Protons in a Mitochondrion? Electron
transfer translocates protons from the mitochondrial matrix
to the external medium, establishing a pH gradient across the
inner membrane (outside more acidic than inside). The ten-
dency of protons to diffuse back into the matrix is the driving
force for ATP synthesis by ATP synthase. During oxidative
phosphorylation by a suspension of mitochondria in a medium
of pH 7.4, the pH of the matrix has been measured as 7.7.
(a) Calculate [H
H11001
] in the external medium and in the
matrix under these conditions.
(b) What is the outside-to-inside ratio of [H
H11001
]? Comment
on the energy inherent in this concentration difference. (Hint:
See Eqn 11–3, p. 398.)
(c) Calculate the number of protons in a respiring liver
mitochondrion, assuming its inner matrix compartment is a
sphere of diameter 1.5 H9262m.
(d) From these data, is the pH gradient alone sufficient
to generate ATP?
(e) If not, suggest how the necessary energy for syn-
thesis of ATP arises.
15. Rate of ATP Turnover in Rat Heart Muscle Rat
heart muscle operating aerobically fills more than 90% of its
ATP needs by oxidative phosphorylation. Each gram of tis-
sue consumes O
2
at the rate of 10.0 H9262mol/min, with glucose
as the fuel source.
(a) Calculate the rate at which the heart muscle con-
sumes glucose and produces ATP.
(b) For a steady-state concentration of ATP of 5.0 H9262mol/g
of heart muscle tissue, calculate the time required (in sec-
onds) to completely turn over the cellular pool of ATP. What
does this result indicate about the need for tight regulation
of ATP production? (Note: Concentrations are expressed as
micromoles per gram of muscle tissue because the tissue is
mostly water.)
16. Rate of ATP Breakdown in Flight Muscle ATP pro-
duction in the flight muscle of the fly Lucilia sericata results
almost exclusively from oxidative phosphorylation. During
flight, 187 ml of O
2
/hrjg of body weight is needed to main-
tain an ATP concentration of 7.0 H9262mol/g of flight muscle.
Assuming that flight muscle makes up 20% of the weight of
the fly, calculate the rate at which the flight-muscle ATP pool
turns over. How long would the reservoir of ATP last in the
absence of oxidative phosphorylation? Assume that reducing
equivalents are transferred by the glycerol 3-phosphate shut-
tle and that O
2
is at 25 H11034C and 101.3 kPa (1 atm).
17. Transmembrane Movement of Reducing Equiva-
lents Under aerobic conditions, extramitochondrial NADH
must be oxidized by the mitochondrial electron-transfer
chain. Consider a preparation of rat hepatocytes containing
mitochondria and all the cytosolic enzymes. If [4-
3
H]NADH is
introduced, radioactivity soon appears in the mitochondrial
matrix. However, if [7-
14
C]NADH is introduced, no radioac-
tivity appears in the matrix. What do these observations re-
veal about the oxidation of extramitochondrial NADH by the
electron-transfer chain?
18. NAD Pools and Dehydrogenase Activities Although
both pyruvate dehydrogenase and glyceraldehyde 3-phosphate
dehydrogenase use NAD
H11001
as their electron acceptor, the two
enzymes do not compete for the same cellular NAD pool. Why?
19. Photochemical Efficiency of Light at Different
Wavelengths The rate of photosynthesis, measured by O
2
production, is higher when a green plant is illuminated with
light of wavelength 680 nm than with light of 700 nm. How-
ever, illumination by a combination of light of 680 nm and
700 nm gives a higher rate of photosynthesis than light of ei-
ther wavelength alone. Explain.
20. Balance Sheet for Photosynthesis In 1804 Theodore
de Saussure observed that the total weights of oxygen and
dry organic matter produced by plants is greater than the
weight of carbon dioxide consumed during photosynthesis.
Where does the extra weight come from?
21. Role of H
2
S in Some Photosynthetic Bacteria Il-
luminated purple sulfur bacteria carry out photosynthesis in
the presence of H
2
O and
14
CO
2
, but only if H
2
S is added and
O
2
is absent. During the course of photosynthesis, measured
by formation of [
14
C]carbohydrate, H
2
S is converted to ele-
mental sulfur, but no O
2
is evolved. What is the role of the
conversion of H
2
S to sulfur? Why is no O
2
evolved?
22. Boosting the Reducing Power of Photosystem I by
Light Absorption When photosystem I absorbs red light
at 700 nm, the standard reduction potential of P700 changes
O
14
C
H
N
NH
2
R
[7-
14
C]NADH
O
3
H
C
N
3
H
NH
2
R
[4-
3
H]NADH
H
8885d_c19_690-750 3/1/04 11:32 AM Page 749 mac76 mac76:385_reb:
Chapter 19 Oxidative Phosphorylation and Photophosphorylation750
from 0.40 V to about H110021.2 V. What fraction of the absorbed
light is trapped in the form of reducing power?
23. Limited ATP Synthesis in the Dark In a laboratory
experiment, spinach chloroplasts are illuminated in the ab-
sence of ADP and P
i
, then the light is turned off and ADP
and P
i
are added. ATP is synthesized for a short time in the
dark. Explain this finding.
24. Mode of Action of the Herbicide DCMU When
chloroplasts are treated with 3-(3,4-dichlorophenyl)-1,1-
dimethylurea (DCMU, or diuron), a potent herbicide, O
2
evo-
lution and photophosphorylation cease. Oxygen evolution, but
not photophosphorylation, can be restored by addition of an
external electron acceptor, or Hill reagent. How does DCMU
act as a weed killer? Suggest a location for the inhibitory ac-
tion of this herbicide in the scheme shown in Figure 19–49.
Explain.
25. Bioenergetics of Photophosphorylation The steady-
state concentrations of ATP, ADP, and P
i
in isolated spinach
chloroplasts under full illumination at pH 7.0 are 120.0, 6.0,
and 700.0 H9262m, respectively.
(a) What is the free-energy requirement for the syn-
thesis of 1 mol of ATP under these conditions?
(b) The energy for ATP synthesis is furnished by light-
induced electron transfer in the chloroplasts. What is the min-
imum voltage drop necessary (during transfer of a pair of
electrons) to synthesize ATP under these conditions? (You
may need to refer to Eqn 13–6, p. 510.)
26. Light Energy for a Redox Reaction Suppose you
have isolated a new photosynthetic microorganism that oxi-
dizes H
2
S and passes the electrons to NAD
H11001
. What wavelength
of light would provide enough energy for H
2
S to reduce NAD
H11001
under standard conditions? Assume 100% efficiency in the
photochemical event, and use EH11032H11034 of H11002243 mV for H
2
S and
H11002320 mV for NAD
H11001
. See Figure 19–39 for energy equivalents
of wavelengths of light.
27. Equilibrium Constant for Water-Splitting Reac-
tions The coenzyme NADP
H11001
is the terminal electron ac-
ceptor in chloroplasts, according to the reaction
2H
2
O H11001 2NADP
H11001
On 2NADPH H11001 2H
H11001
H11001 O
2
Use the information in Table 19–2 to calculate the equilib-
rium constant for this reaction at 25 H11034C. (The relationship
between KH11032
eq
and H9004GH11032H11034 is discussed on p. 492.) How can the
chloroplast overcome this unfavorable equilibrium?
28. Energetics of Phototransduction During photosyn-
thesis, eight photons must be absorbed (four by each photo-
system) for every O
2
molecule produced:
2H
2
O H11001 2NADP
H11001
H11001 8 photons 88n 2NADPH H11001 2H
H11001
H11001 O
2
Assuming that these photons have a wavelength of 700 nm
(red) and that the absorption and use of light energy are 100%
efficient, calculate the free-energy change for the process.
29. Electron Transfer to a Hill Reagent Isolated spin-
ach chloroplasts evolve O
2
when illuminated in the presence
of potassium ferricyanide (a Hill reagent), according to the
equation
2H
2
O H11001 4Fe
3H11001
88n O
2
H11001 4H
H11001
H11001 4Fe
2H11001
where Fe
3H11001
represents ferricyanide and Fe
2H11001
, ferrocyanide.
Is NADPH produced in this process? Explain.
30. How Often Does a Chlorophyll Molecule Absorb a
Photon? The amount of chlorophyll a (M
r
892) in a spinach
leaf is about 20 H9262g/cm
2
of leaf. In noonday sunlight (average
energy 5.4 J/cm
2
jmin), the leaf absorbs about 50% of the ra-
diation. How often does a single chlorophyll molecule absorb
a photon? Given that the average lifetime of an excited chloro-
phyll molecule in vivo is 1 ns, what fraction of the chlorophyll
molecules is excited at any one time?
31. Effect of Monochromatic Light on Electron Flow
The extent to which an electron carrier is oxidized or reduced
during photosynthetic electron transfer can sometimes be ob-
served directly with a spectrophotometer. When chloroplasts
are illuminated with 700 nm light, cytochrome f, plastocyanin,
and plastoquinone are oxidized. When chloroplasts are illu-
minated with 680 nm light, however, these electron carriers
are reduced. Explain.
32. Function of Cyclic Photophosphorylation When
the [NADPH]/[NADP
H11001
] ratio in chloroplasts is high, photo-
phosphorylation is predominantly cyclic (see Fig. 19–49). Is
O
2
evolved during cyclic photophosphorylation? Is NADPH
produced? Explain. What is the main function of cyclic photo-
phosphorylation?
8885d_c19_690-750 3/1/04 11:32 AM Page 750 mac76 mac76:385_reb:
chapter
W
e have now reached a turning point in our study of
cellular metabolism. Thus far in Part II we have de-
scribed how the major metabolic fuels—carbohydrates,
fatty acids, and amino acids—are degraded through con-
verging catabolic pathways to enter the citric acid cy-
cle and yield their electrons to the respiratory chain,
and how this exergonic flow of electrons to oxygen is
coupled to the endergonic synthesis of ATP. We now
turn to anabolic pathways, which use chemical energy
in the form of ATP and NADH or NADPH to synthesize
cellular components from simple precursor molecules.
Anabolic pathways are generally reductive rather than
oxidative. Catabolism and anabolism proceed simulta-
neously in a dynamic steady state, so the energy-
yielding degradation of cellular components is counter-
balanced by biosynthetic processes, which create and
maintain the intricate orderliness of living cells.
Plants must be especially versatile in their handling
of carbohydrates, for several reasons. First, plants are
autotrophs, able to convert inorganic carbon (as CO
2
)
into organic compounds. Second, biosynthesis occurs
primarily in plastids, membrane-bounded organelles
unique to plants, and the movement of intermediates be-
tween cellular compartments is an important aspect of
metabolism. Third, plants are not motile: they cannot
move to find better supplies of water, sunlight, or nutri-
ents. They must have sufficient metabolic flexibility to
allow them to adapt to changing conditions in the place
where they are rooted. Finally, plants have thick cell walls
made of carbohydrate polymers, which must be assem-
bled outside the plasma membrane and which constitute
a significant proportion of the cell’s carbohydrate.
The chapter begins with a description of the process
by which CO
2
is assimilated into trioses and hexoses,
then considers photorespiration, an important side re-
action during CO
2
fixation, and the ways in which cer-
tain plants avoid this side reaction. We then look at how
the biosynthesis of sucrose (for sugar transport) and
starch (for energy storage) is accomplished by mech-
anisms analogous to those employed by animal cells to
make glycogen. The next topic is the synthesis of the
cellulose of plant cell walls and the peptidoglycan of bac-
terial cell walls, illustrating the problems of energy-
dependent biosynthesis outside the plasma membrane.
Finally, we discuss how the various pathways that share
pools of common intermediates are segregated within
organelles yet integrated with one another.
20.1 Photosynthetic Carbohydrate Synthesis
The synthesis of carbohydrates in animal cells always
employs precursors having at least three carbons, all of
which are less oxidized than the carbon in CO
2
. Plants
20
751
CARBOHYDRATE BIOSYNTHESIS
IN PLANTS AND BACTERIA
20.1 Photosynthetic Carbohydrate Synthesis 751
20.2 Photorespiration and the C
4
and CAM
Pathways 766
20.3 Biosynthesis of Starch and Sucrose 771
20.4 Synthesis of Cell Wall Polysaccharides: Plant
Cellulose and Bacterial Peptidoglycan 775
20.5 Integration of Carbohydrate Metabolism in the
Plant Cell 780
. . . the discovery of the long-lived isotope of carbon,
carbon-14, by Samuel Ruben and Martin Kamen in 1940
provided the ideal tool for the tracing of the route along
which carbon dioxide travels on its way to carbohydrate.
—Melvin Calvin, Nobel Address, 1961
8885d_c20_751–786 2/18/04 1:56 PM Page 751 mac76 mac76:385_reb:
and photosynthetic microorganisms, by contrast, can
synthesize carbohydrates from CO
2
and water, reducing
CO
2
at the expense of the energy and reducing power
furnished by the ATP and NADPH that are generated
by the light-dependent reactions of photosynthesis (Fig.
20–1). Plants (and other autotrophs) can use CO
2
as the
sole source of the carbon atoms required for the biosyn-
thesis of cellulose and starch, lipids and proteins, and
the many other organic components of plant cells. By
contrast, heterotrophs cannot bring about the net re-
duction of CO
2
to achieve a net synthesis of glucose.
Green plants contain in their chloroplasts unique
enzymatic machinery that catalyzes the conversion of
CO
2
to simple (reduced) organic compounds, a process
called CO
2
assimilation. This process has also been
called CO
2
fixation or carbon fixation, but we re-
serve these terms for the specific reaction in which CO
2
is incorporated (fixed) into a three-carbon organic com-
pound, the triose phosphate 3-
phosphoglycerate. This simple
product of photosynthesis is
the precursor of more com-
plex biomolecules, including
sugars, polysaccharides, and
the metabolites derived from
them, all of which are synthe-
sized by metabolic pathways
similar to those of animal tis-
sues. Carbon dioxide is assim-
ilated via a cyclic pathway, its
key intermediates constantly
regenerated. The pathway
was elucidated in the early
1950s by Melvin Calvin, Andrew Benson, and James A.
Bassham, and is often called the Calvin cycle or, more
descriptively, the photosynthetic carbon reduction
cycle.
Carbohydrate metabolism is more complex in plant
cells than in animal cells or in nonphotosynthetic mi-
croorganisms. In addition to the universal pathways of
glycolysis and gluconeogenesis, plants have the unique
reaction sequences for reduction of CO
2
to triose phos-
phates and the associated reductive pentose phosphate
pathway—all of which must be coordinately regulated
to ensure proper allocation of carbon to energy pro-
duction and synthesis of starch and sucrose. Key en-
zymes are regulated, as we shall see, by (1) reduction
of disulfide bonds by electrons flowing from photosys-
tem I and (2) changes in pH and Mg
2H11001
concentration
that result from illumination. When we look at other as-
pects of plant carbohydrate metabolism, we also find en-
zymes that are modulated by (3) conventional allosteric
regulation by one or more metabolic intermediates and
(4) covalent modification (phosphorylation).
Plastids Are Organelles Unique to
Plant Cells and Algae
Most of the biosynthetic activities in plants (including
CO
2
assimilation) occur in plastids, a family of self-
reproducing organelles bounded by a double membrane
and containing a small genome that encodes some of
their proteins. Most proteins destined for plastids are
encoded in nuclear genes, which are transcribed and
translated like other nuclear genes; then the proteins
are imported into plastids. Plastids reproduce by binary
fission, replicating their genome (a single circular DNA
molecule) and using their own enzymes and ribosomes
to synthesize the proteins encoded by that genome.
Chloroplasts (see Fig. 19–38) are the sites of CO
2
as-
similation. The enzymes for this process are contained
in the stroma, the soluble phase bounded by the inner
chloroplast membrane. Amyloplasts are colorless
plastids (that is, they lack chlorophyll and other pig-
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria752
Sucrose
(transport)
Starch
(storage)
Cellulose
(cell wall)
Hexose phosphates
Pentose phosphates
Metabolic
intermediates
DNA
RNA
Protein
Lipid
Triose phosphates
ADP,
NADP
H11001
ATP,
NADPH
CO
2
, H
2
O
Light-dependent
reactions of
photosynthesis
FIGURE 20–1 Assimilation of CO
2
into biomass in plants. The light-
driven synthesis of ATP and NADPH, described in Chapter 19, pro-
vides energy and reducing power for the fixation of CO
2
into trioses,
from which all the carbon-containing compounds of the plant cell are
synthesized. The processes shown with red arrows are the focus of this
chapter.
Melvin Calvin,
1911–1997
8885d_c20_752 2/20/04 2:45 PM Page 752 mac76 mac76:385_reb:
ments found in chloroplasts). They have no internal
membranes analogous to the photosynthetic mem-
branes (thylakoids) of chloroplasts, and in plant tissues
rich in starch these plastids are packed with starch gran-
ules (Fig. 20–2). Chloroplasts can be converted to pro-
plastids by the loss of their internal membranes and
chlorophyll, and proplastids are interconvertible with
amyloplasts (Fig. 20–3). In turn, both amyloplasts and
proplastids can develop into chloroplasts. The relative
proportions of the plastid types depend on the type of
plant tissue and on the intensity of illumination. Cells of
green leaves are rich in chloroplasts, whereas amylo-
plasts dominate in nonphotosynthetic tissues that store
starch in large quantities, such as potato tubers.
The inner membranes of all types of plastids are im-
permeable to polar and charged molecules. Traffic
across these membranes is mediated by sets of specific
transporters.
Carbon Dioxide Assimilation Occurs in Three Stages
The first stage in the assimilation of CO
2
into biomole-
cules (Fig. 20–4) is the carbon-fixation reaction:
condensation of CO
2
with a five-carbon acceptor, ribu-
lose 1,5-bisphosphate, to form two molecules of 3-
phosphoglycerate. In the second stage, the 3-phos-
phoglycerate is reduced to triose phosphates. Overall,
three molecules of CO
2
are fixed to three molecules of
ribulose 1,5-bisphosphate to form six molecules of glyc-
eraldehyde 3-phosphate (18 carbons) in equilibrium
with dihydroxyacetone phosphate. In the third stage,
five of the six molecules of triose phosphate (15 car-
bons) are used to regenerate three molecules of ribu-
lose 1,5-bisphosphate (15 carbons), the starting mate-
rial. The sixth molecule of triose phosphate, the net
product of photosynthesis, can be used to make hex-
oses for fuel and building materials, sucrose for trans-
port to nonphotosynthetic tissues, or starch for storage.
Thus the overall process is cyclical, with the continuous
conversion of CO
2
to triose and hexose phosphates.
Fructose 6-phosphate is a key intermediate in stage 3
of CO
2
assimilation; it stands at a branch point, leading
either to regeneration of ribulose 1,5-bisphosphate or to
synthesis of starch. The pathway from hexose phos-
phate to pentose bisphosphate involves many of the
same reactions used in animal cells for the conversion
of pentose phosphates to hexose phosphates during
the nonoxidative phase of the pentose phosphate
pathway (see Fig. 14–22). In the photosynthetic as-
similation of CO
2
, essentially the same set of reactions
operates in the other direction, converting hexose phos-
phates to pentose phosphates. This reductive pentose
phosphate cycle uses the same enzymes as the oxida-
tive pathway, and several more enzymes that make the
reductive cycle irreversible. All 13 enzymes of the path-
way are in the chloroplast stroma.
20.1 Photosynthetic Carbohydrate Synthesis 753
FIGURE 20–2 Amyloplasts filled with starch (dark granules) are
stained with iodine in this section of Ranunculus root cells. Starch
granules in various tissues range from 1 to 100 H9262m in diameter.
FIGURE 20–3 Plastids: their origins and interconversions. All types
of plastids are bounded by a double membrane, and some (notably
the mature chloroplast) have extensive internal membranes. The in-
ternal membranes can be lost (when a mature chloroplast becomes a
proplastid) and resynthesized (as a proplastid gives rise to a pregranal
plastid and then a mature chloroplast). Proplastids in nonphotosyn-
thetic tissues (such as root) give rise to amyloplasts, which contain
large quantities of starch. All plant cells have plastids, and these or-
ganelles are the site of other important processes, including the syn-
thesis of essential amino acids, thiamine, pyridoxal phosphate, flavins,
and vitamins A, C, E, and K.
Chloroplast
ProplastidPregranal
plastid
Amyloplast
8885d_c20_751–786 2/18/04 1:56 PM Page 753 mac76 mac76:385_reb:
Stage 1: Fixation of CO
2
into 3-Phosphoglycerate An im-
portant clue to the nature of the CO
2
-assimilation mech-
anisms in photosynthetic organisms came in the late
1940s. Calvin and his associates illuminated a suspen-
sion of green algae in the presence of radioactive car-
bon dioxide (
14
CO
2
) for just a few seconds, then quickly
killed the cells, extracted their contents, and with the
help of chromatographic methods searched for the
metabolites in which the labeled carbon first appeared.
The first compound that became labeled was 3-phos-
phoglycerate, with the
14
C predominantly located in
the carboxyl carbon atom. These experiments strongly
suggested that 3-phosphoglycerate is an early interme-
diate in photosynthesis. The many plants in which this
three-carbon compound is the first intermediate are
called C
3
plants, in contrast with the C
4
plants de-
scribed below.
The enzyme that catalyzes incorporation of CO
2
into
an organic form is ribulose 1,5-bisphosphate carbox-
ylase/oxygenase, a name mercifully shortened to ru-
bisco. As a carboxylase, rubisco catalyzes the covalent
attachment of CO
2
to the five-carbon sugar ribulose 1,5-
bisphosphate and cleavage of the unstable six-carbon in-
termediate to form two molecules of 3-phosphoglycerate,
one of which bears the carbon introduced as CO
2
in its
carboxyl group (Fig. 20–4). The enzyme’s oxygenase ac-
tivity is discussed in Section 20.2.
Plant rubisco, the crucial enzyme in the production
of biomass from CO
2
, has a complex structure (Fig.
20–5a), with eight identical large subunits (M
r
53,000;
encoded in the chloroplast genome, or plastome), each
containing a catalytic site, and eight identical small sub-
units (M
r
14,000; encoded in the nuclear genome) of
uncertain function. The rubisco of photosynthetic bac-
teria is simpler in structure, having two subunits that in
many respects resemble the large subunits of the plant
enzyme (Fig. 20–5b). This similarity is consistent with
the endosymbiont hypothesis for the origin of chloro-
plasts (p. 35). The plant enzyme has an exceptionally
low turnover number; only three molecules of CO
2
are
fixed per second per molecule of rubisco at 25 H11034C. To
achieve high rates of CO
2
fixation, plants therefore need
large amounts of this enzyme. In fact, rubisco makes up
almost 50% of soluble protein in chloroplasts and is prob-
ably one of the most abundant enzymes in the biosphere.
Central to the proposed mechanism for plant ru-
bisco is a carbamoylated Lys side chain with a bound
Mg
2H11001
ion. The Mg
2H11001
ion brings together and orients
the reactants at the active site (Fig. 20–6) and polar-
izes the CO
2
, opening it to nucleophilic attack by the
five-carbon enediolate reaction intermediate formed on
the enzyme (Fig. 20–7). The resulting six-carbon inter-
mediate breaks down to yield two molecules of 3-
phosphoglycerate.
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria754
Ribulose 1,5-
bisphosphate
(3)
C
CHOH
O
P
CHOH
CH
2
O
PCH
2
O
CHOH
PCH
2
O
CHO
Glyceraldehyde 3-phosphate
(6)
(3)
H11001 H
H11001
ATP
(1)
(6)
ADP
(3)
NADP
H11001
(6)
(6)ADP
(6)
P
i
(6)
(5)
COO
H11002
PCH
2
O
3-Phosphoglycerate
(6)
CHOH
Energy
production
via glycolysis;
starch or
sugar
synthesis
Stage 3:
Regeneration
of acceptor Stage 1:
Fixation
Stage 2:
Reduction
(3)
CO
2
ATP
NADPH
FIGURE 20–4 The three stages of CO
2
assimilation in photosynthetic organisms. Stoichiome-
tries of three key intermediates (numbers in parentheses) reveal the fate of carbon atoms
entering and leaving the cycle. As shown here, three CO
2
are fixed for the net synthesis of one
molecule of glyceraldehyde 3-phosphate. This cycle is the photosynthetic carbon reduction
cycle, or the Calvin cycle.
8885d_c20_751–786 2/18/04 1:56 PM Page 754 mac76 mac76:385_reb:
20.1 Photosynthetic Carbohydrate Synthesis 755
(b)
(a)
Top view
Side view
FIGURE 20–5 Structure of ribulose 1,5-bisphosphate
carboxylase (rubisco). (a) Top and side view of a ribbon
model of rubisco from spinach (PDB ID 8RUC). The
enzyme has eight large subunits (blue) and eight small
ones (gray), tightly packed into a structure of M
r
550,000. Rubisco is present at a concentration of about
250 mg/mL in the chloroplast stroma, corresponding to
an extraordinarily high concentration of active sites (~4
mM). Amino acid residues of the active site are shown in
yellow, Mg
2H11001
in green. (b) Ribbon model of rubisco
from the bacterium Rhodospirillum rubrum (PDB ID
9RUB). The subunits are in gray and blue. A Lys residue
at the active site that is carboxylated to a carbamate in
the active enzyme is shown in red. The substrate,
ribulose 1,5-bisphosphate, is yellow; Mg
2+
is green.
Asp
203
Glu
204
H
2
O (in CO
2
site)
Carbamoyl-Lys
201
Ribulose 1,5-bisphosphate
FIGURE 20–6 Central role of Mg
2H11545
in the catalytic
mechanism of rubisco. (Derived from PDB ID 1RXO)
Mg
2H11001
is coordinated in a roughly octahedral complex
with six oxygen atoms: one oxygen in the carbamate
on Lys
201
; two in the carboxyl groups of Glu
204
and
Asp
203
; two at C-2 and C-3 of the substrate, ribulose
1,5-bisphosphate; and one in the other substrate, CO
2
.
A water molecule occupies the Co2–binding site in this
crystal structure. (Residue numbers refer to the spinach
enzyme.)
8885d_c20_755 2/20/04 12:00 PM Page 755 mac76 mac76:385_reb:
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria756
Glu
Asp
Lys
201
N
H
O
–
OH
H
O
O
HC
OH
Enediolate
intermediate
C
C
C
C
+
Mg
2+
O
Glu
Rubisco
Asp
Lys
201
N
H
O
–
O
–
H
O
O
O
CH
2
P
CO
2
HC
OH
H
2
O
Ribulose 1,5-bisphosphate
3-Phosphoglycerate
Carbamoylated
Lys side chain
C
C
H
C
+
N
H
His
294
b-Keto acid
intermediate
Hydrated intermediate
2
3
4
:
:
N
Mg
2+
O
CH
2
O
H
N
H
His
294
:N
5
1
O
CH
2
P
O
CH
2
P
O
CH
2
P
P O
CH
2
P
O
CH
2
P
Glu
Asp
Lys
201
N
H
O
– –
O
O
O
OH
HC
OHC
C
C
C
+
Mg
2+
H
H
O
–
OO
C
HCOH
–
OO
C
CH
2
O P
Glu
Asp
Lys
201
N
H
O
–
–
O
O
–
O
OH
H
O
HC
OHC
C
C
C
+
Mg
2+
OH
H
O
O
CH
2
P
O
CH
2
P
Glu
Asp
Lys
201
N
H
O
OH
C
+
O
H
–
C
Mg
2+
Lys
175
N
+
H
H
H
–
OO
C
HCOH
CH
2
O P
3-Phosphoglycerate
MECHANISM FIGURE 20–7 First stage of CO
2
assimilation: rubisco’s
carboxylase activity. The CO
2
-fixation reaction is catalyzed by ribulose
1,5-bisphosphate carboxylase/oxygenase (rubisco). 1 Ribulose 1,5-
bisphosphate forms an enediolate at the active site. 2 CO
2
, polarized
by the proximity of the Mg
2H11001
ion, undergoes nucleophilic attack by
the enediolate, producing a branched six-carbon sugar. 3 Hydroxy-
lation at C-3 of this sugar, followed by aldol cleavage 4 , forms one
molecule of 3-phosphoglycerate, which leaves the enzyme active site.
5 The carbanion of the remaining three-carbon fragment is proto-
nated by the nearby side chain of Lys
175
, generating a second molecule
of 3-phosphoglycerate. The overall reaction therefore accomplishes
the combination of one CO
2
and one ribulose 1,5-bisphosphate to
form two molecules of 3-phosphoglycerate, one of which contains the
carbon atom from CO
2
(red). Rubisco Mechanism; Rubisco
Tutorial
8885d_c20_756 2/20/04 12:01 PM Page 756 mac76 mac76:385_reb:
As the catalyst for the first step of photosynthetic
CO
2
assimilation, rubisco is a prime target for regula-
tion. The enzyme is inactive until carbamoylated on
the H9255 amino group of Lys
201
(Fig. 20–8). Ribulose 1,5-
bisphosphate inhibits carbamoylation by binding tightly
to the active site and locking the enzyme in the “closed”
conformation, in which Lys
201
is inaccessible. Rubisco
activase overcomes the inhibition by promoting ATP-
dependent release of the ribulose 1,5-bisphosphate, ex-
posing the Lys amino group to nonenzymatic car-
bamoylation by CO
2
; this is followed by Mg
2H11001
binding,
which activates the rubisco. Rubisco activase in some
species is activated by light through a redox mechanism
(see Fig. 20–19).
Another regulatory mechanism involves the “noc-
turnal inhibitor” 2-carboxyarabinitol 1-phosphate, a nat-
urally occurring transition-state analog (see Box 6–3)
with a structure similar to that of the H9252-keto acid in-
termediate of the rubisco reaction (Fig. 20–7; see also
Fig. 20–20). This compound, synthesized in the dark in
some plants, is a potent inhibitor of carbamoylated ru-
bisco. It is either broken down when light returns or is
expelled by rubisco activase, activating the rubisco.
Stage 2: Conversion of 3-Phosphoglycerate to Glyceraldehyde
3-Phosphate The 3-phosphoglycerate formed in stage 1
is converted to glyceraldehyde 3-phosphate in two steps
that are essentially the reversal of the corresponding
steps in glycolysis, with one exception: the nucleotide
cofactor for the reduction of 1,3-bisphosphoglycerate is
NADPH rather than NADH (Fig. 20–9). The chloroplast
stroma contains all the glycolytic enzymes except phos-
phoglycerate mutase. The stromal and cytosolic enzymes
are isozymes; both sets of enzymes catalyze the same
reactions, but they are the products of different genes.
CH
2
HO
OH
OH
PO
3
2
H11002
CH
2
OH
O
CC
2-Carboxyarabinitol 1-phosphate
C
C
H
H
O
H11002
O
20.1 Photosynthetic Carbohydrate Synthesis 757
ADP
HCOH
HCOH
rubisco
activase
CO
2
Ribulose
1,5-bisphosphate
ATP-dependent
removal of ribulose
1,5 bisphosphate
uncovers e-amino
group of Lys
201
.
e-Amino group of Lys
201
is carbamoylated by CO
2
;
Mg
2+
binds to carbamoyl-
Lys, activating rubisco.
Rubisco with
unmodified Lys
201
and bound ribulose
1,5-bisphosphate
is inactive.
O
Mg
2+
Mg
2+
O
O
C
CLys
201
NH
3
+
Lys
201
NH
3
+
+
Lys
201
N
H
Rubisco
ATP
O
CH
2
P
O
CH
2
P
FIGURE 20–8 Role of rubisco activase in the carbamoylation of Lys
201
of rubisco. When the substrate ribulose 1,5-bisphosphate is bound to
the active site, Lys
201
is not accessible. Rubisco activase couples ATP
hydrolysis to expulsion of the bound sugar bisphosphate, exposing
Lys
201
; this Lys residue can now be carbamoylated with CO
2
in a re-
action that is apparently not enzyme-mediated. Mg
2H11001
is attracted to
and binds to the negatively charged carbamoyl-Lys, and the enzyme
is thus activated.
In the first step of stage 2, the stromal 3-phospho-
glycerate kinase catalyzes the transfer of a phospho-
ryl group from ATP to 3-phosphoglycerate, yielding
1,3-bisphosphoglycerate. Next, NADPH donates elec-
trons in a reduction catalyzed by the chloroplast-specific
isozyme of glyceraldehyde 3-phosphate dehydroge-
nase, producing glyceraldehyde 3-phosphate and P
i
.
Triose phosphate isomerase then interconverts glycer-
aldehyde 3-phosphate and dihydroxyacetone phosphate.
8885d_c20_751–786 2/18/04 1:56 PM Page 757 mac76 mac76:385_reb:
Most of the triose phosphate thus produced is used to
regenerate ribulose 1,5-bisphosphate; the rest is either
converted to starch in the chloroplast and stored for
later use or immediately exported to the cytosol and con-
verted to sucrose for transport to growing regions of the
plant. In developing leaves, a significant portion of the
triose phosphate may be degraded by glycolysis to pro-
vide energy.
Stage 3: Regeneration of Ribulose 1,5-Bisphosphate from Triose
Phosphates The first reaction in the assimilation of CO
2
into triose phosphates consumes ribulose 1,5-bisphos-
phate and, for continuous flow of CO
2
into carbohydrate,
ribulose 1,5-bisphosphate must be constantly regener-
ated. This is accomplished in a series of reactions (Fig.
20–10) that, together with stages 1 and 2, constitute the
cyclic pathway shown in Figure 20–4. The product of
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria758
Sucrose
ATP
Starch
Dihydroxyacetone
phosphate
Fructose 6-phosphate
glyceraldehyde 3-phosphate
dehydrogenase
NADP
+
P
i
NADPH + H
+
ADP
triose phosphate
isomerase
transaldolase
fructose
1,6-bisphosphatase
glycolysis
P
i
–triose
phosphate
antiporter
Stroma
Cytosol
CH
2
O
C
COO
H5008
P
C
C
CH
2
O P
C O
H
OH
OH
H
H
HO
CH
2
O
C
P
C
C
CH
2
OH
C O
H
OH
OH
H
H
HO
CH
2
O
CHOH
P
C
CH
2
O
CHOH
P
PO
O
CH
CH
2
O
CHOH
P
O
CH
2
OH
CH
2
O
C
P
O
3-phosphoglycerate
kinase
Fructose
6-phosphate
Glyceraldehyde 3-phosphate
Fructose
1,6-bisphosphate
Glyceraldehyde
3-phosphate
Dihydroxyacetone
phosphate
Fructose
1,6-bisphosphate
1,3-Bisphosphoglycerate
3-Phosphoglycerate
ATP
FIGURE 20–9 Second stage of CO
2
assimilation. 3-Phosphoglycerate
is converted to glyceraldehyde 3-phosphate (red arrows). Also shown
are the alternative fates of the fixed carbon of glyceraldehyde
3-phosphate (blue arrows). Most of the glyceraldehyde 3-phosphate is
recycled to ribulose 1,5-bisphosphate as shown in Figure 20–10. A
small fraction of the “extra” glyceraldehyde 3-phosphate may be used
immediately as a source of energy, but most is converted to sucrose
for transport or is stored in the chloroplast as starch. In the latter case,
glyceraldehyde 3-phosphate condenses with dihydroxyacetone phos-
phate in the stroma to form fructose 1,6-bisphosphate, a precursor of
starch. In other situations the glyceraldehyde 3-phosphate is converted
to dihydroxyacetone phosphate, which leaves the chloroplast via a
specific transporter (see Fig. 20–15) and, in the cytosol, can be
degraded glycolytically to provide energy or used to form fructose
6-phosphate and hence sucrose.
8885d_c20_758 2/20/04 12:01 PM Page 758 mac76 mac76:385_reb:
the first assimilation reaction (3-phosphoglycerate)
thus undergoes transformations that regenerate ribu-
lose 1,5-bisphosphate. The intermediates in this path-
way include three-, four-, five-, six-, and seven-carbon
sugars. In the following discussion, all step numbers re-
fer to Figure 20–10.
Steps 1 and 4 are catalyzed by the same enzyme,
transaldolase. It first catalyzes the reversible conden-
sation of glyceraldehyde 3-phosphate with dihydroxy-
acetone phosphate, yielding fructose 1,6-bisphosphate
(step 1 ); this is cleaved to fructose 6-phosphate and
P
i
by fructose 1,6-bisphosphatase (FBPase-1) in step
2 . The reaction is strongly exergonic and essentially
irreversible. Step 3 is catalyzed by transketolase,
which contains thiamine pyrophosphate (TPP) as its
prosthetic group (see Fig. 14–13a) and requires Mg
2H11001
.
20.1 Photosynthetic Carbohydrate Synthesis 759
Glyceraldehyde 3-phosphate Dihydroxyacetone phosphate
transaldolase
fructose 1,6-bisphosphatase
transketolase
ribulose
5-phosphate
epimerase
sedoheptulose
1,7-bisphosphatase
ADP
ADP
ribulose 5-phosphate
epimerase
P
i
Fructose 1,6-bisphosphate
Fructose 6-phosphate
Erythrose 4-phosphateDihydroxyacetone phosphate
Glyceraldehyde 3-phosphate
Xylulose 5-phosphate
Ribulose 5-phosphate
Ribulose 5-phosphate
ribose
5-phosphate
isomerase
ATP
ATP
ATP
ADP
ribulose
5-phosphate
kinase
transketolase
Glyceraldehyde 3-phosphate
Ribose 5-phosphate
Ribulose 5-phosphate Ribulose 1,5-bisphosphate
Ribulose 1,5-
bisphosphate
Sedoheptulose 1,7-bisphosphate
Sedoheptulose 7-phosphate
Xylulose 5-phosphate
P
i
transaldolase
1
2
3
4
5
9
8
6
7
9
8
FIGURE 20–10 Third stage of CO
2
assimilation. This schematic
diagram shows the interconversions of triose phosphates and pentose
phosphates. Black dots represent the number of carbons in each
compound. The starting materials are glyceraldehyde 3-phosphate and
dihydroxyacetone phosphate. Reactions catalyzed by transaldolase
( 1 and 4 ) and transketolase ( 3 and 6 ) produce pentose phos-
phates that are converted to ribulose 1,5-bisphosphate—ribose
5-phosphate by ribose 5-phosphate isomerase ( 7 ) and xylulose
5-phosphate by ribulose 5-phosphate epimerase ( 8 ). In step 9 ,
ribulose 5-phosphate is phosphorylated, regenerating ribulose 1,5-
bisphosphate. The steps with blue arrows are exergonic and make
the whole process irreversible: steps 2 fructose 1,6-bisphosphatase,
5 sedoheptulose bisphosphatase, and 9 ribulose 5-phosphate
kinase.
8885d_c20_759 2/20/04 12:01 PM Page 759 mac76 mac76:385_reb:
Transketolase catalyzes the reversible transfer of a
2-carbon ketol group (CH
2
OHOCOO) from a ketose
phosphate donor, fructose 6-phosphate, to an aldose
phosphate acceptor, glyceraldehyde 3-phosphate (Fig.
20–11a, b), forming the pentose xylulose 5-phosphate
and the tetrose erythrose 4-phosphate. In step 4 ,
transaldolase acts again, combining erythrose 4-phos-
phate with dihydroxyacetone phosphate to form the
seven-carbon sedoheptulose 1,7-bisphosphate. An
enzyme unique to plastids, sedoheptulose 1,7-bisphos-
phatase, converts the bisphosphate to sedoheptulose
7-phosphate (step 5 ); this is the second irreversible
reaction in the pathway. Transketolase now acts again,
converting sedoheptulose 7-phosphate and glyceralde-
hyde 3-phosphate to two pentose phosphates in step 6
(Fig. 20–11c). Figure 20–12 shows how a two-carbon
fragment is temporarily carried on the transketolase
cofactor TPP and condensed with the three carbons of
glyceraldehyde 3-phosphate in step 6 .
The pentose phosphates formed in the transketo-
lase reactions—ribose 5-phosphate and xylulose 5-phos-
phate—are converted to ribulose 5-phosphate (steps
7 and 8 ), which in the final step ( 9 ) of the cycle is
phosphorylated to ribulose 1,5-bisphosphate by ribulose
5-phosphate kinase (Fig. 20–13). This is the third very
exergonic reaction of the pathway, as the phosphate an-
hydride bond in ATP is swapped for a phosphate ester
in ribulose 1,5-bisphosphate.
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria760
CH
2
O
CO
CHOH
P
H11001
Ketose
donor
Aldose
acceptor
TPP
transketolase
Fructose
6-phosphate
Glyceraldehyde
3-phosphate
Xylulose
5-phosphate
Erythrose
4-phosphate
Sedoheptulose
7-phosphate
Glyceraldehyde
3-phosphate
(a)
(b)
(c)
CH
2
OH
R
2
C
O
R
1
CHOH
H11001
R
2
R
1
CH
2
OH
C O
O
C
C
H11001
HO HC
CO
CH
2
OH
C H11001
C
O
C
CH
2
OP
OHH
H OH
CH
2
O
HO
OH
P
HC
CO
H11001
CH
2
OH
C
O
C
C
C
CH
OHH
CH
2
OP
HO
OH
HC
CO
CH
2
OH
C
H
CH
2
OP
H11001
H
C
O
C
CH
2
OP
OHH
H
CH OH
OH
OH
H
OHH
HO
OH
Xylulose
5-phosphate
HC
CO
CH
2
OH
CH
CH
2
OP
H
H
H
H
Ribose
5-phosphate
C
O
C
C OHH
OHH
CH
2
OP
OHCH
H
FIGURE 20–11 Transketolase-catalyzed reactions of the Calvin cycle. (a) General reaction
catalyzed by transketolase: the transfer of a two-carbon group, carried temporarily on enzyme-
bound TPP, from a ketose donor to an aldose acceptor. (b) Conversion of a hexose and a triose
to a four-carbon and a five-carbon sugar (step 3 of Fig. 20–10). (c) Conversion of seven-
carbon and three-carbon sugars to two pentoses (step 6 of Fig. 20–10).
8885d_c20_751–786 2/18/04 1:56 PM Page 760 mac76 mac76:385_reb:
20.1 Photosynthetic Carbohydrate Synthesis 761
Ribose 5-phosphate
Xylulose
5-phosphate
Glyceraldehyde
3-phosphate
(aldose acceptor)
Sedoheptulose
7-phosphate
(ketose donor)
Thiamine pyrophosphate
(cofactor of transketolase)
Carbanion,
stabilized by
resonance
OH
CH
2
OH
C
C
C
C
C
CH
2
O P
OH
OH
O
HHO
H
H
H
CH
2
O P
P
OHC
C
C
OH
OH
H
H
H
C
HO
CH
2
O P
OHCH
C
HO
OH
CH
2
OH
C
C
C
CH
2
O P
O
HHO
H
OH
H11002
C CH
2
OH
CH
3
C
RH11032
SN
H11001
R
OH
C CH
2
OH
CH
3
C
RH11032
SN
H11001
R
H
OH
C
OH
H
C
OH
H
C
OH
H
C
O
CH
2
OCH
2
C
P
H
P
RH11032R
CH
2
CH
3
CH
3
CH
2
S
N
N
NH
2
N
H11001
FIGURE 20–12 TPP as a cofactor for transketolase. Transketolase
transfers a two-carbon group from sedoheptulose 7-phosphate to
glyceraldehyde 3-phosphate, producing two pentose phosphates
(step 6 in Fig. 20–10). Thiamine pyrophosphate serves as a
temporary carrier of the two-carbon unit and as an electron sink
(see Fig. 14–13) to facilitate the reactions.
O
CHOH
CH
2
O
C
OC
HC
H
H
HO
Xylulose 5-phosphate
ribose
5-phosphate
isomerase
ribose
5-phosphate
epimerase
P
CHOH
CH
2
OH
OHC
OHCH
Ribose 5-phosphate
PCH
2
O
CHOH
CH
2
O
OC
OHCH
Ribulose
5-phosphate
P
CH
2
OH
CHOH
CH
2
O
OC
OHCH
Ribulose
1,5-bisphosphate
P
ATP ADP
CH
2
O P
ribulose
5-phosphate
kinase
FIGURE 20–13 Regeneration of ribulose 1,5-bisphosphate. The starting material
for the Calvin cycle, ribulose 1,5-bisphosphate, is regenerated from two pentose
phosphates produced in the cycle. This pathway involves the action of an
isomerase and an epimerase, then phosphorylation by a kinase, with ATP as
phosphate group donor (steps 7 , 8 , and 9 of Fig. 20–10).
8885d_c20_751–786 2/18/04 1:56 PM Page 761 mac76 mac76:385_reb:
Synthesis of Each Triose Phosphate from CO
2
Requires Six NADPH and Nine ATP
The net result of three turns of the Calvin cycle is the
conversion of three molecules of CO
2
and one molecule
of phosphate to a molecule of triose phosphate. The stoi-
chiometry of the overall path from CO
2
to triose phos-
phate, with regeneration of ribulose 1,5-bisphosphate,
is shown in Figure 20–14. Three molecules of ribulose
1,5-bisphosphate (a total of 15 carbons) condense with
three CO
2
(3 carbons) to form six molecules of 3-phos-
phoglycerate (18 carbons). These six molecules of 3-
phosphoglycerate are reduced to six molecules of glyc-
eraldehyde 3-phosphate (which is in equilibrium with
dihydroxyacetone phosphate), with the expenditure of
six ATP (in the synthesis of 1,3-bisphosphoglycerate)
and six NADPH (in the reduction of 1,3-bisphospho-
glycerate to glyceraldehyde 3-phosphate). The isozyme
of glyceraldehyde 3-phosphate dehydrogenase present
in chloroplasts can use NADP as its electron carrier and
normally functions in the direction of 1,3-bisphospho-
glycerate reduction. The cytosolic isozyme uses NAD,
as does the glycolytic enzyme of animals and other eu-
karyotes, and in the dark this isozyme acts in glycolysis
to oxidize glyceraldehyde 3-phosphate. Both glycer-
aldehyde 3-phosphate dehydrogenase isozymes, like all
enzymes, catalyze the reaction in both directions.
One molecule of glyceraldehyde 3-phosphate is the
net product of the carbon assimilation pathway. The
other five triose phosphate molecules (15 carbons) are
rearranged in steps 1 to 9 of Figure 20–10 to form
three molecules of ribulose 1,5-bisphosphate (15 car-
bons). The last step in this conversion requires one ATP
per ribulose 1,5-bisphosphate, or a total of three ATP.
Thus, in summary, for every molecule of triose phos-
phate produced by photosynthetic CO
2
assimilation, six
NADPH and nine ATP are required.
NADPH and ATP are produced in the light-
dependent reactions of photosynthesis in about the
same ratio (2:3) as they are consumed in the Calvin cy-
cle. Nine ATP molecules are converted to ADP and phos-
phate in the generation of a molecule of triose phos-
phate; eight of the phosphates are released as P
i
and
combined with eight ADP to regenerate ATP. The ninth
phosphate is incorporated into the triose phosphate it-
self. To convert the ninth ADP to ATP, a molecule of P
i
must be imported from the cytosol, as we shall see.
In the dark, the production of ATP and NADPH by
photophosphorylation, and the incorporation of CO
2
into triose phosphate (by the so-called dark reactions),
cease. The “dark reactions” of photosynthesis were so
named to distinguish them from the primary light-
driven reactions of electron transfer to NADP
H11001
and syn-
thesis of ATP, described in Chapter 19. They do not, in
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria762
Glyceraldehyde 3-phosphate1
6 ATP
6ADP
6 NADPH + 6H
+
6P
i
6NADP
+
3ADP
3 ATP
2P
i
1,3-Bisphosphoglycerate6
3-Phosphoglycerate6
Ribulose 1,5-bisphosphate3
Ribulose 5-phosphate3
Glyceraldehyde 3-phosphate
5
CO
2
+ H
2
O3
Dihydroxyacetone phosphate
Glyceraldehyde 3-phosphate
Dihydroxyacetone phosphate
6
FIGURE 20–14 Stoichiometry of CO
2
assimilation in the Calvin cycle. For every three CO
2
molecules fixed, one molecule of triose phosphate (glyceraldehyde 3-phosphate) is produced
and nine ATP and six NADPH are consumed.
8885d_c20_751–786 2/18/04 1:56 PM Page 762 mac76 mac76:385_reb:
fact, occur at significant rates in the dark and are thus
more appropriately called the carbon-assimilation re-
actions. Later in this section we describe the regula-
tory mechanisms that turn on carbon assimilation in the
light and turn it off in the dark.
The chloroplast stroma contains all the enzymes
necessary to convert the triose phosphates produced by
CO
2
assimilation (glyceraldehyde 3-phosphate and di-
hydroxyacetone phosphate) to starch, which is tem-
porarily stored in the chloroplast as insoluble granules.
Aldolase condenses the trioses to fructose 1,6-bisphos-
phate; fructose 1,6-bisphosphatase produces fructose 6-
phosphate; phosphohexose isomerase yields glucose 6-
phosphate; and phosphoglucomutase produces glucose
1-phosphate, the starting material for starch synthesis
(see Section 20.3).
All the reactions of the Calvin cycle except those
catalyzed by rubisco, sedoheptulose 1,7-bisphospha-
tase, and ribulose 5-phosphate kinase also take place in
animal tissues. Lacking these three enzymes, animals
cannot carry out net conversion of CO
2
to glucose.
A Transport System Exports Triose Phosphates
from the Chloroplast and Imports Phosphate
The inner chloroplast membrane is impermeable to most
phosphorylated compounds, including fructose 6-phos-
phate, glucose 6-phosphate, and fructose 1,6-bisphos-
phate. It does, however, have a specific antiporter that
catalyzes the one-for-one exchange of P
i
with a triose
phosphate, either dihydroxyacetone phosphate or 3-
phosphoglycerate (Fig. 20–15; see also Fig. 20–9). This
antiporter simultaneously moves P
i
into the chloroplast,
where it is used in photophosphorylation, and moves
triose phosphate into the cytosol, where it can be used
to synthesize sucrose, the form in which the fixed car-
bon is transported to distant plant tissues.
Sucrose synthesis in the cytosol and starch synthe-
sis in the chloroplast are the major pathways by which
the excess triose phosphate from photosynthesis is “har-
vested.” Sucrose synthesis (described below) releases
four P
i
molecules from the four triose phosphates re-
quired to make sucrose. For every molecule of triose
phosphate removed from the chloroplast, one P
i
is trans-
ported into the chloroplast, providing the ninth P
i
men-
tioned above, to be used in regenerating ATP. If this ex-
change were blocked, triose phosphate synthesis would
quickly deplete the available P
i
in the chloroplast, slow-
ing ATP synthesis and suppressing assimilation of CO
2
into starch.
The P
i
–triose phosphate antiport system serves one
additional function. ATP and reducing power are needed
in the cytosol for a variety of synthetic and energy-
requiring reactions. These requirements are met to an
as yet undetermined degree by mitochondria, but a sec-
ond potential source of energy is the ATP and NADPH
generated in the chloroplast stroma during the light
reactions. However, neither ATP nor NADPH can cross
the chloroplast membrane. The P
i
–triose phosphate
antiport system has the indirect effect of moving ATP
20.1 Photosynthetic Carbohydrate Synthesis 763
Stroma
Cytosol
Chloroplast
inner membrane
P
i
–triose
phosphate antiporter
Dihydroxy-
acetone
phosphate
Sucrose
P
i
P
i
Dihydroxy-
acetone
phosphate
photosynthesis
Calvin
cycle
9ADP + 8P
i
+ P
i
9ATP
9ATP 9ADP + 9P
i
Light
FIGURE 20–15 The P
i
–triose phosphate antiport system of the inner
chloroplast membrane. This transporter facilitates the exchange of
cytosolic P
i
for stromal dihydroxyacetone phosphate. The products of
photosynthetic carbon assimilation are thus moved into the cytosol
where they serve as a starting point for sucrose biosynthesis, and P
i
required for photophosphorylation is moved into the stroma. This same
antiporter can transport 3-phosphoglycerate and acts in the shuttle for
exporting ATP and reducing equivalents (see Fig. 20–16).
8885d_c20_763 2/20/04 12:02 PM Page 763 mac76 mac76:385_reb:
equivalents and reducing equivalents from the chloro-
plast to the cytosol (Fig. 20–16). Dihydroxyacetone
phosphate formed in the stroma is transported to the
cytosol, where it is converted by glycolytic enzymes
to 3-phosphoglycerate, generating ATP and NADH. 3-
Phosphoglycerate reenters the chloroplast, completing
the cycle.
Four Enzymes of the Calvin Cycle Are Indirectly
Activated by Light
The reductive assimilation of CO
2
requires a lot of ATP
and NADPH, and their stromal concentrations increase
when chloroplasts are illuminated (Fig. 20–17). The
light-induced transport of protons across the thylakoid
membrane (Chapter 19) also increases the stromal pH
from about 7 to about 8, and it is accompanied by a flow
of Mg
2H11001
from the thylakoid compartment into the
stroma, raising the [Mg
2H11001
] from 1 to 3 mM to 3 to 6 mM.
Several stromal enzymes have evolved to take advan-
tage of these light-induced conditions, which signal the
availability of ATP and NADPH: the enzymes are more
active in an alkaline environment and at high [Mg
2H11001
].
For example, activation of rubisco by formation of the
carbamoyllysine is faster at alkaline pH, and high stro-
mal [Mg
2H11001
] favors formation of the enzyme’s active Mg
2H11001
complex. Fructose 1,6-bisphosphatase requires Mg
2H11001
and is very dependent on pH (Fig. 20–18); its activity
increases more than 100-fold when pH and [Mg
2H11001
] rise
during chloroplast illumination.
Four Calvin cycle enzymes are subject to a special
type of regulation by light. Ribulose 5-phosphate kinase,
fructose 1,6-bisphosphatase, sedoheptulose 1,7-bisphos-
phatase, and glyceraldehyde 3-phosphate dehydroge-
nase are activated by light-driven reduction of disulfide
bonds between two Cys residues critical to their cat-
alytic activities. When these Cys residues are disulfide-
bonded (oxidized), the enzymes are inactive; this is the
normal situation in the dark. With illumination, electrons
flow from photosystem I to ferredoxin (see Fig. 19–49),
which passes electrons to a small, soluble, disulfide-
containing protein called thioredoxin (Fig. 20–19), in
a reaction catalyzed by ferredoxin-thioredoxin re-
ductase. Reduced thioredoxin donates electrons for the
reduction of the disulfide bonds of the light-activated
enzymes, and these reductive cleavage reactions are
accompanied by conformational changes that increase
enzyme activities. At nightfall, the Cys residues in the
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria764
NADH
+ H
+
Stroma
Chloroplast
inner membrane
Cytosol
P
i
–triose
phosphate antiporter
P
i
P
i
3-Phosphoglycerate
1,3-Bisphosphoglycerate
Dihydroxyacetone
phosphate
NAD
+
ADP
P
i
P
i
phospho-
glycerate
kinase
glyceraldehyde
3-phosphate
dehydrogenase
1,3-Bisphosphoglycerate
Glyceraldehyde
3-phosphate
ADP
Dihydroxyacetone
phosphate
3-Phosphoglycerate
P
i
–triose
phosphate antiporter
ATPATP
Glyceraldehyde
3-phosphate
NADPH
+ H
+
NADP
+
triose
phosphate
isomerase
FIGURE 20–16 Role of the P
i
–triose phosphate
antiporter in the transport of ATP and reducing
equivalents. Dihydroxyacetone phosphate leaves
the chloroplast and is converted to glyceraldehyde
3-phosphate in the cytosol. The cytosolic glycer-
aldehyde 3-phosphate dehydrogenase and
phosphoglycerate kinase reactions then produce
NADH, ATP, and 3-phosphoglycerate. The latter
reenters the chloroplast and is reduced to dihy-
droxyacetone phosphate, completing a cycle that
effectively moves ATP and reducing equivalents
(NADPH/NADH) from chloroplast to cytosol.
8885d_c20_751–786 2/18/04 1:56 PM Page 764 mac76 mac76:385_reb:
four enzymes are reoxidized to their disulfide forms, the
enzymes are inactivated, and ATP is not expended in
CO
2
assimilation. Instead, starch synthesized and stored
during the daytime is degraded to fuel glycolysis at night.
Glucose 6-phosphate dehydrogenase, the first en-
zyme in the oxidative pentose phosphate pathway, is
also regulated by this light-driven reduction mechanism,
but in the opposite sense. During the day, when photo-
synthesis produces plenty of NADPH, this enzyme is not
needed for NADPH production. Reduction of a critical
disulfide bond by electrons from ferredoxin inactivates
the enzyme.
20.1 Photosynthetic Carbohydrate Synthesis 765
Light
CO
2
-assimilation
cycle
Glyceraldehyde 3-phosphate
CO
2
P
i
H
+
Mg
2+Photosynthetic
electron transfer
Stroma
Thylakoid
+ H
+
ATP NADPH NADP
+
ADP + P
i
FIGURE 20–17 Source of ATP and NADPH. ATP and NADPH pro-
duced by the light reactions are essential substrates for the reduction
of CO
2
. The photosynthetic reactions that produce ATP and NADPH
are accompanied by movement of protons (red) from the stroma into
the thylakoid, creating alkaline conditions in the stroma. Magnesium
ions pass from the thylakoid into the stroma, increasing the stromal
[Mg
2H11001
].
FIGURE 20–18 Activation of chloroplast fructose 1,6-bisphosphatase.
Reduced fructose 1,6-bisphosphatase (FBPase-1) is activated by light
and by the combination of high pH and high [Mg
2H11001
] in the stroma,
both of which are produced by illumination.
FBPase-1 activity (units/mg)
0
150
[MgCl
2
] (mM)
100
50
2015105
pH 8.0
pH 7.5
pH 7.0
0
FIGURE 20–19 Light activation of several enzymes of the Calvin
cycle. The light activation is mediated by thioredoxin, a small,
disulfide-containing protein. In the light, thioredoxin is reduced by
electrons moving from photosystem I through ferredoxin (Fd) (blue
arrows), then thioredoxin reduces critical disulfide bonds in each
of the enzymes sedoheptulose 1,7-bisphosphatase, fructose 1,6-
bisphosphatase, ribulose 5-phosphate kinase, and glyceraldehye
3-phosphate dehydrogenase, activating these enzymes. In the dark,
the OSH groups undergo reoxidation to disulfides, inactivating the
enzymes.
Fd
ox
Fd
red
ferredoxin-
thioredoxin
reductase
Thioredoxin
HS SH
Enzyme
(active)
O
2
(in dark)
Photosystem I
HS SH
Enzyme
(inactive)
SS
Thioredoxin
SS
Light
8885d_c20_765 2/20/04 12:02 PM Page 765 mac76 mac76:385_reb:
SUMMARY 20.1 Photosynthetic Carbohydrate
Synthesis
■ Photosynthesis in vascular plants takes place in
chloroplasts. In the CO
2
-assimilating reactions
(the Calvin cycle), ATP and NADPH are used
to reduce CO
2
to triose phosphates. These
reactions occur in three stages: the fixation
reaction itself, catalyzed by rubisco; reduction
of the resulting 3-phosphoglycerate to
glyceraldehyde 3-phosphate; and regeneration
of ribulose 1,5-bisphosphate from triose
phosphates.
■ Rubisco condenses CO
2
with ribulose
1,5-bisphosphate, forming an unstable hexose
bisphosphate that splits into two molecules of
3-phosphoglycerate. Rubisco is activated by
covalent modification (carbamoylation of
Lys
201
) catalyzed by rubisco activase and is
inhibited by a natural transition-state analog,
whose concentration rises in the dark and falls
during daylight.
■ Stromal isozymes of the glycolytic enzymes
catalyze reduction of 3-phosphoglycerate to
glyceraldehyde 3-phosphate; each molecule
reduced requires one ATP and one NADPH.
■ Stromal enzymes, including transketolase and
transaldolase, rearrange the carbon skeletons
of triose phosphates, generating intermediates
of three, four, five, six, and seven carbons
and eventually yielding pentose phosphates.
The pentose phosphates are converted to
ribulose 5-phosphate, then phosphorylated to
ribulose 1,5-bisphosphate to complete the
Calvin cycle.
■ The cost of fixing three CO
2
into one triose
phosphate is nine ATP and six NADPH, which
are provided by the light-dependent reactions
of photosynthesis.
■ An antiporter in the inner chloroplast
membrane exchanges P
i
in the cytosol for
3-phosphoglycerate or dihydroxyacetone
phosphate produced by CO
2
assimilation in the
stroma. Oxidation of dihydroxyacetone phos-
phate in the cytosol generates ATP and NADH,
thus moving ATP and reducing equivalents
from the chloroplast to the cytosol.
■ Four enzymes of the Calvin cycle are activated
indirectly by light and are inactive in the dark,
so that hexose synthesis does not compete
with glycolysis—which is required to provide
energy in the dark.
20.2 Photorespiration and the C
4
and CAM Pathways
As we have seen, photosynthetic cells produce O
2
(by
the splitting of H
2
O) during the light-driven reactions
(Chapter 19) and use CO
2
during the light-independent
processes (described above), so the net gaseous change
during photosynthesis is the uptake of CO
2
and release
of O
2
:
CO
2
H11001 H
2
O 88n O
2
H11001 (CH
2
O)
In the dark, plants also carry out mitochondrial res-
piration, the oxidation of substrates to CO
2
and the
conversion of O
2
to H
2
O. And there is another process
in plants that, like mitochondrial respiration, consumes
O
2
and produces CO
2
and, like photosynthesis, is driven
by light. This process, photorespiration, is a costly
side reaction of photosynthesis, a result of the lack of
specificity of the enzyme rubisco. In this section we de-
scribe this side reaction and the strategies plants use to
minimize its metabolic consequences.
Photorespiration Results from Rubisco’s
Oxygenase Activity
Rubisco is not absolutely specific for CO
2
as a substrate.
Molecular oxygen (O
2
) competes with CO
2
at the
active site, and about once in every three or four
turnovers, rubisco catalyzes the condensation of O
2
with ribulose 1,5-bisphosphate to form 3-phosphoglyc-
erate and 2-phosphoglycolate (Fig. 20–20), a meta-
bolically useless product. This is the oxygenase activ-
ity referred to in the full name of the enzyme: ribulose
1,5-bisphosphate carboxylase/oxygenase. The reaction
with O
2
results in no fixation of carbon and appears to
be a net liability to the cell; salvaging the carbons from
2-phosphoglycolate (by the pathway outlined below)
consumes significant amounts of cellular energy and
releases some previously fixed CO
2
.
Given that the reaction with oxygen is deleterious
to the organism, why did the evolution of rubisco pro-
duce an active site unable to discriminate well between
CO
2
and O
2
? Perhaps much of this evolution occurred
before the time, about 2.5 billion years ago, when pro-
duction of O
2
by photosynthetic organisms started to
raise the oxygen content of the atmosphere. Before that
time, there was no selective pressure for rubisco to dis-
criminate between CO
2
and O
2
. The K
m
for CO
2
is about
9 H9262M, and that for O
2
is about 350 H9262M. The modern at-
mosphere contains about 20% O
2
and only 0.04% CO
2
,
so an aqueous solution in equilibrium with air at room
temperature contains about 250 H9262M O
2
and 11 H9262M CO
2
—
concentrations that allow significant O
2
“fixation” by ru-
bisco and thus a significant waste of energy. The tem-
perature dependence of the solubilities of O
2
and CO
2
is
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria766
8885d_c20_751–786 2/18/04 1:56 PM Page 766 mac76 mac76:385_reb:
such that at higher temperatures, the ratio of O
2
to CO
2
in solution increases. In addition, the affinity of rubisco
for CO
2
decreases with increasing temperature, exacer-
bating its tendency to catalyze the wasteful oxygenase
reaction. And as CO
2
is consumed in the assimilation re-
actions, the ratio of O
2
to CO
2
in the air spaces of a leaf
increases, further favoring the oxygenase reaction.
The Salvage of Phosphoglycolate Is Costly
The glycolate pathway converts two molecules of 2-
phosphoglycolate to a molecule of serine (three carbons)
and a molecule of CO
2
(Fig. 20–21). In the chloroplast, a
phosphatase converts 2-phosphoglycolate to glycolate,
which is exported to the peroxisome. There, glycolate is
20.2 Photorespiration and the C
4
and CAM Pathways 767
2-Phospho-
glycolate
CO
2
released in
photorespiration
Mitochondrion
NAD
+
Serine
Hydroxypyruvate
H
2
O
2
O
2
Glyoxylate
Glycine
Glycolate
glycolic
acid
oxidase
Peroxisome
Serine 2 Glycine
-hydroxy
acid
reductase
GlycolateGlycerate
Calvin
cycle
ADP
P
i
Chloroplast
stroma
NADH
+ H
+
NAD
+
NH
3
glycine
decarboxylase
[ NH
2
]
transamination
+ H
+
CH
2
O P
COO
H5008
CH
2
OH
COO
H5008
CH
2
OH
COO
H5008
CHO
COO
H5008
COO
H5008
CH
2
NH
3
H11001
COO
H5008
CH
2
NH
3
H11001
OH
CH
NH
3
H11001
COO
H5008
CH
2
OOH
C COO
H5008
CH
2
Glycerate
OHOH
CH COO
H5008
CH
2
OHOH
CH COO
H5008
CH
2
OH
CH
NH
3
H11001
COO
H5008
CH
2
NADH
H9251
O
2
Ribulose 1,5-
bisphosphate
3-Phospho-
glycerate
ATP
CH
2
O
C
OH
HOHC
CO
CH
2
O
CH
2
O
C
O
HOH
C
CHOH
CH
2
O
Ribulose
1,5-bisphosphate
Enediol form
Enzyme-bound
intermediate
CH
2
O
C
OH
H
OH
C
COH
CH
2
O
O
2
O
O
H
OH
H5008
H
2
O
CH
2
O
C
H11001
H5008
O
C
CH
2
O
OHH
O
H5008
O
C
O
2-Phosphoglycolate 3-Phosphoglycerate
P
P
P
P
P
P
P
P
FIGURE 20–20 Oxygenase activity of rubisco. Rubisco can incorpo-
rate O
2
rather than CO
2
into ribulose 1,5-bisphosphate. The unstable
intermediate thus formed splits into 2-phosphoglycolate (recycled as
described in Fig. 20–21) and 3-phosphoglycerate, which can reenter
the Calvin cycle.
FIGURE 20–21 Glycolate pathway. This pathway, which salvages 2-
phosphoglycolate (shaded pink) by its conversion to serine and even-
tually 3-phosphoglycerate, involves three cellular compartments. Gly-
colate formed by dephosphorylation of 2-phosphoglycolate in
chloroplasts is oxidized to glyoxylate in peroxisomes and then
transaminated to glycine. In mitochondria, two glycine molecules con-
dense to form serine and the CO
2
released during photorespiration
(shaded green). This reaction is catalyzed by glycine decarboxylase,
an enzyme present at very high levels in the mitochondria of C
3
plants
(see text). The serine is converted to hydroxypyruvate and then to glyc-
erate in peroxisomes; glycerate reenters the chloroplasts to be phos-
phorylated, rejoining the Calvin cycle. Oxygen (shaded blue) is con-
sumed at two steps during photorespiration.
8885d_c20_751–786 2/18/04 1:56 PM Page 767 mac76 mac76:385_reb:
oxidized by molecular oxygen, and the resulting aldehyde
(glyoxylate) undergoes transamination to glycine. The hy-
drogen peroxide formed as a side product of glycolate ox-
idation is rendered harmless by peroxidases in the per-
oxisome. Glycine passes from the peroxisome to the
mitochondrial matrix, where it undergoes oxidative de-
carboxylation by the glycine decarboxylase complex, an
enzyme similar in structure and mechanism to two mito-
chondrial complexes we have already encountered: the
pyruvate dehydrogenase complex and the H9251-ketoglutarate
dehydrogenase complex (Chapter 16). The glycine de-
carboxylase complex oxidizes glycine to CO
2
and
NH
3
, with the concomitant reduction of NAD
H11001
to NADH
and transfer of the remaining carbon from glycine to the
cofactor tetrahydrofolate (Fig. 20–22). The one-carbon
unit carried on tetrahydrofolate is then transferred to a
second glycine by serine hydroxymethyltransferase,
producing serine. The net reaction catalyzed by the
glycine decarboxylase complex and serine hydrox-
ymethyltransferase is
2 Glycine H11001 NAD
H11001
H11001 H
2
O 88n
serine H11001 CO
2
H11001 NH
3
H11001 NADH H11001 H
H11001
The serine is converted to hydroxypyruvate, to glycer-
ate, and finally to 3-phosphoglycerate, which is used to
regenerate ribulose 1,5-bisphosphate, completing the
long, expensive cycle (Fig. 20–21).
In bright sunlight, the flux through the glycolate sal-
vage pathway can be very high, producing about five
times more CO
2
than is typically produced by all the ox-
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria768
FAD
FAD
FAD
–
OOC CH
2
NH
+
HC
PLP
H
C
S
S
H
Glycine
COO
–
H
3
N
+
CO
2
H
2
N
H
2
O
CH
2
N
5
,N
10
-methylene H
4
F
H
2
O
NH
3
FAD
H
P
2
1
5
3
4
L
T
PLP
S
S
H
P
L
O
HC
T
PLP
S
S
H
P
L
FADH
2
O
HC
T
PLP
H
P
L
O
HC
T
PLP
S
HS
HS
HS
H
P
L
O
HC
T
NADH + H
+
NAD
+
H
4
F
FIGURE 20–22 The glycine decarboxylase system. Glycine decar-
boxylase in plant mitochondria is a complex of four types of subunits,
with the stoichiometry P
4
H
27
T
9
L
2
. Protein H has a covalently attached
lipoic acid residue that can undergo reversible oxidation. Step 1 is
formation of a Schiff base between pyridoxal phosphate (PLP) and
glycine, catalyzed by protein P (named for its bound PLP). In step
2 , protein P catalyzes oxidative decarboxylation of glycine, releas-
ing CO
2
; the remaining methylamine group is attached to one of the
OSH groups of reduced lipoic acid. 3 Protein T (which uses tetrahy-
drofolate (H
4
F) as cofactor) now releases NH
3
from the methylamine
moiety and transfers the remaining one-carbon fragment to tetrahy-
drofolate, producing N
5
,N
10
-methylene tetrahydrofolate. 4 Protein L
oxidizes the two OSH groups of lipoic acid to a disulfide, passing
electrons through FAD to NAD
H11001
5 , thus completing the cycle. The
N
5
,N
10
-methylene tetrahydrofolate formed in this process is used by
serine hydroxymethyltransferase to convert a molecule of glycine to
serine, regenerating the tetrahydrofolate that is essential for the reac-
tion catalyzed by protein T. The L subunit of glycine decarboxylase is
identical to the dihydrolipoyl dehydrogenase (E
3
) of pyruvate dehy-
drogenase and H9251-ketoglutarate dehydrogenase (see Fig. 16–6).
8885d_c20_751–786 2/18/04 1:56 PM Page 768 mac76 mac76:385_reb:
idations of the citric acid cycle. To generate this large
flux, mitochondria contain prodigious amounts of the
glycine decarboxylase complex: the four proteins of the
complex make up half of all the protein in the mito-
chondrial matrix in the leaves of pea and spinach plants!
In nonphotosynthetic parts of a plant, such as potato tu-
bers, mitochondria have very low concentrations of the
glycine decarboxylase complex.
The combined activity of the rubisco oxygenase
and the glycolate salvage pathway consumes O
2
and
produces CO
2
—hence the name photorespiration.
This pathway is perhaps better called the oxidative
photosynthetic carbon cycle or C
2
cycle, names
that do not invite comparison with respiration in mi-
tochondria. Unlike mitochondrial respiration, “pho-
torespiration” does not conserve energy and may
actually inhibit net biomass formation as much as 50%.
This inefficiency has led to evolutionary adaptations
in the carbon-assimilation processes, particularly in
plants that have evolved in warm climates.
In C
4
Plants, CO
2
Fixation and Rubisco Activity
Are Spatially Separated
In many plants that grow in the tropics (and in temper-
ate-zone crop plants native to the tropics, such as maize,
sugarcane, and sorghum) a mechanism has evolved to
circumvent the problem of wasteful photorespiration.
The step in which CO
2
is fixed into a three-carbon prod-
uct, 3-phosphoglycerate, is preceded by several steps,
one of which is temporary fixation of CO
2
into a four-
carbon compound. Plants that use this process are re-
ferred to as C
4
plants, and the assimilation process as
C
4
metabolism or the C
4
pathway. Plants that use the
carbon-assimilation method we have described thus far,
in which the first step is reaction of CO
2
with ribulose
1,5-bisphosphate to form 3-phosphoglycerate, are called
C
3
plants.
The C
4
plants, which typically grow at high light in-
tensity and high temperatures, have several important
characteristics: high photosynthetic rates, high growth
rates, low photorespiration rates, low rates of water loss,
and a specialized leaf structure. Photosynthesis in the
leaves of C
4
plants involves two cell types: mesophyll
and bundle-sheath cells (Fig. 20–23a). There are three
variants of C
4
metabolism, worked out in the 1960s by
Marshall Hatch and Rodger Slack (Fig. 20–23b).
In plants of tropical origin, the first intermediate
into which
14
CO
2
is fixed is oxaloacetate, a four-carbon
compound. This reaction, which occurs in the cytosol of
leaf mesophyll cells, is catalyzed by phosphoenolpyru-
vate carboxylase, for which the substrate is HCO
3
H11002
,
not CO
2
. The oxaloacetate thus formed is either reduced
to malate at the expense of NADPH (as shown in Fig.
20–23b) or converted to aspartate by transamination:
Oxaloacetate H11001 H9251-amino acid 88n L-aspartate H11001 H9251-keto acid
20.2 Photorespiration and the C
4
and CAM Pathways 769
The malate or aspartate formed in the mesophyll cells
then passes into neighboring bundle-sheath cells
through plasmodesmata, protein-lined channels that
connect two plant cells and provide a path for move-
ment of metabolites and even small proteins between
cells. In the bundle-sheath cells, malate is oxidized and
decarboxylated to yield pyruvate and CO
2
by the action
of malic enzyme, reducing NADP
H11001
. In plants that use
aspartate as the CO
2
carrier, aspartate arriving in
bundle-sheath cells is transaminated to form oxaloac-
etate and reduced to malate, then the CO
2
is released
by malic enzyme or PEP carboxykinase. As labeling ex-
periments show, the free CO
2
released in the bundle-
sheath cells is the same CO
2
molecule originally fixed
into oxaloacetate in the mesophyll cells. This CO
2
is now
fixed again, this time by rubisco, in exactly the same re-
action that occurs in C
3
plants: incorporation of CO
2
into
C-1 of 3-phosphoglycerate.
The pyruvate formed by decarboxylation of malate
in bundle-sheath cells is transferred back to the meso-
phyll cells, where it is converted to PEP by an unusual
enzymatic reaction catalyzed by pyruvate phosphate
dikinase (Fig. 20–23b). This enzyme is called a dikinase
because two different molecules are simultaneously
phosphorylated by one molecule of ATP: pyruvate to
PEP, and phosphate to pyrophosphate. The pyro-
phosphate is subsequently hydrolyzed to phosphate, so
two high-energy phosphate groups of ATP are used in
regenerating PEP. The PEP is now ready to receive an-
other molecule of CO
2
in the mesophyll cell.
The PEP carboxylase of mesophyll cells has a high
affinity for HCO
3
H11002
(which is favored relative to CO
2
in
aqueous solution and can fix CO
2
more efficiently than
can rubisco). Unlike rubisco, it does not use O
2
as an
alternative substrate, so there is no competition be-
tween CO
2
and O
2
. The PEP carboxylase reaction, then,
serves to fix and concentrate CO
2
in the form of malate.
Release of CO
2
from malate in the bundle-sheath cells
yields a sufficiently high local concentration of CO
2
for
rubisco to function near its maximal rate, and for sup-
pression of the enzyme’s oxygenase activity.
Once CO
2
is fixed into 3-phosphoglycerate in the
bundle-sheath cells, the other reactions of the Calvin cy-
cle take place exactly as described earlier. Thus in C
4
plants, mesophyll cells carry out CO
2
assimilation by the
C
4
pathway and bundle-sheath cells synthesize starch
and sucrose by the C
3
pathway.
Three enzymes of the C
4
pathway are regulated by
light, becoming more active in daylight. Malate dehy-
drogenase is activated by the thioredoxin-dependent re-
duction mechanism shown in Figure 20–19; PEP car-
boxylase is activated by phosphorylation of a Ser
residue; and pyruvate phosphate dikinase is activated
by dephosphorylation. In the latter two cases, the de-
tails of how light effects phosphorylation or dephos-
phorylation are not known.
8885d_c20_751–786 2/18/04 1:56 PM Page 769 mac76 mac76:385_reb:
The pathway of CO
2
assimilation has a greater en-
ergy cost in C
4
plants than in C
3
plants. For each mol-
ecule of CO
2
assimilated in the C
4
pathway, a molecule
of PEP must be regenerated at the expense of two high-
energy phosphate groups of ATP. Thus C
4
plants need
five ATP molecules to assimilate one molecule of CO
2
,
whereas C
3
plants need only three (nine per triose
phosphate). As the temperature increases (and the
affinity of rubisco for CO
2
decreases, as noted above),
a point is reached (at about 28 to 30 H11034C) at which the
gain in efficiency from the elimination of photorespira-
tion more than compensates for this energetic cost. C
4
plants (crabgrass, for example) outgrow most C
3
plants
during the summer, as any experienced gardener can
attest.
In CAM Plants, CO
2
Capture and Rubisco Action
Are Temporally Separated
Succulent plants such as cactus and pineapple, which
are native to very hot, very dry environments, have an-
other variation on photosynthetic CO
2
fixation, which
reduces loss of water vapor through the pores (stom-
ata) by which CO
2
and O
2
must enter leaf tissue. In-
stead of separating the initial trapping of CO
2
and its
fixation by rubisco across space (as do the C
4
plants),
they separate these two events over time. At night, when
the air is cooler and moister, the stomata open to allow
entry of CO
2
, which is then fixed into oxaloacetate by
PEP carboxylase. The oxaloacetate is reduced to malate
and stored in the vacuoles, to protect cytosolic and plas-
tid enzymes from the low pH produced by malic acid
dissociation. During the day the stomata close, pre-
venting the water loss that would result from high day-
time temperatures, and the CO
2
trapped overnight in
malate is released as CO
2
by the NADP-linked malic en-
zyme. This CO
2
is now assimilated by the action of ru-
bisco and the Calvin cycle enzymes. Because this
method of CO
2
fixation was first discovered in
stonecrops, perennial flowering plants of the family
Crassulaceae, it is called crassulacean acid metabolism,
and the plants are called CAM plants.
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria770
FIGURE 20–23 Carbon assimilation in C
4
plants. The C
4
pathway, in-
volving mesophyll cells and bundle-sheath cells, predominates in
plants of tropical origin. (a) Electron micrograph showing chloroplasts
of adjacent mesophyll and bundle-sheath cells. The bundle-sheath cell
contains starch granules. Plasmodesmata connecting the two cells are
visible. (b) The C
4
pathway of CO
2
assimilation, which occurs through
a four-carbon intermediate.
NADPH + H
+
3-PhosphoglycerateRibulose
1,5-bisphosphate
AMP
ATP
P
i
Mesophyll
cell
PEP
Pyruvate
pyruvate
phosphate
dikinase
PP
i
+
+
Malate
NADP
+
malic
enzyme
Triose phosphates
Bundle-sheath
cell
Plasma membranes
Oxaloacetate
Malate
malate
dehydrogenase
NADPH
+
H
+
P
i
NADP
+
Pyruvate
–
Plasmodesmata
PEP
carboxylase
H
2
O
(in air)
H
+
CO
2
HCO
3
CO
2
(b)
(a)
Mesophyll
cell
Plasmodesmata
Bundle-
sheath
cell
8885d_c20_770 2/20/04 12:02 PM Page 770 mac76 mac76:385_reb:
SUMMARY 20.2 Photorespiration and the C
4
and CAM Pathways
■ When rubisco uses O
2
rather than CO
2
as
substrate, the 2-phosphoglycolate so formed is
disposed of in an oxygen-dependent pathway.
The result is increased consumption of
O
2
—photorespiration or, more accurately, the
oxidative photosynthetic carbon cycle or C
2
cycle. The 2-phosphoglycolate is converted to
glyoxylate, to glycine, and then to serine in a
pathway that involves enzymes in the
chloroplast stroma, the peroxisome, and the
mitochondrion.
■ In C
4
plants, the carbon-assimilation pathway
minimizes photorespiration: CO
2
is first fixed in
mesophyll cells into a four-carbon compound,
which passes into bundle-sheath cells and
releases CO
2
in high concentrations. The
released CO
2
is fixed by rubisco, and the
remaining reactions of the Calvin cycle occur
as in C
3
plants.
■ In CAM plants, CO
2
is fixed into malate in the
dark and stored in vacuoles until daylight,
when the stomata are closed (minimizing water
loss) and malate serves as a source of CO
2
for
rubisco.
20.3 Biosynthesis of Starch and Sucrose
During active photosynthesis in bright light, a plant leaf
produces more carbohydrate (as triose phosphates)
than it needs for generating energy or synthesizing pre-
cursors. The excess is converted to sucrose and trans-
ported to other parts of the plant, to be used as fuel or
stored. In most plants, starch is the main storage form,
but in a few plants, such as sugar beet and sugarcane,
sucrose is the primary storage form. The synthesis of
sucrose and starch occurs in different cellular com-
partments (cytosol and plastids, respectively), and
these processes are coordinated by a variety of regula-
tory mechanisms that respond to changes in light level
and photosynthetic rate.
ADP-Glucose Is the Substrate for Starch Synthesis
in Plant Plastids and for Glycogen Synthesis
in Bacteria
Starch, like glycogen, is a high molecular weight poly-
mer of D-glucose in (H92511n4) linkage. It is synthesized in
chloroplasts for temporary storage as one of the stable
end products of photosynthesis, and for long-term stor-
age it is synthesized in the amyloplasts of the nonpho-
tosynthetic parts of plants—seeds, roots, and tubers
(underground stems).
The mechanism of glucose activation in starch syn-
thesis is similar to that in glycogen synthesis. An acti-
vated nucleotide sugar, in this case ADP-glucose, is
formed by condensation of glucose 1-phosphate with ATP
in a reaction made essentially irreversible by the pres-
ence in plastids of inorganic pyrophosphatase (p. 502).
Starch synthase then transfers glucose residues from
ADP-glucose to preexisting starch molecules. Although it
has generally been assumed that glucose is added to the
nonreducing end of starch, as in glycogen synthesis
(see Fig. 15–8), evidence now suggests that starch syn-
thase has two equivalent active sites that alternate in in-
serting a glucosyl residue onto the reducing end of the
growing chain. This end remains covalently attached to
the enzyme, first at one active site, then at the other
(Fig. 20–24). Attachment to one active site effectively
activates the reducing end of the growing chain for nu-
cleophilic displacement of the enzyme by the attacking
C-4 hydroxyl of a glucosyl moiety bound to the other ac-
tive site, forming the (H92511n4) linkage characteristic of
starch.
The amylose of starch is unbranched, but amy-
lopectin has numerous (H92511n6)-linked branches (see
Fig. 7–15). Chloroplasts contain a branching enzyme,
similar to glycogen-branching enzyme (see Fig. 15–9),
that introduces the (H92511n6) branches of amylopectin.
Taking into account the hydrolysis by inorganic py-
rophosphatase of the PP
i
produced during ADP-glucose
synthesis, the overall reaction for starch formation from
glucose 1-phosphate is
Starch
n
H11001 glucose 1-phosphate H11001 ATP 88n
starch
nH110011
H11001 ADP H11001 2P
i
H9004GH11032H11034H11005H1100250 kJ/mol
Starch synthesis is regulated at the level of ADP-glucose
formation, as discussed below.
Many types of bacteria store carbohydrate in the
form of glycogen (essentially highly branched starch),
which they synthesize in a reaction analogous to that
catalyzed by glycogen synthase in animals. Bacteria, like
plant plastids, use ADP-glucose as the activated form of
glucose, whereas animal cells use UDP-glucose. Again,
the similarity between plastid and bacterial metabolism
is consistent with the endosymbiont hypothesis for the
origin of organelles (see Fig. 1–36).
UDP-Glucose Is the Substrate for Sucrose Synthesis
in the Cytosol of Leaf Cells
Most of the triose phosphate generated by CO
2
fixation
in plants is converted to sucrose (Fig. 20–25) or starch.
In the course of evolution, sucrose may have been se-
lected as the transport form of carbon because of its un-
usual linkage between the anomeric C-1 of glucose and
the anomeric C-2 of fructose. This bond is not hydrolyzed
by amylases or other common carbohydrate-cleaving
20.3 Biosynthesis of Starch and Sucrose 771
8885d_c20_751–786 2/18/04 1:56 PM Page 771 mac76 mac76:385_reb:
enzymes, and the unavailability of the anomeric carbons
prevents sucrose from reacting nonenzymatically (as
does glucose) with amino acids and proteins.
Sucrose is synthesized in the cytosol, beginning
with dihydroxyacetone phosphate and glyceraldehyde
3-phosphate exported from the chloroplast. After con-
densation of two triose phosphates to form fructose 1,6-
bisphosphate (catalyzed by aldolase), hydrolysis by
fructose 1,6-bisphosphatase yields fructose 6-phosphate.
Sucrose 6-phosphate synthase then catalyzes the
reaction of fructose 6-phosphate with UDP-glucose
to form sucrose 6-phosphate (Fig. 20–25). Finally,
sucrose 6-phosphate phosphatase removes the phos-
phate group, making sucrose available for export to
other tissues. The reaction catalyzed by sucrose 6-phos-
phate synthase is a low-energy process (H9004GH11032H11034 H11005 H110025.7
kJ/mol), but the hydrolysis of sucrose 6-phosphate to
sucrose is sufficiently exergonic (H9004GH11032H11034 H11005 H1100216.5 kJ/mol)
to make the overall synthesis of sucrose essentially
irreversible. Sucrose synthesis is regulated and closely
coordinated with starch synthesis, as we shall see.
One remarkable difference between the cells of
plants and animals is the absence in the plant cell cy-
tosol of the enzyme inorganic pyrophosphatase, which
catalyzes the reaction
PP
i
H11001 H
2
O 88n 2P
i
H9004GH11032H11034 H11005 H1100219.2 kJ/mol
For many biosynthetic reactions that liberate PP
i
, py-
rophosphatase activity makes the process more favor-
able energetically, tending to make these reactions ir-
reversible. In plants, this enzyme is present in plastids
but absent from the cytosol. As a result, the cytosol of
leaf cells contains a substantial concentration of PP
i
—
enough (~0.3 mM) to make reactions such as that cat-
alyzed by UDP-glucose pyrophosphorylase (Fig. 15–7)
readily reversible. Recall from Chapter 14 (p. 527) that
the cytosolic isozyme of phosphofructokinase in plants
uses PP
i
, not ATP, as the phosphoryl donor.
Conversion of Triose Phosphates to Sucrose
and Starch Is Tightly Regulated
Triose phosphates produced by the Calvin cycle in
bright sunlight, as we have noted, may be stored tem-
porarily in the chloroplast as starch, or converted to su-
crose and exported to nonphotosynthetic parts of the
plant, or both. The balance between the two processes
is tightly regulated, and both must be coordinated with
the rate of carbon fixation. Five-sixths of the triose
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria772
Starch
ADP-glucose
–
ADP
X
a
–
X
b
OH
Each of the two reactive groups (X
a
, X
b
) at the
active site of starch synthase makes a nucleophilic
attack on ADP-glucose, displacing ADP and forming
a covalent attachment to C-1 of the glucose unit.
The hydroxyl at C-4 of glucose 3 displaces X
b
from the disaccharide, forming a trisaccharide
attached to X
a
. X
b
, now free, acquires glucose
residue 4 from another ADP-glucose.
The hydroxyl at C-4 of glucose 4 displaces
X
a
, forming a tetrasaccharide, with its reducing
end covalently attached to X
b
.
Many repetitions of this sequence extend the
oligosaccharide, adding glucose residues at its
reducing end, with X
a
and X
b
alternately
carrying the growing starch chain. When the chain
reaches an appropriate length, it is separated
from starch synthase.
1
ADP
X
a
X
b
3
ADP5
ADP2
2
1
:
OH
:
OH
X
a
X
b
ADP4
2 1
3
:
OH
Nonreducing end
5
:
ADP6
X
a
X
b
3
4
2 1
X
a
X
b
34 2 1
The bond holding glucose residue 1 to X
a
undergoes nucleophilic attack by the OH
at C-4 of glucose residue 2 on X
b
, forming an
a(1 4)-disaccharide of residues 2 and 1. This
remains attached through glucose 2 to X
b
. X
a
,
now free, displaces ADP from another ADP-
glucose and becomes attached to glucose 3.
Starch
synthase
FIGURE 20–24 Starch synthesis. Starch synthesis proceeds by a two-
site insertion mechanism, with ADP-glucose as the initial glucosyl
donor. The two identical active sites on starch synthase alternate in
displacing the growing chain from each other, and new glucosyl units
are inserted at the reducing end of the growing chain.
8885d_c20_751–786 2/18/04 1:56 PM Page 772 mac76 mac76:385_reb:
phosphate formed in the Calvin cycle must be recycled
to ribulose 1,5-bisphosphate (Fig. 20–14); if more than
one-sixth of the triose phosphate is drawn out of the
cycle to make sucrose and starch, the cycle will slow or
stop. However, insufficient conversion of triose phos-
phate to starch or sucrose would tie up phosphate, leav-
ing a chloroplast deficient in P
i
, which is also essential
for operation of the Calvin cycle.
The flow of triose phosphates into sucrose is reg-
ulated by the activity of fructose 1,6-bisphosphatase
(FBPase-1) and the enzyme that effectively reverses its
action, PP
i
-dependent phosphofructokinase (PP-PFK-1;
p. 527). These enzymes are therefore critical points for
determining the fate of triose phosphates produced by
photosynthesis. Both enzymes are regulated by fructose
2,6-bisphosphate (F2,6BP), which inhibits FBPase-1
and stimulates PP-PFK-1. In vascular plants, the con-
centration of F2,6BP varies inversely with the rate of
photosynthesis (Fig. 20–26). Phosphofructokinase-2,
responsible for F2,6BP synthesis, is inhibited by dihy-
droxyacetone phosphate or 3-phosphoglycerate and
stimulated by fructose 6-phosphate and P
i
. During ac-
tive photosynthesis, dihydroxyacetone phosphate is
produced and P
i
is consumed, resulting in inhibition of
PFK-2 and lowered concentrations of F2,6BP. This
20.3 Biosynthesis of Starch and Sucrose 773
O
UDP-glucose
sucrose
6-phosphate
synthase
1
HOCH
2
HO
H
OH
H
OHO
P
UDP
CH
2
O
HOH
H
H
HOH
Fructose 6-phosphate
UDP
H
HO
O
2
HOCH
2
H
HH
1
HO
O
CH
2
OH
O
HOH
H
H
HOH
Sucrose
HOH
O
sucrose
6-phosphate
phosphatase
2
HOCH
2
H
OH
HH
1
HO
P
i
CH
2
OH
O
HOH
H
H
HOH
Sucrose 6-phosphate
H
O
H11001
O
CH
2
OH
2
H
HO
HO
PCH
2
O
CH
2
OH
FIGURE 20–25 Sucrose synthesis. Sucrose is synthesized from UDP-
glucose and fructose 6-phosphate, which are synthesized from triose
phosphates in the plant cell cytosol by pathways shown in Figures
15–7 and 20–9. The sucrose 6-phosphate synthase of most plant
species is allosterically regulated by glucose 6-phosphate and P
i
.
FIGURE 20–26 Fructose 2,6-bisphosphate as regulator of sucrose syn-
thesis. The concentration of the allosteric regulator fructose 2,6-
bisphosphate in plant cells is regulated by the products of photosyn-
thetic carbon assimilation and by P
i
. Dihydroxyacetone phosphate and
3-phosphoglycerate produced by CO
2
assimilation inhibit phospho-
fructokinase-2 (PFK-2), the enzyme that synthesizes the regulator; P
i
stimulates PFK-2. The concentration of the regulator is therefore
inversely proportional to the rate of photosynthesis. In the dark, the
concentration of fructose 2,6-bisphosphate increases and stimulates
the glycolytic enzyme PP
i
-dependent phosphofructokinase-1 (PP-PFK-
1), while inhibiting the gluconeogenic enzyme fructose 1,6-
bisphosphatase (FBPase-1). When photosynthesis is active (in the light),
the concentration of the regulator drops and the synthesis of fructose
6-phosphate and sucrose is favored.
Chloroplast
Fructose
2,6-bisphosphate
ADP
ATP
PFK-2 FBPase-2
H
2
O
P
i
Dark
No photosynthesis
P
i
Light
Active photosynthesis
3-Phosphoglycerate
and dihydroxyacetone
phosphate
Fructose
6-phosphate
PP
i
P
i
P
i
FBPase-1
Triose
phosphate
Cytosol
Dark Dark
Fructose
1,6-bisphosphate
PP-PFK-1
Sucrose
Glycolysis
Gluconeogenesis
8885d_c20_751–786 2/18/04 1:56 PM Page 773 mac76 mac76:385_reb:
favors greater flux of triose phosphate into fructose 6-
phosphate formation and sucrose synthesis. With this
regulatory system, sucrose synthesis occurs when the
level of triose phosphate produced by the Calvin cycle
exceeds that needed to maintain the operation of the
cycle.
Sucrose synthesis is also regulated at the level of
sucrose 6-phosphate synthase, which is allosterically
activated by glucose 6-phosphate and inhibited by P
i
.
This enzyme is further regulated by phosphorylation
and dephosphorylation; a protein kinase phosphory-
lates the enzyme on a specific Ser residue, making it
less active, and a phosphatase reverses this inactiva-
tion by removing the phosphate (Fig. 20–27). Inhibi-
tion of the kinase by glucose 6-phosphate, and of the
phosphatase by P
i
, strengthens the effects of these two
compounds on sucrose synthesis. When hexose phos-
phates are abundant, sucrose 6-phosphate synthase is
activated by glucose 6-phosphate; when P
i
is elevated
(as when photosynthesis is slow), sucrose synthesis
is slowed. During active photosynthesis, triose phos-
phates are converted to fructose 6-phosphate, which
is rapidly equilibrated with glucose 6-phosphate by
phosphohexose isomerase. Because the equilibrium
lies far toward glucose 6-phosphate, as soon as fruc-
tose 6-phosphate accumulates, the level of glucose
6-phosphate rises and sucrose synthesis is stimulated.
The key regulatory enzyme in starch synthesis is
ADP-glucose pyrophosphorylase (Fig. 20–28); it is
activated by 3-phosphoglycerate (which accumulates
during active photosynthesis) and inhibited by P
i
(which
accumulates when light-driven condensation of ADP
and P
i
slows). When sucrose synthesis slows, 3-phos-
phoglycerate formed by CO
2
fixation accumulates, acti-
vating this enzyme and stimulating the synthesis of
starch.
SUMMARY 20.3 Biosynthesis of Starch
and Sucrose
■ Starch synthase in chloroplasts and amyloplasts
catalyzes the addition of single glucose
residues, donated by ADP-glucose, to the
reducing end of a starch molecule by a
two-step insertion mechanism. Branches in
amylopectin are introduced by a second
enzyme.
■ Sucrose is synthesized in the cytosol in two
steps from UDP-glucose and fructose
1-phosphate.
■ The partitioning of triose phosphates between
sucrose synthesis and starch synthesis is
regulated by fructose 2,6-bisphosphate
(F2,6BP), an allosteric effector of the enzymes
that determine the level of fructose
6-phosphate. F2,6BP concentration varies
inversely with the rate of photosynthesis, and
F2,6BP inhibits the synthesis of fructose
6-phosphate, the precursor to sucrose.
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria774
Glucose
6-phosphate
Glucose
6-phosphate
(less active)
(more active)
CH
2
O
SPS
kinase
H
2
O
P
i
P
i
P
i
SPS
phosphatase
ADP
ATP
Sucrose
6-phosphate
synthase
Sucrose
6-phosphate
synthase
CH
2
OH
P
Photosynthesis
Bright lightDim light or darkness
ADP-glucose + PP
i
ADP-glucose
pyrophosphorylase
Glucose 1-
phosphate + ATP
3-Phosphoglycerate
slow
ADP + P
i
ATP
FIGURE 20–27 Regulation of sucrose phosphate synthase by phos-
phorylation. A protein kinase (SPS kinase) specific for sucrose phos-
phate synthase (SPS) phosphorylates a Ser residue in SPS, inactivating
it; a specific phosphatase (SPS phosphatase) reverses this inhibition.
The kinase is inhibited allosterically by glucose 6-phosphate, which
also activates SPS allosterically. The phosphatase is inhibited by Pi,
which also inhibits SPS directly. Thus when the concentration of glu-
cose 6-phosphate is high as a result of active photosynthesis, SPS is
activated and produces sucrose phosphate. A high P
i
concentration,
which occurs when photosynthetic conversion of ADP to ATP is slow,
inhibits sucrose phosphate synthesis.
FIGURE 20–28 Regulation of ADP-glucose phosphorylase by 3-
phosphoglycerate and P
i
. This enzyme, which produces the precursor
for starch synthesis, is rate-limiting in starch production. The enzyme
is stimulated allosterically by 3-phosphoglycerate (3-PGA) and inhib-
ited by P
i
; in effect, the ratio [3-PGA]/[P
i
], which rises with increas-
ing rates of photosynthesis, controls starch synthesis at this step.
8885d_c20_751–786 2/18/04 1:56 PM Page 774 mac76 mac76:385_reb:
20.4 Synthesis of Cell Wall
Polysaccharides: Plant Cellulose
and Bacterial Peptidoglycan
Cellulose is a major constituent of plant cell walls, pro-
viding strength and rigidity and preventing the swelling
of the cell and rupture of the plasma membrane that
might result when osmotic conditions favor water entry
into the cell. Each year, worldwide, plants synthesize
more than 10
11
metric tons of cellulose, making this
simple polymer one of the most abundant compounds in
the biosphere. The structure of cellulose is simple: lin-
ear polymers of thousands of H9252(1n4)-linked D-glucose
units, assembled into bundles of about 36 chains, which
aggregate side by side to form a microfibril (Fig. 20–29).
The biosynthesis of cellulose is less well understood
than that of glycogen or starch. As a major component
of the plant cell wall, cellulose must be synthesized from
intracellular precursors but deposited and assembled
outside the plasma membrane. The enzymatic machin-
ery for initiation, elongation, and export of cellulose
chains is more complicated than that needed to syn-
thesize starch or glycogen (which are not exported).
Bacteria face a similar set of problems when they syn-
thesize the complex polysaccharides that make up their
cell walls, and they may employ some of the same mech-
anisms to solve these problems.
Cellulose Is Synthesized by Supramolecular
Structures in the Plasma Membrane
The complex enzymatic machinery that assembles cel-
lulose chains spans the plasma membrane, with one part
positioned to bind the substrate, UDP-glucose, in the
cytosol and another part extending to the outside, re-
sponsible for elongating and crystallizing cellulose mol-
ecules in the extracellular space. Freeze-fracture elec-
tron microscopy shows these terminal complexes,
also called rosettes, to be composed of six large parti-
cles arranged in a regular hexagon (Fig. 20–30). Several
proteins, including the catalytic subunit of cellulose
synthase, make up the terminal complex. Cellulose
synthase has not been isolated in its active form, but its
amino acid sequence has been determined from the nu-
cleotide sequence of the gene that encodes it. From the
primary structure we can use hydropathy plots (see Fig.
11–11) to deduce that the enzyme has eight trans-
membrane segments, connected by short loops on the
outside, and several longer loops exposed to the cytosol.
Much of the recent progress in understanding cellulose
synthesis stems from genetic and molecular genetic
studies of the plant Arabidopsis thaliana, which is es-
pecially amenable to genetic dissection and whose
genome has been sequenced.
20.4 Synthesis of Cell Wall Polysaccharides 775
Cellulose
microfibrils in
plant cell wall
Microfibril
b(1 4)
Cellulose
chains
H
HOH
O O
H
CH
2
OH
HOH
O
H
HOH
H
CH
2
OH
HOH
O
O
FIGURE 20–29 Cellulose structure. The plant cell wall is made up in
part of cellulose molecules arranged side by side to form paracrys-
talline arrays—cellulose microfibrils. Many microfibrils combine to
form a cellulose fiber, seen in the scanning electron microscope as a
structure 5 to 12 nm in diameter, laid down on the cell surface in sev-
eral layers distinguishable by the different orientations of their fibers.
FIGURE 20–30 Rosettes. The outside surface of the plant plasma mem-
brane in a freeze-fractured sample, viewed here with electron mi-
croscopy, contains many hexagonal arrays of particles about 10 nm
in diameter, believed to be composed of cellulose synthase molecules
and associated enzymes.
8885d_c20_751–786 2/18/04 1:56 PM Page 775 mac76 mac76:385_reb:
New cellulose chains appear to be initiated by the
formation of a lipid-linked intermediate unlike anything
involved in starch or glycogen synthesis. Glucose is trans-
ferred from UDP-glucose to a membrane lipid, probably
the plant sterol sitosterol (Fig. 20–31), on the inner face
of the plasma membrane. Here, intracellular cellulose
synthase adds several more glucose residues to the first
one, in (H92521n4) linkage, forming a short oligosaccharide
chain attached to the sitosterol (sitosterol dextrin). Next,
the whole sitosterol dextrin flips across to the outer face
of the plasma membrane, where most of the polysac-
charide chain is removed by endo-1,4-H9252-glucanase. The
shortened sitosterol dextrin primer now associates, per-
haps covalently, with another form of cellulose synthase.
Presumably this entire process occurs in the rosettes.
Whether each of the 36 cellulose chains is initiated on its
own lipid primer, or the primer recycles to start a num-
ber of chains, is not yet clear. In either case, the second
form of cellulose synthase extends the polymer to 500 to
15,000 glucose units, extruding it onto the outer surface
of the cell. The action of the enzyme is processive: one
enzyme molecule adds many glucose units before re-
leasing the growing cellulose chain. The direction of chain
growth (whether addition occurs at the reducing end or
at the nonreducing end) has not been established.
The finished cellulose is in the form of crystalline
microfibrils (Fig. 20–29), each consisting of 36 separate
cellulose chains lying side by side, all with the same
(parallel) orientation of nonreducing and reducing ends.
It seems likely that each particle in the rosette synthe-
sizes six separate cellulose chains simultaneously and in
parallel with the chains made by the other five particles,
so that 36 polymers arrive together on the outer surface
of the cell, already aligned and ready to crystallize as a
microfibril of the cell wall. When the 36 polymers reach
some critical length, their synthesis is terminated by an
unknown mechanism; crystallization into a microfibril
follows.
In addition to its catalytic subunit, cellulose syn-
thase may have subunits that mediate extrusion of the
polysaccharide chain (the pore subunit) and crystal-
lization of the polysaccharide chains outside the cell
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria776
Sitosterol
OH
UDP
Glucose
UDP
Cytosol
short oligosaccharide
primer retained;
sitosterol flips
to cytosolic face
Lengthening
molecule
exits cell
Extracellular space
endo-1,4-b-
glucanase
cellulose
synthase
:
UDP
cellulose
synthase
UDP
n
((
n
((
O
O Sitosterol
dextrin
1
2
3
molecule flips
from inner to
outer leaflet
4
5
7
6
UDP
UDP
n
((
n
((
FIGURE 20–31 Lipid primer for cellulose synthesis. This proposed
pathway begins with 1 the transfer of a glucosyl residue from UDP-
glucose to a lipid “primer” (probably sitosterol) in the inner leaflet of
the plasma membrane. After this initiation, 2 the chain of carbohy-
drate is elongated by transfer of glucosyl residues from UDP-glucose,
until 3 a critical length of oligosaccharide is reached. 4 The sitos-
terol with its attached oligosaccharide now flips from the inner leaflet
to the outer leaflet. 5 An endo-1,4-H9252-glucanase separates the grow-
ing chain from a short oligonucleotide still attached to the lipid. As it
is pushed out of the cell, 6 the lipid-free polymer of glucosyl residues
(the glucan acceptor) is further extended by the addition of glucosyl
residues from UDP-glucose, catalyzed by cellulose synthase. 7 The
lipid-linked oligosaccharide returns to serve as the primer for another
chain of cellulose.
8885d_c20_751–786 2/18/04 1:56 PM Page 776 mac76 mac76:385_reb:
(the crystallization subunit). The potent herbicide CGA
325H11032615, which specifically inhibits cellulose synthesis,
causes rosettes to fall apart; the small amount of cellu-
lose still synthesized remains tightly, perhaps covalently,
bound to the catalytic subunit of cellulose synthase. The
inhibitor may act by dissociating the catalytic subunit
from the pore and crystallization subunits, preventing
the later stages of cellulose synthesis.
The UDP-glucose used for cellulose synthesis is
generated from sucrose produced during photosynthe-
sis, by the reaction catalyzed by sucrose synthase
(named for the reverse reaction):
Sucrose H11001 UDP 88n UDP-glucose H11001 fructose
In one proposed model, cellulose synthase spans the
plasma membrane and uses cytosolic UDP-glucose as the
precursor for extracellular cellulose synthesis. In another,
a membrane-bound form of sucrose synthase forms a
complex with cellulose synthase, feeding UDP-glucose
from sucrose directly into cell wall synthesis (Fig. 20–32).
In the activated precursor of cellulose (UDP-
glucose), the glucose is H9251-linked to the nucleotide, but
in the product (cellulose), glucose residues are H9252(1n4)-
linked, so there is an inversion of configuration at the
anomeric carbon (C-1) as the glycosidic bond forms.
Glycosyltransferases that invert configuration are gen-
erally assumed to use a single-displacement mechanism,
with nucleophilic attack by the acceptor species at the
anomeric carbon of the donor sugar (UDP-glucose).
Certain bacteria (Acetobacter, Agrobacteria, Rhi-
zobia, and Sarcina) and many simple eukaryotes also
carry out cellulose synthesis, apparently by a mecha-
nism similar to that in plants. If the bacteria use a mem-
brane lipid to initiate new chains, it cannot be a sterol—
bacteria do not contain sterols.
Lipid-Linked Oligosaccharides Are Precursors
for Bacterial Cell Wall Synthesis
Like plants, many bacteria have thick, rigid extracellu-
lar walls that protect them from osmotic lysis. The pep-
tidoglycan that gives bacterial envelopes their strength
and rigidity is an alternating linear copolymer of N-
acetylglucosamine (GlcNAc) and N-acetylmuramic acid
(Mur2Ac), linked by (H92521n4) glycosidic bonds and
cross-linked by short peptides attached to the Mur2Ac
(Fig. 20–33). During assembly of the polysaccharide
backbone of this complex macromolecule, both GlcNAc
and Mur2Ac are activated by attachment of a uridine
20.4 Synthesis of Cell Wall Polysaccharides 777
Crystallization
subunit
Sucrose
Cytosol
Fructose
UDP
UDP-
glucose
Cellulose
synthase
Sucrose
synthase
Pore
subunit
Catalytic
subunit
FIGURE 20–32 A plausible model for the structure of cellulose syn-
thase. The enzyme complex includes a catalytic subunit with eight
transmembrane segments and several other subunits that are presumed
to act in threading cellulose chains through the catalytic site and out
of the cell, and in the crystallization of 36 cellulose strands into the
paracrystalline microfibrils shown in Figure 20–29.
FIGURE 20–33 Peptidoglycan structure. This is the peptidoglycan of
the cell wall of Staphylococcus aureus, a gram-positive bacterium.
Peptides (strings of colored spheres) covalently link N-acetylmuramic
acid residues in neighboring polysaccharide chains. Note the mixture
of L and D amino acids in the peptides. Gram-positive bacteria such
as S. aureus have a pentaglycine chain in the cross-link. Gram-negative
bacteria, such as E. coli, lack the pentaglycine; instead, the terminal
D-Ala residue of one tetrapeptide is attached directly to a neighbor-
ing tetrapeptide through either L-Lys or a lysine-like amino acid, di-
aminopimelic acid.
Staphylococcus
aureus
Reducing
end
L-Ala
D-Glu
L-Lys
D-Ala
N-Acetylglucosamine
(GlcNAc)
Pentaglycine
cross-link
(b14)
N-Acetylmuramic
acid (Mur2Ac)
8885d_c20_751–786 2/18/04 1:56 PM Page 777 mac76 mac76:385_reb:
nucleotide at their anomeric carbons. First, GlcNAc 1-
phosphate condenses with UTP to form UDP-GlcNAc
(Fig. 20–34, step 1 ), which reacts with phospho-
enolpyruvate to form UDP-Mur2Ac (step 2 ); five amino
acids are then added (step 3 ). The Mur2Ac-pentapep-
tide moiety is transferred from the uridine nucleotide
to the membrane lipid dolichol, a long-chain isoprenoid
alcohol (see Fig. 10–22f) (step 4 ), and a GlcNAc
residue is donated by UDP-GlcNAc (step 5 ). In many
bacteria, five glycines are added in peptide linkage to
the amino group of the Lys residue of the pentapeptide
(step 6 ). Finally, this disaccharide decapeptide is added
to the nonreducing end of an existing peptidoglycan
molecule (step 7 ). A transpeptidation reaction cross-
links adjacent polysaccharide chains (step 8 ), con-
tributing to a huge, strong, macromolecular wall around
the bacterial cell. Many of the most effective antibiotics
in use today act by inhibiting reactions in the synthesis
of the peptidoglycan (Box 20–1).
Many other oligosaccharides and polysaccharides
are synthesized by similar routes in which sugars are ac-
tivated for subsequent reactions by attachment to nu-
cleotides. In the glycosylation of proteins, for example
(see Fig. 27–34), the precursors of the carbohydrate
moieties include sugar nucleotides and lipid-linked
oligosaccharides.
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria778
+ UTPGlcNAc-1 UDP-GlcNAc + PP
i
PEP
NADPH
NADP
+
UDP-Mur2Ac
UDP-GlcNAc
L-Alanine
1
2
3
5
6
7
D-Glutamate
L-Lysine
D-Ala–D-Alanine
Dolichol
UDP
PP
4
Dolichol
UMP
P
P
Dolichol
5 L-Glycine
D-Alanine
8
transpeptidase
PP DolicholPP
Dolichol
Existing peptidoglycan “primer”
Another peptidoglycan chain
Penicillins
PP
Mature, fully cross-linked
peptidoglycan
FIGURE 20–34 Synthesis of bacterial peptidoglycan.
In the early steps of this pathway ( 1 through 4 ),
N-acetylglucosamine (GlcNAc) and N-acetylmuramic
acid (Mur2Ac) are activated by attachment of a uridine
nucleotide (UDP) to their anomeric carbons and, in
the case of Mur2Ac, of a long-chain isoprenyl alcohol
(dolichol) through a phosphodiester bond. These acti-
vating groups participate in the formation of glycosidic
linkages; they serve as excellent leaving groups. After
5 , 6 assembly of a disaccharide with a peptide side
chain (10 amino acid residues), 7 this precursor is
transferred to the nonreducing end of an existing pepti-
doglycan chain, which serves as a primer for the
polymerization reaction. Finally, 8 in a transpeptida-
tion reaction between the peptide side chains on two
different peptidoglycan molecules, a Gly residue at the
end of one chain displaces a terminal D-Ala in the
other chain, forming a cross-link. This transpeptidation
reaction is inhibited by the penicillins, which kill
bacteria by weakening their cell walls.
8885d_c20_751–786 2/18/04 1:56 PM Page 778 mac76 mac76:385_reb:
20.4 Synthesis of Cell Wall Polysaccharides 779
BOX 20–1 BIOCHEMISTRY IN MEDICINE
The Magic Bullet versus the Bulletproof Vest:
Penicillin and H9252-Lactamase
Because peptidoglycans are unique to bacterial cell
walls, with no known homologous structures in mam-
mals, the enzymes responsible for their synthesis are
ideal targets for antibiotic action. Antibiotics that hit
specific bacterial targets are sometimes called “magic
bullets.” Penicillin and its many synthetic analogs have
been used to treat bacterial infections since these
drugs came into wide application in World War II.
Penicillins and related antibiotics contain the H9252-
lactam ring (Fig. 1), variously modified. All penicillins
have a thiazolidine ring attached to the H9252-lactam, but
they differ in the substituent at position 6, which ac-
counts for the different pharmacological properties of
the penicillins. For example, penicillin V is acid stable
and can be administered orally, but methicillin is acid
labile and must be given intravenously or intramus-
cularly. However, methicillin resists breakdown by
bacterial enzymes (H9252-lactamases) whereas many other
penicillins do not. The H9252-lactams have many of the
properties that make a good drug. First, they target
a metabolic pathway present in bacteria but not in
people. Second, they have half-lives in the body long
enough to be clinically useful. Third, they reach thera-
CCN
H
C
H
O
C
O
R
S CH
3
CH
3
CH
COOH
C
N
CH
3
CH
3
H
2
O
C
H H
Enz Ser OH
CN
H
C
C
O
R
S
CH
COOH
C
N
H
O
Trans-
peptidase
b-Lactamase
Ser O
CH
3
CH
3
C
HH
CN
H
C
C
O
R
S
CH
COOH
C
N
H
O
H11002
O
(a) (b)
Stably derivatized,
inactive transpeptidase
Inactive penicillin
H
CH
3
CH
3
C
H H
CN
H
C
C
O
R
S
CH
COOH
C
N
H
O
Ser O
Penicillin
b-Lactamase
FIGURE 2
H
CN
H
C
O
C
O
R
R groups
S
CH
3
CH
2
CH
3
COOH
CH
C
N
Side chain Thiazolidine ring
General structure of penicillins
O CH
2
OCH
3
OCH
3
Methicillin
Penicillin V
Penicillin G
(Benzylpenicillin)
1
5
6
7
4
3
2
b-Lactam
ring
H
C#
!
!
FIGURE 1
peutic concentrations in most, if not all, tissues and
organs. Finally, they are effective against a broad
range of bacterial species.
Penicillins block formation of the peptide cross-
links in peptidoglycans, acting as mechanism-based
(suicide) inhibitors. The normal catalytic mechanism
of the target enzyme activates the inhibitor, which then
covalently modifies a critical residue in the active site.
Transpeptidases employ a reaction mechanism (involv-
ing Ser residues) similar to that of chymotrypsin (see
Fig. 6–21); the reaction activates H9252-lactams such as
penicillin, which in turn inactivate the transpeptidases.
After penicillin enters the transpeptidase active site, the
proton on the hydroxyl group of an active-site Ser res-
idue is abstracted to the nitrogen of the H9252-lactam ring,
and the activated oxygen of the Ser hydroxyl attacks
the carbonyl carbon at position 7 of the H9252-lactam, open-
ing the ring and forming a stable penicilloyl-enzyme
derivative that inactivates the enzyme (Fig. 2a).
Widespread use of antibiotics has driven the se-
lection and evolution of antibiotic resistance in many
pathogenic bacteria. The most important mechanism of
resistance is inactivation of the antibiotic by enzymatic
hydrolysis of the lactam ring, catalyzed by bacterial
(continued on next page)
8885d_c20_751–786 2/18/04 1:56 PM Page 779 mac76 mac76:385_reb:
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria780
SUMMARY 20.4 Synthesis of Cell Wall
Polysaccharides: Plant Cellulose and
Bacterial Peptidoglycan
■ Cellulose synthesis takes place in terminal
complexes (rosettes) in the plasma membrane.
Each cellulose chain begins as a sitosterol
dextrin formed inside the cell. It then flips to
the outside, where the oligosaccharide portion
is transferred to cellulose synthase in the
rosette and is then extended. Each rosette
produces 36 separate cellulose chains
simultaneously and in parallel. The chains
crystallize into one of the microfibrils that form
the cell wall.
■ Synthesis of the bacterial cell wall
peptidoglycan also involves lipid-linked
oligosaccharides formed inside the cell and
flipped to the outside for assembly.
20.5 Integration of Carbohydrate
Metabolism in the Plant Cell
Carbohydrate metabolism in a typical plant cell is more
complex in several ways than that in a typical animal
cell. The plant cell carries out the same processes that
generate energy in animal cells (glycolysis, citric acid
cycle, and oxidative phosphorylation); it can generate
hexoses from three- or four-carbon compounds by glu-
coneogenesis; it can oxidize hexose phosphates to pen-
tose phosphates with the generation of NADPH (the ox-
idative pentose phosphate pathway); and it can produce
a polymer of (H92511n4)-linked glucose (starch) and de-
grade it to generate hexoses. But besides these carbo-
hydrate transformations that it shares with animal cells,
the photosynthetic plant cell can fix CO
2
into organic
compounds (the rubisco reaction); use the products of
fixation to generate trioses, hexoses, and pentoses (the
Calvin cycle); and convert acetyl-CoA generated from
fatty acid breakdown to four-carbon compounds (the
glyoxylate cycle) and the four-carbon compounds to
hexoses (gluconeogenesis). These processes, unique to
the plant cell, are segregated in several compartments
not found in animal cells: the glyoxylate cycle in gly-
oxysomes, the Calvin cycle in chloroplasts, starch syn-
thesis in amyloplasts, and organic acid storage in vac-
uoles. The integration of events among these various
compartments requires specific transporters in the
membranes of each organelle, to move products from
one organelle to another or into the cytosol.
Gluconeogenesis Converts Fats and Proteins
to Glucose in Germinating Seeds
Many plants store lipids and proteins in their seeds, to
be used as sources of energy and as biosynthetic pre-
cursors during germination, before photosynthetic
mechanisms have developed. Active gluconeogenesis in
germinating seeds provides glucose for the synthesis of
sucrose, polysaccharides, and many metabolites derived
from hexoses. In plant seedlings, sucrose provides much
of the chemical energy needed for initial growth.
We noted earlier (Chapter 14) that animal cells
can carry out gluconeogenesis from three- and four-
carbon precursors, but not from the two acetyl carbons
BOX 20–1 BIOCHEMISTRY IN MEDICINE (continued from previous page)
H9252-lactamases, which provide bacteria with a bullet-
proof vest (Fig. 2b). A H9252-lactamase forms a temporary
covalent adduct with the carboxyl group of the opened
H9252-lactam ring, which is immediately hydrolyzed, regen-
erating active enzyme. One approach to circumvent-
ing antibiotic resistance of this type is to synthesize
penicillin analogs, such as methicillin, that are poor
substrates for H9252-lactamases. Another approach is to
administer along with antibiotics a H9252-lactamase in-
hibitor such as clavulanate or sulbactam.
Antibiotic resistance is a significant threat to pub-
lic health. Some bacterial infections are now essen-
tially untreatable with antibiotics. By the early 1990s,
20% to 40% of Staphylococcus aureus (the causative
agent of “staph” infections) was resistant to methicillin,
and 32% of Neisseria gonorrhoeae (the causative
agent of gonorrhea) was resistant to penicillin. By
1986, 32% of Shigella (a pathogen responsible for se-
vere forms of dysentery, some with a lethality of up to
15%) was resistant to ampicillin. Significantly, many
of these pathogens are also resistant to many other
antibiotics. In the future, we will need to develop new
drugs that circumvent the bacterial resistance mech-
anisms or that act on different bacterial targets.
8885d_c20_751–786 2/18/04 1:56 PM Page 780 mac76 mac76:385_reb:
20.5 Integration of Carbohydrate Metabolism in the Plant Cell 781
of acetyl-CoA. Because the pyruvate dehydrogenase
reaction is effectively irreversible (pp. 602–603), animal
cells have no way to convert acetyl-CoA to pyruvate or
oxaloacetate. Unlike animals, plants and some micro-
organisms can convert acetyl-CoA derived from fatty
acid oxidation to glucose. Some of the enzymes essen-
tial to this conversion are sequestered in glyoxysomes,
where glyoxysome-specific isozymes of H9252-oxidation
break down fatty acids to acetyl-CoA (see Fig. 16–22).
The physical separation of the glyoxylate cycle and
H9252-oxidation enzymes from the mitochondrial citric acid
cycle enzymes prevents further oxidation of acetyl-CoA
to CO
2
. Instead, the acetyl-CoA is converted to succi-
nate in the glyoxylate cycle (see Fig. 16–20). The suc-
cinate passes into the mitochondrial matrix, where it is
converted by citric acid cycle enzymes to oxaloacetate,
which moves into the cytosol. Cytosolic oxaloacetate
is converted by gluconeogenesis to fructose 6-phos-
phate, the precursor of sucrose. Thus the integration of
reaction sequences in three subcellular compartments
is required for the production of fructose 6-phosphate
or sucrose from stored lipids. Because only three of
the four carbons in each molecule of oxaloacetate are
converted to hexose in the cytosol, about 75% of the
carbon in the fatty acids stored as seed lipids is con-
verted to carbohydrate by the combined pathways of
Figure 20–35. The other 25% is lost as CO
2
in the
conversion of oxaloacetate to phosphoenolpyruvate.
Hydrolysis of storage triacylglycerols also produces
glycerol 3-phosphate, which can enter the gluconeo-
genic pathway after its oxidation to dihydroxyacetone
phosphate (Fig. 20–36).
Glucogenic amino acids (see Table 14–4) derived
from the breakdown of stored seed proteins also yield
precursors for gluconeogenesis, following transamina-
tion and oxidation to succinyl-CoA, pyruvate, oxaloac-
etate, fumarate, and H9251-ketoglutarate (Chapter 18)—all
good starting materials for gluconeogenesis.
Pools of Common Intermediates Link Pathways
in Different Organelles
Although we have described metabolic transformations
in plant cells in terms of individual pathways, these
pathways interconnect so completely that we should in-
stead consider pools of metabolic intermediates shared
among these pathways and connected by readily re-
versible reactions (Fig. 20–37). One such metabolite
pool includes the hexose phosphates glucose 1-phos-
phate, glucose 6-phosphate, and fructose 6-phosphate;
a second includes the 5-phosphates of the pentoses ri-
bose, ribulose, and xylulose; a third includes the triose
phosphates dihydroxyacetone phosphate and glycer-
aldehyde 3-phosphate. Metabolite fluxes through these
Glucose 6-phosphate
Fructose 6-phosphate
Sucrose
Oxaloacetate
Malate
Succinate
Acetyl-CoA
Fatty acid
oxidation
Phosphoenolpyruvate
Glyoxysome
Mitochondrion
Cytosol
gluconeogenesis
H11002
OOC CH
2
CH
2
COO
H11002
CH
3
C
O
S-CoA
Acetyl-CoA
a
b
Citric
acid
cycle
Glyoxylate
cycle
Succinate
Fumarate
Isocitrate
Oxaloacetate
Citrate
Malate
Succinyl-CoA
Oxaloacetate
CH
2
CO
2
C
O
H11002
OOC COO
H11002
-Ketoglutarate
Citrate
Isocitrate
Glyoxylate
CH
3
C
O
S-CoA
FIGURE 20–35 Conversion of stored fatty acids to sucrose in germi-
nating seeds. This pathway begins in glyoxysomes. Succinate is pro-
duced and exported to mitochondria, where it is converted to oxalo-
acetate by enzymes of the citric acid cycle. Oxaloacetate enters the
cytosol and serves as the starting material for gluconeogenesis and for
the synthesis of sucrose, the transport form of carbon in plants.
8885d_c20_751–786 2/18/04 1:56 PM Page 781 mac76 mac76:385_reb:
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria782
CH
2
O
C S-CoA
CH
2
CH
3
O
C
O C
O
CH O C
O
CH
2
Triacylglycerol
Fatty acids
Glycerol
Glycerol 3-phosphate
NADH H11001 H
H11001
NAD
H11001
Acetyl-CoA
Dihydroxyacetone
phosphate
P
CH
2
OH
H9252 oxidation
lipase
glycerol
kinase
O C
O
CHOH
CH
2
OH
CHOH
CH
2
OH
ATP
ADP
CH
2
OH
CH
2
OP
O
glycerol
3-phosphate
dehydrogenase
FIGURE 20–36 Conversion of the glycerol moiety of triacylglycerols
to sucrose in germinating seeds. The glycerol of triacylglycerols is ox-
idized to dihydroxyacetone phosphate, which enters the gluconeo-
genic pathway at the triose phosphate isomerase reaction.
Starch
ADP-Glucose UDP-Glucose Sucrose
Glucose 1-phosphate Glucose 6-phosphate Frucose 6-phosphate
6-Phosphogluconate Fructose 1,6-bisphosphate
Ribose 5-phosphate
Glyceraldehyde 3-phosphate Dihydroxyacetone
phosphate
Ribulose 5-phosphate
Xylulose 5-phosphate
FIGURE 20–37 Pools of pentose
phosphates, triose phosphates, and
hexose phosphates. The compounds
in each pool are readily intercon-
vertible by reactions that have small
standard free-energy changes.
When one component of the pool
is temporarily depleted, a new
equilibrium is quickly established to
replenish it. Movement of the sugar
phosphates between intracellular
compartments is limited; specific
transporters must be present in an
organelle membrane.
pools change in magnitude and direction in response to
changes in the circumstances of the plant, and they vary
with tissue type. Transporters in the membranes of each
organelle move specific compounds in and out, and the
regulation of these transporters presumably influences
the degree to which the pools mix.
During daylight hours, triose phosphates produced
in leaf tissue by the Calvin cycle move out of the chloro-
plast and into the cytosolic hexose phosphate pool,
where they are converted to sucrose for transport to
nonphotosynthetic tissues. In these tissues, sucrose is
converted to starch for storage or is used as an energy
source via glycolysis. In growing plants, hexose phos-
phates are also withdrawn from the pool for the syn-
thesis of cell walls. At night, starch is metabolized by
glycolysis to provide energy, essentially as in nonphoto-
synthetic organisms, and NADPH and ribose 5-phos-
phate are obtained through the oxidative pentose phos-
phate pathway.
SUMMARY 20.5 Integration of Carbohydrate
Metabolism in the Plant Cell
■ Plants can synthesize sugars from acetyl-CoA,
the product of fatty acid breakdown, by the
combined actions of the glyoxylate cycle and
gluconeogenesis.
■ The individual pathways of carbohydrate
metabolism in plants overlap extensively; they
share pools of common intermediates, including
hexose phosphates, pentose phosphates, and
triose phosphates. Transporters in the
membranes of chloroplasts, mitochondria,
amyloplasts, and peroxisomes mediate the
movement of sugar phosphates between
organelles. The direction of metabolite flow
through the pools changes from day to night.
8885d_c20_751–786 2/18/04 1:56 PM Page 782 mac76 mac76:385_reb:
Chapter 20 Further Reading 783
Key Terms
Calvin cycle 752
plastids 752
chloroplast 752
amyloplast 752
carbon-fixation reaction 753
ribulose 1,5-bisphosphate 753
3-phosphoglycerate 753
pentose phosphate pathway 753
reductive pentose phosphate
pathway 753
C
3
plants 754
ribulose 1,5-bisphosphate
carboxylase/oxygenase
(rubisco) 754
rubisco activase 757
transaldolase 759
transketolase 759
sedoheptulose 1,7-bisphosphate 760
ribulose 5-phosphate 760
carbon-assimilation
reactions 763
thioredoxin 764
ferredoxin-thioredoxin
reductase 764
photorespiration 766
2-phosphoglycolate 766
glycolate pathway 767
oxidative photosynthetic carbon cycle
(C
2
cycle) 769
C
4
plants 769
phosphoenolpyruvate
carboxylase 769
malic enzyme 769
pyruvate phosphate dikinase 769
CAM plants 770
nucleotide sugars 771
ADP-glucose 771
starch synthase 771
sucrose 6-phosphate synthase 772
fructose 2,6-bisphosphate 773
ADP-glucose pyrophosphorylase 774
cellulose synthase 775
peptidoglycan 777
metabolite pools 781
Terms in bold are defined in the glossary.
Further Reading
General References
Blankenship, R.E. (2002) Molecular Mechanisms of
Photosynthesis, Blackwell Science, Oxford.
Very readable, well-illustrated, intermediate-level treatment of
all aspects of photosynthesis, including the carbon metabolism
covered in this chapter and the light-driven reactions described
in Chapter 19.
Buchanan, B.B., Gruissem, W., & Jones, R.L. (eds) (2000)
Biochemistry and Molecular Biology of Plants, American Society
of Plant Physiology, Rockville, MD.
This wonderful book—up to date and authoritative—covers all
aspects of plant biochemistry and molecular biology. The fol-
lowing chapters cover carbohydrate synthesis in greater depth:
Malkin, R. & Niyogi, K., Chapter 12, Photosynthesis (pp.
568–629); Dennis, D.T. & Blakeley, S.D., Chapter 13, Carbo-
hydrate Metabolism (pp. 630–675); Siedow, J.N. & Day, D.A.,
Chapter 14, Respiration and Photorespiration (pp. 676–729).
Heldt, H.-W. (1997) Plant Biochemistry and Molecular Biology,
Oxford University Press, Oxford.
Another excellent textbook of plant biochemistry. Especially
useful are Chapter 6, Photosynthetic CO
2
Assimilation by the
Calvin Cycle; Chapter 7, Photorespiration; and Chapter 9,
Polysaccharides.
Photosynthetic Carbohydrate Synthesis
Andersson, I., Knight, S., Schneider, G., Lindqvist, Y.,
Lundqvist, T., Br?ndén, C.-I., & Lorimer, G.H. (1989) Crystal
structure of the active site of ribulose-bisphosphate carboxylase.
Nature 337, 229–234.
Benson, A.A. (2002) Following the path of carbon in photosyn-
thesis: a personal story—history of photosynthesis. Photosynth.
Res. 73, 31–49.
Cleland, W.W., Andrews, T.J., Gutteridge, S., Hartman, F.C.,
& Lorimer, G.H. (1998) Mechanism of rubisco—the carbamate as
general base. Chem. Rev. 98, 549–561.
Review with a special focus on the carbamate at the active site.
Dietz, K.J., Link, G., Pistorius, E.K., & Scheibe, R. (2002)
Redox regulation in oxygenic photosynthesis. Prog. Botany 63,
207–245.
Flügge, U.-I. (1999) Phosphate translocaters in plastids. Annu.
Rev. Plant Physiol. Plant Mol. Biol. 50, 27–45.
Review of the transporters that carry P
i
and various sugar
phosphates across plastid membranes.
Fridlyand, L.E. & Scheibe R. (1999) Homeostatic regulation
upon changes of enzyme activities in the Calvin cycle as an
example for general mechanisms of flux control: what can we
expect from transgenic plants? Photosynth. Res. 61, 227–239.
Hartman, F.C. & Harpel, M.R. (1994) Structure, function,
regulation and assembly of D-ribulose-1,5-bisphosphate
carboxylase/oxygenase. Annu. Rev. Biochem. 63, 197–234.
Horecker, B.L. (2002) The pentose phosphate pathway. J. Biol.
Chem. 277, 47,965–47,971.
Portis, A.R., Jr. (2003) Rubisco activase: rubisco’s catalytic
chaperon. Photosynth. Res. 75, 11–27.
Structure, regulation, mechanism, and importance of rubisco
activase.
Raven, J. A. & Girard-Bascou, J. (2001) Algal model systems
and the elucidation of photosynthetic metabolism. J. Phycol. 37,
943–950.
Short, intermediate-level review.
Schneider, G., Lindqvist, Y., Br?ndén, C.-I., & Lorimer, G.
(1986) Three-dimensional structure of ribulose-1,5-bisphosphate
carboxylase/oxygenase from Rhodospirillum rubrum at 2.9 ?
resolution. EMBO J. 5, 3409–3415.
Smith, A.M., Denyer, K., & Martin, C. (1997) The synthesis of
the starch granule. Annu. Rev. Plant Physiol. Plant Mol. Biol.
48, 67–87.
Review of the role of ADP-glucose pyrophosphorylase in the
synthesis of amylose and amylopectin in starch granules.
8885d_c20_783 2/20/04 12:03 PM Page 783 mac76 mac76:385_reb:
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria784
Spreitzer, R.J. & Salvucci, M.E. (2002). Rubisco: structure,
regulatory interactions, and possibilities for a better enzyme.
Annu. Rev. Plant Biol. 53, 449–475.
Advanced review on rubisco and rubisco activase.
Stitt, M. (1995) Regulation of metabolism in transgenic plants.
Annu. Rev. Plant Physiol. Plant Mol. Biol. 46, 341–368.
Review of the genetic approaches to defining points of
regulation in vivo.
Tabita, F.R. (1999) Microbial ribulose 1,5-bisphosphate carboxy-
lase/oxygenase: a different perspective. Photosynth. Res. 60, 1–28.
Discussion of biochemical and genetic studies of the microbial
rubisco, and comparison with the enzyme from plants.
Wolosiuk, R., Ballicora, M., & Hagelin, K. (1993) The reduc-
tive pentose phosphate cycle for photosynthetic CO
2
assimilation:
enzyme modulation. FASEB J. 7, 622–637.
Photorespiration and the C
4
and CAM Pathways
Douce, R., Bourguignon, J., Neuburger, M., & Rébeillé, F.
(2001) The glycine decarboxylase system: a fascinating complex.
Trends Plant Sci. 6, 167–176.
Intermediate-level description of the structure and the reaction
mechanism of the enzyme.
Douce, R. & Neuberger, M. (1999) Biochemical dissection of
photorespiration. Curr. Opin. Plant Biol. 2, 214–222.
Hatch, M.C. (1987) C
4
photosynthesis: a unique blend of modified
biochemistry, anatomy and ultrastructure. Biochim. Biophys. Acta
895, 81–106.
Intermediate-level review by one of the discoverers of the C
4
pathway.
Tolbert, N.E. (1997) The C
2
oxidative photosynthetic carbon
cycle. Annu. Rev. Plant Physiol. Plant Mol. Biol. 48, 1–25.
A fascinating personal account of the development of our
understanding of photorespiration, by one who was central in
this development.
Biosynthesis of Starch and Sucrose
Ball, S.G., van de Wal, M.H.B.J., & Visser, R.G.F. (1998)
Progress in understanding the biosynthesis of amylose. Trends
Plant Sci. 3, 462–467.
Delmer, D.P. (1999) Cellulose biosynthesis: exciting times for a
difficult field of study. Annu. Rev. Plant Physiol. Plant Mol. Biol.
50, 245–276.
Doblin, M.S., Kurek, I., Javob-Wilk, D., & Delmer, D.P.
(2002) Cellulose biosynthesis in plants: from genes to rosettes.
Plant Cell Physiol. 43, 1407–1420.
Errington, J., Daniel, R.A., & Scheffers, D.J. (2003) Cyto-
kinesis in bacteria. Microbiol. Mol. Biol. Rev. 67, 52–65.
Fernie, A.R., Willmitzer, L., & Trethewey, R.N. (2002) Sucrose
to starch: a transition in molecular plant physiology. Trends Plant
Sci. 7, 35–41.
Intermediate-level review of the genes and proteins involved in
starch synthesis in plant tubers.
Huber, S.C. & Huber, J.L. (1996) Role and regulation of sucrose-
phosphate synthase in higher plants. Annu. Rev. Plant Physiol.
Plant Mol. Biol. 47, 431–444.
Short review of factors that regulate this critical enzyme.
Leloir, L.F. (1971) Two decades of research on the biosynthesis of
saccharides. Science 172, 1299–1303.
Leloir’s Nobel address, including a discussion of the role of
sugar nucleotides in metabolism.
Mukerjea, R., Yu, L., & Robyt, J.F. (2002) Starch biosynthesis:
mechanism for the elongation of starch chains. Carbohydrate Res.
337, 1015–1022.
Recent evidence that the starch chain grows at its reducing end.
Synthesis of Cellulose and Peptidoglycan
Delmer, D.P. (1999) Cellulose biosynthesis: exciting times for a
difficult field of study. Annu. Rev. Plant Physiol. Plant Mol. Biol.
50, 245–276.
Dhugga, K.S. (2001) Building the wall: genes and enzyme
complexes for polysaccharide synthases. Curr. Opin. Plant Biol.
4, 488–493.
Peng, L., Kawagoe, Y., Hogan, P., & Delmer, D. (2002)
Sitosterol-H9252-glucoside as primer for cellulose synthesis in plants.
Science 295, 147–150.
Recent evidence for the lipid-oligosaccharide intermediate in
cellulose synthesis.
Reid, J.S.G. (2000) Cementing the wall: cell wall polysaccharide
synthesising enzymes. Curr. Opin. Plant Biol. 3, 512–516.
Saxena, I.M. & Brown, R.M., Jr. (2000) Cellulose synthases and
related enzymes. Curr. Opin. Plant Biol. 3, 523–531.
Williamson, R.E., Burn, J.E., & Hocart, C.H. (2002) Towards
the mechanism of cellulose synthesis. Trends Plant Sci. 7, 461–467.
1. Segregation of Metabolism in Organelles What are
the advantages to the plant cell of having different organelles
to carry out different reaction sequences that share inter-
mediates?
2. Phases of Photosynthesis When a suspension of green
algae is illuminated in the absence of CO
2
and then incubated
with
14
CO
2
in the dark,
14
CO
2
is converted to [
14
C]glucose for
a brief time. What is the significance of this observation with
regard to the CO
2
-assimilation process, and how is it related
to the light reactions of photosynthesis? Why does the con-
version of
14
CO
2
to [
14
C]glucose stop after a brief time?
3. Identification of Key Intermediates in CO
2
Assim-
ilation Calvin and his colleagues used the unicellular green
alga Chlorella to study the carbon-assimilation reactions of
photosynthesis. They incubated
14
CO
2
with illuminated sus-
pensions of algae and followed the time course of appearance
Problems
8885d_c20_751–786 2/18/04 1:56 PM Page 784 mac76 mac76:385_reb:
Chapter 20 Problems 785
of
14
C in two compounds, X and Y, under two sets of condi-
tions. Suggest the identities of X and Y, based on your un-
derstanding of the Calvin cycle.
(a) Illuminated Chlorella were grown with unlabeled
CO
2
, then the light was turned off and
14
CO
2
was added (ver-
tical dashed line in the graph below). Under these conditions,
X was the first compound to become labeled with
14
C; Y was
unlabeled.
(b) Illuminated Chlorella cells were grown with
14
CO
2
.
Illumination was continued until all the
14
CO
2
had disap-
peared (vertical dashed line in the graph below). Under these
conditions, X became labeled quickly but lost its radioactiv-
ity with time, whereas Y became more radioactive with time.
4. Regulation of the Calvin Cycle Iodoacetate reacts
irreversibly with the free OSH groups of Cys residues in
proteins.
Predict which Calvin cycle enzyme(s) would be inhibited
by iodoacetate, and explain why.
5. Thioredoxin in Regulation of Calvin Cycle Enzymes
Motohashi and colleagues
*
used thioredoxin as a hook to fish
out from plant extracts the proteins that are activated by
thioredoxin. To do this, they prepared a mutant thioredoxin
in which one of the reactive Cys residues was replaced with
a Ser. Explain why this modification was necessary for their
experiments.
Radioactivity
0
Time
14
CO
2
Y
X
Radioactivity
CO
2
,
light
0
Time
14
CO
2
,
dark
Y
X
6. Comparison of the Reductive and Oxidative Pentose
Phosphate Pathways The reductive pentose phosphate
pathway generates a number of intermediates identical to
those of the oxidative pentose phosphate pathway (Chapter
14). What role does each pathway play in cells where it is
active?
7. Photorespiration and Mitochondrial Respiration
Compare the oxidative photosynthetic carbon cycle (C
2
cy-
cle), also called photorespiration, with the mitochondrial
respiration that drives ATP synthesis. Why are both
processes referred to as respiration? Where in the cell do they
occur, and under what circumstances? What is the path of
electron flow in each?
8. Rubisco and the Composition of the Atmosphere
N. E. Tolbert
?
has argued that the dual specificity of rubisco
for CO
2
and O
2
is not simply a leftover from evolution in a
low-oxygen environment. He suggests that the relative activ-
ities of the carboxylase and oxygenase activities of rubisco
actually have set, and now maintain, the ratio of CO
2
to O
2
in the earth’s atmosphere. Discuss the pros and cons of this
hypothesis, in molecular terms and in global terms. How does
the existence of C
4
organisms bear on the hypothesis?
9. Role of Sedoheptulose 1,7-Bisphosphatase What ef-
fect on the cell and the organism might result from a defect
in sedoheptulose 1,7-bisphosphatase in (a) a human hepato-
cyte and (b) the leaf cell of a green plant?
10. Pathway of CO
2
Assimilation in Maize If a maize
(corn) plant is illuminated in the presence of
14
CO
2
, after
about 1 second, more than 90% of all the radioactivity in-
corporated in the leaves is found at C-4 of malate, aspartate,
and oxaloacetate. Only after 60 seconds does
14
C appear at
C-1 of 3-phosphoglycerate. Explain.
11. Identifying CAM Plants Given some
14
CO
2
and all
the tools typically present in a biochemistry research lab, how
would you design a simple experiment to determine whether
a plant was a typical C
4
plant or a CAM plant?
12. Chemistry of Malic Enzyme: Variation on a Theme
Malic enzyme, found in the bundle-sheath cells of C
4
plants,
carries out a reaction that has a counterpart in the citric acid
cycle. What is the analogous reaction? Explain your choice.
13. The Cost of Storing Glucose as Starch Write the
sequence of steps and the net reaction required to calculate
the cost, in ATP molecules, of converting a molecule of cy-
tosolic glucose 6-phosphate to starch and back to glucose 6-
phosphate. What fraction of the maximum number of ATP
molecules available from complete catabolism of glucose 6-
phosphate to CO
2
and H
2
O does this cost represent?
CH
2
O
NAD
C
Inactive enzyme
O
ICH
2
O
CH11001
O
S
Iodoacetate
NAD
H11001H11001
H11002
H11002
SH
Cys Cys
HI
*Motohashi, K., Kondoh, A., Stumpp, M.T., & Hisabori, T. (2001) Com-
prehensive survey of proteins targeted by chloroplast thioredoxin.
Proc. Natl. Acad. Sci. USA 98, 11,224–11,229.
?
Tolbert, N.E. (1994) The role of photosynthesis and photorespiration
in regulating atmospheric CO
2
and O
2
. In Regulation of Atmospheric
CO
2
and O
2
by Photosynthetic Carbon Metabolism (Tolbert, N.E.
& Preiss, J., eds), pp. 8–33, Oxford University Press, New York.
8885d_c20_751–786 2/18/04 1:56 PM Page 785 mac76 mac76:385_reb:
Chapter 20 Carbohydrate Biosynthesis in Plants and Bacteria786
14. Inorganic Pyrophosphatase The enzyme inorganic
pyrophosphatase contributes to making many biosynthetic re-
actions that generate inorganic pyrophosphate essentially ir-
reversible in cells. By keeping the concentration of PP
i
very
low, the enzyme “pulls” these reactions in the direction of PP
i
formation. The synthesis of ADP-glucose in chloroplasts is
one reaction that is pulled in the forward direction by this
mechanism. However, the synthesis of UDP-glucose in the
plant cytosol, which produces PP
i
, is readily reversible in vivo.
How do you reconcile these two facts?
15. Regulation of Starch and Sucrose Synthesis Su-
crose synthesis occurs in the cytosol and starch synthesis in
the chloroplast stroma, yet the two processes are intricately
balanced. What factors shift the reactions in favor of (a)
starch synthesis and (b) sucrose synthesis?
16. Regulation of Sucrose Synthesis In the regulation
of sucrose synthesis from the triose phosphates produced
during photosynthesis, 3-phosphoglycerate and P
i
play criti-
cal roles (see Fig. 20–26). Explain why the concentrations of
these two regulators reflect the rate of photosynthesis.
17. Sucrose and Dental Caries The most prevalent in-
fection in humans worldwide is dental caries, which stems
from the colonization and destruction of tooth enamel by
a variety of acidifying microorganisms. These organisms
synthesize and live within a water-insoluble network of
dextrans, called dental plaque, composed of (H92511n6)-linked
polymers of glucose with many (H92511n3) branch points. Poly-
merization of dextran requires dietary sucrose, and the
reaction is catalyzed by a bacterial enzyme, dextran-sucrose
glucosyltransferase.
(a) Write the overall reaction for dextran polymerization.
(b) In addition to providing a substrate for the forma-
tion of dental plaque, how does dietary sucrose also provide
oral bacteria with an abundant source of metabolic energy?
18. Differences between C
3
and C
4
Plants The plant
genus Atriplex includes some C
3
and some C
4
species. From
the data in the plots below (species 1, black curve; species
2, red curve), identify which is a C
3
plant and which is a C
4
plant. Justify your answer in molecular terms that account
for the data in all three plots.
19. C
4
Pathway in a Single Cell In typical C
4
plants, the
initial capture of CO
2
occurs in one cell type, and the Calvin
cycle reactions occur in another (see Fig. 20–23). Voznesen-
skaya and colleagues
??
have described a plant, Bienertia cy-
cloptera—which grows in salty depressions of semidesert in
Central Asia—that shows the biochemical properties of a C
4
plant but unlike typical C
4
plants does not segregate the re-
actions of CO
2
fixation into two cell types. PEP carboxylase
and rubisco are present in the same cell. However, the cells
have two types of chloroplasts, which are localized differently,
as shown in the micrograph. One type, relatively poor in grana
(thylakoids), is confined to the periphery; the more typical
chloroplasts are clustered in the center of the cell, separated
from the peripheral chloroplasts by large vacuoles. Thin cy-
tosolic bridges pass through the vacuoles, connecting the pe-
ripheral and central cytosol.
10 H9262m
Uptake of CO
2
Uptake of CO
2
Uptake of CO
2
Light intensity Leaf temperature [CO
2
]
in intracellular space
??
Voznesenskaya, E.V., Fraceschi, V.R., Kiirats, O., Artyusheva, E.G.,
Freitag, H., & Edwards, G.E. (2002) Proof of C
4
photosynthesis with-
out Kranz anatomy in Bienertia cycloptera (Chenopodiaceae).
Plant J. 31, 649–662.
In this plant, where would you expect to find (a) PEP
carboxylase, (b) rubisco, and (c) starch granules? Explain
your answers with a model for CO
2
fixation in these C
4
cells.
8885d_c20_786 2/20/04 12:03 PM Page 786 mac76 mac76:385_reb:
chapter
L
ipids play a variety of cellular roles, some only re-
cently recognized. They are the principal form of
stored energy in most organisms and major constituents
of cellular membranes. Specialized lipids serve as pig-
ments (retinal, carotene), cofactors (vitamin K), deter-
gents (bile salts), transporters (dolichols), hormones
(vitamin D derivatives, sex hormones), extracellular and
intracellular messengers (eicosanoids, phosphatidylino-
sitol derivatives), and anchors for membrane proteins
(covalently attached fatty acids, prenyl groups, and
phosphatidylinositol). The ability to synthesize a vari-
ety of lipids is essential to all organisms. This chapter
describes the biosynthetic pathways for some of the
most common cellular lipids, illustrating the strategies
employed in assembling these water-insoluble products
from water-soluble precursors such as acetate. Like
other biosynthetic pathways, these reaction sequences
are endergonic and reductive. They use ATP as a source
of metabolic energy and a reduced electron carrier (usu-
ally NADPH) as reductant.
We first describe the biosynthesis of fatty acids, the
primary components of both triacylglycerols and phos-
pholipids, then examine the assembly of fatty acids into
triacylglycerols and the simpler membrane phospho-
lipids. Finally, we consider the synthesis of cholesterol,
a component of some membranes and the precursor of
steroids such as the bile acids, sex hormones, and
adrenocortical hormones.
21.1 Biosynthesis of Fatty Acids
and Eicosanoids
After the discovery that fatty acid oxidation takes place
by the oxidative removal of successive two-carbon
(acetyl-CoA) units (see Fig. 17–8), biochemists thought
the biosynthesis of fatty acids might proceed by a sim-
ple reversal of the same enzymatic steps. However, as
they were to find out, fatty acid biosynthesis and break-
down occur by different pathways, are catalyzed by dif-
ferent sets of enzymes, and take place in different parts
of the cell. Moreover, biosynthesis requires the partici-
pation of a three-carbon intermediate, malonyl-CoA,
that is not involved in fatty acid breakdown.
We focus first on the pathway of fatty acid synthe-
sis, then turn our attention to regulation of the pathway
and to the biosynthesis of longer-chain fatty acids, un-
saturated fatty acids, and their eicosanoid derivatives.
Malonyl-CoA Is Formed from Acetyl-CoA
and Bicarbonate
The formation of malonyl-CoA from acetyl-CoA is an
irreversible process, catalyzed by acetyl-CoA carbox-
ylase. The bacterial enzyme has three separate poly-
peptide subunits (Fig. 21–1); in animal cells, all three
CH
2
C
O
C
S-CoAO
Malonyl-CoA
O
H5008
LIPID BIOSYNTHESIS
21.1 Biosynthesis of Fatty Acids and Eicosanoids 787
21.2 Biosynthesis of Triacylglycerols 804
21.3 Biosynthesis of Membrane Phospholipids 808
21.4 Biosynthesis of Cholesterol, Steroids, and
Isoprenoids 816
How the division of “spoils” came about I do not recall—it
may have been by drawing lots. At any rate, David Shemin
“drew” amino acid metabolism, which led to his classical
work on heme biosynthesis. David Rittenburg was to
continue his interest in protein synthesis and turnover,
and lipids were to be my territory.
—Konrad Bloch, on how his career turned to problems of lipid
metabolism after the death of his mentor, Rudolf Schoen-
heimer; article in Annual Review of Biochemistry, 1987
21
787
8885d_c21_787-832 2/26/04 9:35 AM Page 787 mac76 mac76:385_reb:
activities are part of a single multifunctional polypep-
tide. Plant cells contain both types of acetyl-CoA
carboxylase. In all cases, the enzyme contains a biotin
prosthetic group covalently bound in amide linkage to
the H9255-amino group of a Lys residue in one of the three
polypeptides or domains of the enzyme molecule. The
two-step reaction catalyzed by this enzyme is very
similar to other biotin-dependent carboxylation reac-
tions, such as those catalyzed by pyruvate carboxylase
(see Fig. 16–16) and propionyl-CoA carboxylase (see
Fig. 17–11). The carboxyl group, derived from bicar-
bonate (HCO
3
H11002
), is first transferred to biotin in an ATP-
dependent reaction. The biotinyl group serves as a tem-
porary carrier of CO
2
, transferring it to acetyl-CoA in
the second step to yield malonyl-CoA.
Fatty Acid Synthesis Proceeds in a Repeating
Reaction Sequence
The long carbon chains of fatty acids are assembled in
a repeating four-step sequence (Fig. 21–2). A saturated
acyl group produced by this set of reactions becomes
the substrate for subsequent condensation with an ac-
tivated malonyl group. With each passage through the
cycle, the fatty acyl chain is extended by two carbons.
When the chain length reaches 16 carbons, the product
Chapter 21 Lipid Biosynthesis788
OC
HN
NH
S
Biotin
carboxylase
N
C
O
H5008
H11001HCO3
H5008
ATP ADP H11001 P
i
transcarboxylase
O
HN
NH
S
Biotin
arm
Lys
side arm
NH
C
O
C
O
H5008
O
biotin
carboxylase
Malonyl-CoA
CH
3
S-CoA
C
O
OC
HN
NH
S
NH
C
O
H11001
C
O
Acetyl-CoA
CH
3
S-CoA
C
OO
C
NH
CO
O
H5008
C
O
C
H5008
O
Malonyl-CoA
S-CoA
C
O
Acetyl-CoA
CH
3
S-CoA
C
OO
NH
C
O
C
H5008
O
CH
2
O
Lys
Biotin
carrier
protein
Biotin
carrier
protein
Transcarboxylase
Biotin
FIGURE 21–1 The acetyl-CoA carboxylase reaction. Acetyl-CoA
carboxylase has three functional regions: biotin carrier protein (gray);
biotin carboxylase, which activates CO
2
by attaching it to a nitrogen
in the biotin ring in an ATP-dependent reaction (see Fig. 16–16); and
transcarboxylase, which transfers activated CO
2
(shaded green) from
biotin to acetyl-CoA, producing malonyl-CoA. The long, flexible bi-
otin arm carries the activated CO
2
from the biotin carboxylase region
to the transcarboxylase active site, as shown in the diagrams below
the reaction arrows. The active enzyme in each step is shaded blue.
8885d_c21_787-832 2/26/04 9:35 AM Page 788 mac76 mac76:385_reb:
(palmitate, 16:0; see Table 10–1) leaves the cycle. Car-
bons C-16 and C-15 of the palmitate are derived from
the methyl and carboxyl carbon atoms, respectively, of
an acetyl-CoA used directly to prime the system at the
outset (Fig. 21–3); the rest of the carbon atoms in the
chain are derived from acetyl-CoA via malonyl-CoA.
Both the electron-carrying cofactor and the acti-
vating groups in the reductive anabolic sequence differ
from those in the oxidative catabolic process. Recall that
in H9252 oxidation, NAD
H11001
and FAD serve as electron ac-
ceptors and the activating group is the thiol (OSH)
group of coenzyme A (see Fig. 17–8). By contrast, the
reducing agent in the synthetic sequence is NADPH and
the activating groups are two different enzyme-bound
OSH groups, as described below.
All the reactions in the synthetic process are cat-
alyzed by a multienzyme complex, fatty acid synthase.
Although the details of enzyme structure differ in
prokaryotes such as Escherichia coli and in eukary-
otes, the four-step process of fatty acid synthesis is the
same in all organisms. We first describe the process as
it occurs in E. coli, then consider differences in enzyme
structure in other organisms.
The Fatty Acid Synthase Complex
Has Seven Different Active Sites
The core of the E. coli fatty acid synthase system con-
sists of seven separate polypeptides (Table 21–1), and
at least three others act at some stage of the process.
The proteins act together to catalyze the formation of
fatty acids from acetyl-CoA and malonyl-CoA. Through-
out the process, the intermediates remain covalently at-
tached as thioesters to one of two thiol groups of the
synthase complex. One point of attachment is the OSH
group of a Cys residue in one of the seven synthase pro-
teins (H9252-ketoacyl-ACP synthase); the other is the OSH
group of acyl carrier protein.
21.1 Biosynthesis of Fatty Acids and Eicosanoids 789
S
H5008
O
C
CH
2
C
O
S
O
H
Acetyl group
(first acyl group)
Fatty acid
synthase
dehydration
CH
2
CH
3
C
C
O
O
H
2
O
3
Saturated acyl group,
lengthened by two carbons
SCH
3
C
OO
CH
2
C
CO
2
1condensation
SCH
3
C
O
OH
CH
2
C
H
NADP
H11001
4reduction
NADPH H11001 H
H11001
SCH
3
C
O
CC
H
CH
3
CH
2
Malonyl group S
HS
HS
HS
NADP
H11001
2reduction
NADPH H11001 H
H11001
HS
b
a
FIGURE 21–2 Addition of two carbons to a growing fatty acyl chain:
a four-step sequence. Each malonyl group and acetyl (or longer acyl)
group is activated by a thioester that links it to fatty acid synthase, a
multienzyme complex described later in the text. 1 Condensation
of an activated acyl group (an acetyl group from acetyl-CoA is the first
acyl group) and two carbons derived from malonyl-CoA, with elimi-
nation of CO
2
from the malonyl group, extends the acyl chain by two
carbons. The mechanism of the first step of this reaction is given to il-
lustrate the role of decarboxylation in facilitating condensation. The
H9252-keto product of this condensation is then reduced in three more
steps nearly identical to the reactions of H9252 oxidation, but in the re-
verse sequence: 2 the H9252-keto group is reduced to an alcohol, 3
elimination of H
2
O creates a double bond, and 4 the double bond
is reduced to form the corresponding saturated fatty acyl group.
8885d_c21_787-832 2/26/04 9:35 AM Page 789 mac76 mac76:385_reb:
Acyl carrier protein (ACP) of E. coli is a small
protein (M
r
8,860) containing the prosthetic group
4H11541-phosphopantetheine (Fig. 21–4; compare this with
the panthothenic acid and H9252-mercaptoethylamine moi-
ety of coenzyme A in Fig. 8–41). Hydrolysis of thioesters
is highly exergonic, and the energy released helps
to make two different steps ( 1 and 5 in Fig. 21–5) in
fatty acid synthesis (condensation) thermodynamically
favorable. The 4H11032-phosphopante-theine prosthetic
group of ACP is believed to serve as a flexible arm,
tethering the growing fatty acyl chain to the surface
of the fatty acid synthase complex while carrying the
reaction intermediates from one enzyme active site to
the next.
Fatty Acid Synthase Receives the Acetyl
and Malonyl Groups
Before the condensation reactions that build up the fatty
acid chain can begin, the two thiol groups on the en-
zyme complex must be charged with the correct acyl
groups (Fig. 21–5, top). First, the acetyl group of acetyl-
CoA is transferred to the Cys OSH group of the H9252-
ketoacyl-ACP synthase. This reaction is catalyzed by
acetyl-CoA–ACP transacetylase (AT in Fig. 21–5).
The second reaction, transfer of the malonyl group from
malonyl-CoA to the OSH group of ACP, is catalyzed by
malonyl-CoA–ACP transferase (MT), also part of the
complex. In the charged synthase complex, the acetyl
Chapter 21 Lipid Biosynthesis790
C
H5008
O
CH
3
CH
2
C
O
O
CO
2
H11001
CH
3
4H
H11001
4e
H5008
CH
2
CH
2
O
S
CO
S
C
COO
H5008
O
CH
2
S
CO
2
H11001
CH
3
4H
H11001
4e
H5008
CH
2
CH
2
CH
2
CH
2
C
S
C
O
O
CH
2
S
CO
2
H11001
CH
3
4H
H11001
4e
H5008
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
C
S
C
O
O
CH
2
S
four more
additions
CH
3
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
CH
2
C
HS
H11001
Palmitate
Fatty acid
synthase
S
HS
COO
H5008
COO
H5008
COO
H5008
FIGURE 21–3 The overall process of palmitate synthesis. The fatty acyl chain grows
by two-carbon units donated by activated malonate, with loss of CO
2
at each step.
The initial acetyl group is shaded yellow, C-1 and C-2 of malonate are shaded pink,
and the carbon released as CO
2
is shaded green. After each two-carbon addition,
reductions convert the growing chain to a saturated fatty acid of four, then six, then
eight carbons, and so on. The final product is palmitate (16:0).
TABLE 21–1 Proteins of the Fatty Acid Synthase Complex of E. coli
Component Function
Acyl carrier protein (ACP) Carries acyl groups in thioester linkage
Acetyl-CoA–ACP transacetylase (AT) Transfers acyl group from CoA to Cys residue of KS
H9252-Ketoacyl-ACP synthase (KS) Condenses acyl and malonyl groups (KS has at least three isozymes)
Malonyl-CoA–ACP transferase (MT) Transfers malonyl group from CoA to ACP
H9252-Ketoacyl-ACP reductase (KR) Reduces H9252-keto group to H9252-hydroxyl group
H9252-Hydroxyacyl-ACP dehydratase (HD) Removes H
2
O from H9252-hydroxyacyl-ACP, creating double bond
Enoyl-ACP reductase (ER) Reduces double bond, forming saturated acyl-ACP
8885d_c21_790 2/27/04 12:40 PM Page 790 mac116 mac116:
and malonyl groups are very close to each other and are
activated for the chain-lengthening process. The first
four steps of this process are now considered in some
detail; all step numbers refer to Figure 21–5.
Step 1 Condensation The first reaction in the formation
of a fatty acid chain is condensation of the activated
acetyl and malonyl groups to form acetoacetyl-ACP,
an acetoacetyl group bound to ACP through the phos-
phopantetheine OSH group; simultaneously, a molecule
of CO
2
is produced. In this reaction, catalyzed by H9252-
ketoacyl-ACP synthase (KS), the acetyl group is
transferred from the Cys OSH group of the enzyme to
the malonyl group on the OSH of ACP, becoming the
methyl-terminal two-carbon unit of the new acetoacetyl
group.
The carbon atom of the CO
2
formed in this reaction
is the same carbon originally introduced into malonyl-
CoA from HCO
3
H11002
by the acetyl-CoA carboxylase reaction
(Fig. 21–1). Thus CO
2
is only transiently in covalent
linkage during fatty acid biosynthesis; it is removed as
each two-carbon unit is added.
Why do cells go to the trouble of adding CO
2
to make
a malonyl group from an acetyl group, only to lose the
CO
2
during the formation of acetoacetate? Recall that
in the H9252 oxidation of fatty acids (see Fig. 17–8), cleav-
age of the bond between two acyl groups (cleavage of
an acetyl unit from the acyl chain) is highly exergonic,
so the simple condensation of two acyl groups (two
acetyl-CoA molecules, for example) is highly ender-
gonic. The use of activated malonyl groups rather than
acetyl groups is what makes the condensation reactions
thermodynamically favorable. The methylene carbon
(C-2) of the malonyl group, sandwiched between car-
bonyl and carboxyl carbons, is chemically situated to act
as a good nucleophile. In the condensation step (step
1 ), decarboxylation of the malonyl group facilitates the
nucleophilic attack of the methylene carbon on the
thioester linking the acetyl group to H9252-ketoacyl-ACP
synthase, displacing the enzyme’s OSH group. Coupling
the condensation to the decarboxylation of the malonyl
group renders the overall process highly exergonic. A
similar carboxylation-decarboxylation sequence facili-
tates the formation of phosphoenolpyruvate from pyru-
vate in gluconeogenesis (see Fig. 14–17).
By using activated malonyl groups in the synthesis
of fatty acids and activated acetate in their degradation,
the cell makes both processes energetically favorable,
although one is effectively the reversal of the other. The
extra energy required to make fatty acid synthesis
favorable is provided by the ATP used to synthesize
malonyl-CoA from acetyl-CoA and HCO
3
H11002
(Fig. 21–1).
Step 2 Reduction of the Carbonyl Group The acetoacetyl-
ACP formed in the condensation step now undergoes
reduction of the carbonyl group at C-3 to form D-H9252-
hydroxybutyryl-ACP. This reaction is catalyzed by H9252-
ketoacyl-ACP reductase (KR) and the electron donor
is NADPH. Notice that the D-H9252-hydroxybutyryl group
does not have the same stereoisomeric form as the L-H9252-
hydroxyacyl intermediate in fatty acid oxidation (see
Fig. 17–8).
Step 3 Dehydration The elements of water are now re-
moved from C-2 and C-3 of D-H9252-hydroxybutyryl-ACP to
yield a double bond in the product, trans-H9004
2
- butenoyl-
ACP. The enzyme that catalyzes this dehydration is H9252-
hydroxyacyl-ACP dehydratase (HD).
Step 4 Reduction of the Double Bond Finally, the double
bond of trans-H9004
2
-butenoyl-ACP is reduced (saturated)
to form butyryl-ACP by the action of enoyl-ACP re-
ductase (ER); again, NADPH is the electron donor.
The Fatty Acid Synthase Reactions Are Repeated
to Form Palmitate
Production of the four-carbon, saturated fatty acyl–ACP
completes one pass through the fatty acid synthase
21.1 Biosynthesis of Fatty Acids and Eicosanoids 791
FIGURE 21–4 Acyl carrier protein (ACP). The prosthetic group is 4H11032-
phosphopantetheine, which is covalently attached to the hydroxyl
group of a Ser residue in ACP. Phosphopantetheine contains the B vi-
tamin pantothenic acid, also found in the coenzyme A molecule. Its
OSH group is the site of entry of malonyl groups during fatty acid
synthesis.
H5008
O
CH
2
C
O
CH
3
SH
OP
HN
O
CH
2
OC
CH
3
CHOH
CH
2
C O
HN
CH
2
CH
2
ACP
Malonyl groups
are esterified to
the
4H11032-Phospho-
pantetheine
Pantothenic
acid
Ser
side chain
SH group.
CH
2
8885d_c21_791 2/27/04 12:40 PM Page 791 mac116 mac116:
Chapter 21 Lipid Biosynthesis792
KS
AT
MT
KR
HDER
ACP
HS
S
CoA-SH
Malonyl-CoA
KS
AT
MT
KR
HDER
ACP
S
S
KS
AT
MT
KR
HDER
ACP
S
HS
CO
2
condensation
KS
AT
MT
KR
HDER
ACP
HS
HS
CoA-SH
Acetyl-CoA
1
Fatty acid synthase
complex charged
with an acetyl and
a malonyl group
b-Ketobutyryl-ACP
KS
AT
MT
KR
HDER
ACP
S
b-Ketoacyl-ACP
synthase
Malonyl-CoA–
ACP transferase
Enoyl-ACP
reductase
b-Hydroxyacyl-ACP
dehydratase
Acetyl-CoA–ACP
transacetylase
b-Ketoacyl-ACP
reductase
KS
AT
MT
KR
HDER
HS
KS
AT
MT
KR
HDER
ACP
S
HS
reduction of
double bond
trans-D
2
-Butenoyl-ACP
NADP
H11001
NADPH H11001 H
H11001
Butyryl-ACP
KS
AT
MT
KR
HDER
ACP
SH
S
translocation of
butyryl group to
Cys on b-ketoacyl-ACP
synthase (KS)
KS
AT
MT
KR
HDER
ACP
S
HS
b-Hydroxybutyryl-ACP
dehydration
H
2
O
reduction of
b-keto group
NADP
H11001
NADPH H11001 H
H11001
CH
3
C
O
S-CoA
CH
3
C
O
CH
2
C
O
S-CoA
C
O
H11002
O
CH
2
C
O
CH
3
CH
2
CH
2
C
O
C
O
H11002
O
CH
3
O
C
CH
2
C
O
CH
3
O
C
CH
2
C
O
CHCH
3
OH
CH C
O
CHCH
3
C
O
CH
3
CH
2
CH
2
2 3
4
5
FIGURE 21–5 Sequence of events during synthesis of a
fatty acid. The fatty acid synthase complex is shown
schematically. Each segment of the disk represents one of
the six enzymatic activities of the complex. At the center is
acyl carrier protein (ACP), with its phosphopantetheine arm
ending in an OSH. The enzyme shown in blue is the one
that will act in the next step. As in Figure 21–3, the initial
acetyl group is shaded yellow, C-1 and C-2 of malonate
are shaded pink, and the carbon released as CO
2
is
shaded green. Steps 1 to 4 are described in the text.
8885d_c21_787-832 2/26/04 9:35 AM Page 792 mac76 mac76:385_reb:
complex. The butyryl group is now transferred from the
phosphopantetheine OSH group of ACP to the Cys OSH
group of H9252-ketoacyl-ACP synthase, which initially bore
the acetyl group (Fig. 21–5). To start the next cycle of
four reactions that lengthens the chain by two more car-
bons, another malonyl group is linked to the now unoc-
cupied phosphopantetheine OSH group of ACP (Fig.
21–6). Condensation occurs as the butyryl group, act-
ing like the acetyl group in the first cycle, is linked to
two carbons of the malonyl-ACP group with concurrent
loss of CO
2
. The product of this condensation is a six-
carbon acyl group, covalently bound to the phospho-
pantetheine OSH group. Its H9252-keto group is reduced in
the next three steps of the synthase cycle to yield the
saturated acyl group, exactly as in the first round of re-
actions—in this case forming the six-carbon product.
Seven cycles of condensation and reduction pro-
duce the 16-carbon saturated palmitoyl group, still
bound to ACP. For reasons not well understood, chain
elongation by the synthase complex generally stops at
this point and free palmitate is released from the ACP
by a hydrolytic activity in the complex. Small amounts
of longer fatty acids such as stearate (18:0) are also
formed. In certain plants (coconut and palm, for exam-
ple) chain termination occurs earlier; up to 90% of the
fatty acids in the oils of these plants are between 8 and
14 carbons long.
We can consider the overall reaction for the syn-
thesis of palmitate from acetyl-CoA in two parts. First,
the formation of seven malonyl-CoA molecules:
7 Acetyl-CoA H11001 7CO
2
H11001 7ATP n
7 malonyl-CoA H11001 7ADP H11001 7P
i
(21–1)
then seven cycles of condensation and reduction:
Acetyl-CoA H11001 7 malonyl-CoA H11001 14NADPH H11001 14H
H11001
n
palmitate H11001 7CO
2
H11001 8 CoA H11001 14NADP
H11001
H11001 6H
2
O (21–2)
The overall process (the sum of Eqns 21–1 and 21–2)
is
8 Acetyl-CoA H11001 7ATP H11001 14NADPH H11001 14H
H11001
n
palmitate H11001 8 CoA H11001 7ADP H11001 7P
i
H11001 14NADP
H11001
H11001 6H
2
O
(21–3)
The biosynthesis of fatty acids such as palmitate thus
requires acetyl-CoA and the input of chemical energy in
two forms: the group transfer potential of ATP and the
reducing power of NADPH. The ATP is required to at-
tach CO
2
to acetyl-CoA to make malonyl-CoA; the
NADPH is required to reduce the double bonds. We re-
turn to the sources of acetyl-CoA and NADPH soon, but
first let’s consider the structure of the remarkable
enzyme complex that catalyzes the synthesis of fatty
acids.
21.1 Biosynthesis of Fatty Acids and Eicosanoids 793
CH
2
C
O
S-CoA
C
O
H5008
O
CH
2
C
O
CH
3
CH
2
KS
AT
MT
KR
HDER
ACP
SH
S
Butyryl group
CoA-SH
Malonyl-CoA
KS
AT
MT
KR
HDER
ACP
S
S
KS
AT
MT
KR
HDER
ACP
S
SH
CO
2
condensation
-Ketoacyl-ACP
CH
2
C
O
C
O
H5008
O
CH
2
C
O
CH
2
CH
3
CH
2
C
O
CH
2
C
O
CH
2
CH
3
b
FIGURE 21–6 Beginning of the second round of the fatty acid syn-
thesis cycle. The butyryl group is on the Cys OSH group. The incoming
malonyl group is first attached to the phosphopantetheine OSH group.
Then, in the condensation step, the entire butyryl group on the Cys
OSH is exchanged for the carboxyl group of the malonyl residue,
which is lost as CO
2
(green). This step is analogous to step 1 in Fig-
ure 21–5. The product, a six-carbon H9252-ketoacyl group, now contains
four carbons derived from malonyl-CoA and two derived from the
acetyl-CoA that started the reaction. The H9252-ketoacyl group then un-
dergoes steps 2 through 4 , as in Figure 21–5.
8885d_c21_787-832 2/26/04 9:35 AM Page 793 mac76 mac76:385_reb:
The Fatty Acid Synthase of Some Organisms Consists
of Multifunctional Proteins
In E. coli and some plants, the seven active sites for
fatty acid synthesis (six enzymes and ACP) reside in
seven separate polypeptides (Fig. 21–7, top). In these
complexes, each enzyme is positioned with its active site
near that of the preceding and succeeding enzymes of
the sequence. The flexible pantetheine arm of ACP can
reach all the active sites, and it carries the growing fatty
acyl chain from one site to the next; intermediates are
not released from the enzyme complex until it has
formed the finished product. As we have seen in earlier
chapters, this channeling of intermediates from one ac-
tive site to the next increases the efficiency of the over-
all process.
The fatty acid synthases of yeast and of vertebrates
are also multienzyme complexes, and their integration
is even more complete than in E. coli and plants. In
yeast, the seven distinct active sites reside in two large,
multifunctional polypeptides, with three activities on
the H9251 subunit and four on the H9252 subunit. In vertebrates,
a single large polypeptide (M
r
240,000) contains all
seven enzymatic activities as well as a hydrolytic activ-
ity that cleaves the finished fatty acid from the ACP-like
part of the enzyme complex. The vertebrate enzyme
functions as a dimer (M
r
480,000) in which the two iden-
tical subunits lie head-to-tail. The subunits appear to
function independently. When all the active sites in one
subunit are inactivated by mutation, palmitate synthe-
sis is only modestly reduced.
Fatty Acid Synthesis Occurs in the Cytosol of Many
Organisms but in the Chloroplasts of Plants
In most higher eukaryotes, the fatty acid synthase com-
plex is found exclusively in the cytosol (Fig. 21–8), as
are the biosynthetic enzymes for nucleotides, amino
acids, and glucose. This location segregates synthetic
processes from degradative reactions, many of which
take place in the mitochondrial matrix. There is a cor-
responding segregation of the electron-carrying cofac-
tors used in anabolism (generally a reductive process)
and those used in catabolism (generally oxidative).
Usually, NADPH is the electron carrier for anabolic
reactions, and NAD
H11001
serves in catabolic reactions. In
hepatocytes, the [NADPH]/[NADP
H11001
] ratio is very high
(about 75) in the cytosol, furnishing a strongly reducing
environment for the reductive synthesis of fatty acids
and other biomolecules. The cytosolic [NADH]/[NAD
H11001
]
ratio is much smaller (only about 8 H11003 10
H110024
), so the
NAD
H11001
-dependent oxidative catabolism of glucose can
take place in the same compartment, and at the same
time, as fatty acid synthesis. The [NADH]/[NAD
H11001
] ratio
in the mitochondrion is much higher than in the cytosol,
because of the flow of electrons to NAD
H11001
from the ox-
idation of fatty acids, amino acids, pyruvate, and acetyl-
CoA. This high mitochondrial [NADH]/[NAD
H11001
] ratio
favors the reduction of oxygen via the respiratory chain.
In hepatocytes and adipocytes, cytosolic NADPH is
largely generated by the pentose phosphate pathway
(see Fig. 14–21) and by malic enzyme (Fig. 21–9a).
The NADP-linked malic enzyme that operates in the
carbon-assimilation pathway of C
4
plants (see Fig.
20–23) is unrelated in function. The pyruvate produced
in the reaction shown in Figure 21–9a reenters the mi-
tochondrion. In hepatocytes and in the mammary gland
of lactating animals, the NADPH required for fatty acid
biosynthesis is supplied primarily by the pentose phos-
phate pathway (Fig. 21–9b).
In the photosynthetic cells of plants, fatty acid syn-
thesis occurs not in the cytosol but in the chloroplast
stroma (Fig. 21–8). This makes sense, given that
NADPH is produced in chloroplasts by the light reac-
tions of photosynthesis:
Again, the resulting high [NADPH]/[NADP
H11001
] ratio
provides the reducing environment that favors reduc-
tive anabolic processes such as fatty acid synthesis.
Acetate Is Shuttled out of Mitochondria as Citrate
In nonphotosynthetic eukaryotes, nearly all the acetyl-
CoA used in fatty acid synthesis is formed in mito-
H
2
OO
22
1
H
H11001
H11001NADPHH11001NADP
H11001
H11001
light
Chapter 21 Lipid Biosynthesis794
Bacteria, Plants
Seven activities
in seven separate
polypeptides
Yeast
Seven activities
in two separate
polypeptides
Vertebrates
Seven activities
in one large
polypeptide
KS
AT
MT
KR
HDER
KS
AT
MT
KR
HDER
ACP
KS
AT
KR
HDER
ACP
ACP
MT
FIGURE 21–7 Structure of fatty acid synthases. The fatty acid syn-
thase of bacteria and plants is a complex of at least seven different
polypeptides. In yeast, all seven activities reside in only two polypep-
tides; the vertebrate enzyme is a single large polypeptide.
8885d_c21_787-832 2/26/04 9:35 AM Page 794 mac76 mac76:385_reb:
chondria from pyruvate oxidation and from the catabo-
lism of the carbon skeletons of amino acids. Acetyl-CoA
arising from the oxidation of fatty acids is not a signifi-
cant source of acetyl-CoA for fatty acid biosynthesis in
animals, because the two pathways are reciprocally reg-
ulated, as described below.
The mitochondrial inner membrane is impermeable
to acetyl-CoA, so an indirect shuttle transfers acetyl
group equivalents across the inner membrane (Fig.
21–10). Intramitochondrial acetyl-CoA first reacts with
oxaloacetate to form citrate, in the citric acid cycle re-
action catalyzed by citrate synthase (see Fig. 16–7).
Citrate then passes through the inner membrane on the
citrate transporter. In the cytosol, citrate cleavage
by citrate lyase regenerates acetyl-CoA in an ATP-
dependent reaction. Oxaloacetate cannot return to the
mitochondrial matrix directly, as there is no oxaloacetate
transporter. Instead, cytosolic malate dehydrogenase re-
duces the oxaloacetate to malate, which returns to the
mitochondrial matrix on the malate–H9251-ketoglutarate
transporter in exchange for citrate. In the matrix, malate
is reoxidized to oxaloacetate to complete the shuttle. Al-
ternatively, the malate produced in the cytosol is used
to generate cytosolic NADPH through the activity of
malic enzyme (Fig. 21–9a).
Fatty Acid Biosynthesis Is Tightly Regulated
When a cell or organism has more than enough meta-
bolic fuel to meet its energy needs, the excess is gen-
erally converted to fatty acids and stored as lipids such
as triacylglycerols. The reaction catalyzed by acetyl-CoA
21.1 Biosynthesis of Fatty Acids and Eicosanoids 795
Animal cells, yeast cells
Mitochondria
Endoplasmic reticulum
Cytosol Chloroplasts Peroxisomes
Plant cells
No fatty acid
oxidation
Fatty acid oxidation
Acetyl-CoA production
Ketone body synthesis
Fatty acid elongation
NADPH production (pentose phosphate
pathway; malic enzyme)
[NADPH]/[NADP
H11001
] high
Isoprenoid and sterol synthesis
(early stages)
Fatty acid synthesis
NADPH, ATP
production
[NADPH]/[NADP
H11001
]
high
Fatty acid
synthesis
Fatty acid
oxidation
(
Phospholipid synthesis
Sterol synthesis (late stages)
Fatty acid elongation
Fatty acid desaturation
Η
2
Ο
2
)
Catalase,
peroxidase:
Η
2
Ο
2
Η
2
Ο
FIGURE 21–8 Subcellular localization of lipid metabolism. Yeast and
vertebrate cells differ from higher plant cells in the compartmentation
of lipid metabolism. Fatty acid synthesis takes place in the compart-
ment in which NADPH is available for reductive synthesis (i.e., where
the [NADPH]/[NADP
H11001
] ratio is high). Processes in red type are cov-
ered in this chapter.
malic enzyme
pentose phosphate pathway
CHOH
COO
H11002
CH
2
COO
H11002
C
CO
2
O
C
H11001
H
3
COO
H11002
Malate Pyruvate
NADP
H11001
Glucose
6-phosphate
Ribulose
5-phosphate
H11001 H
H11001
NADPH
NADP
H11001
NADPH
(a)
(b)
NADP
H11001
NADPH
FIGURE 21–9 Production of NADPH. Two routes to NADPH, cat-
alyzed by (a) malic enzyme and (b) the pentose phosphate pathway.
8885d_c21_787-832 2/26/04 9:35 AM Page 795 mac76 mac76:385_reb:
carboxylase is the rate-limiting step in the biosynthesis
of fatty acids, and this enzyme is an important site of
regulation. In vertebrates, palmitoyl-CoA, the principal
product of fatty acid synthesis, is a feedback inhibitor
of the enzyme; citrate is an allosteric activator (Fig.
21–11a), increasing V
max
. Citrate plays a central role in
diverting cellular metabolism from the consumption
(oxidation) of metabolic fuel to the storage of fuel as fatty
acids. When the concentrations of mitochondrial acetyl-
CoA and ATP increase, citrate is transported out of mi-
tochondria; it then becomes both the precursor of cyto-
solic acetyl-CoA and an allosteric signal for the activation
of acetyl-CoA carboxylase. At the same time, citrate in-
hibits the activity of phosphofructokinase-1 (see Fig.
15–18), reducing the flow of carbon through glycolysis.
Acetyl-CoA carboxylase is also regulated by cova-
lent modification. Phosphorylation, triggered by the
hormones glucagon and epinephrine, inactivates the
enzyme and reduces its sensitivity to activation by cit-
rate, thereby slowing fatty acid synthesis. In its active
(dephosphorylated) form, acetyl-CoA carboxylase poly-
merizes into long filaments (Fig. 21–11b); phosphoryla-
tion is accompanied by dissociation into monomeric
subunits and loss of activity.
Chapter 21 Lipid Biosynthesis796
Matrix Cytosol
NADH
+ H
+
Inner
membrane
Oxaloacetate
Citrate
transporter
citrate
synthase
citrate
lyase
malate
dehydrogenase
Malate
malic
enzyme
malate
dehydrogenase
α
Malate–
-ketoglutarate
transporter
Malate
pyruvate
carboxylase
Oxaloacetate
NAD
+
NADP
+
NAD
+
ADP + P
i
ADP + P
i
ATP
CO
2
CO
2
CoA-SH
Outer
membrane
Citrate
Citrate
CoAOSH
pyruvate
dehydrogenase
Amino acids
Acetyl-CoA
Pyruvate
Glucose
Acetyl-CoA
Fatty acid
synthesis
PyruvatePyruvate
Pyruvate
transporter
NADH
+ H
+
NADPH
+ H
+
ATP
FIGURE 21–10 Shuttle for transfer of acetyl groups from mitochon-
dria to the cytosol. The mitochondrial outer membrane is freely per-
meable to all these compounds. Pyruvate derived from amino acid
catabolism in the mitochondrial matrix, or from glucose by glycolysis
in the cytosol, is converted to acetyl-CoA in the matrix. Acetyl groups
pass out of the mitochondrion as citrate; in the cytosol they are de-
livered as acetyl-CoA for fatty acid synthesis. Oxaloacetate is reduced
to malate, which returns to the mitochondrial matrix and is converted
to oxaloacetate. An alternative fate for cytosolic malate is oxidation
by malic enzyme to generate cytosolic NADPH; the pyruvate pro-
duced returns to the mitochondrial matrix.
8885d_c21_787-832 2/26/04 9:35 AM Page 796 mac76 mac76:385_reb:
The acetyl-CoA carboxylase of plants and bacteria
is not regulated by citrate or by a phosphorylation-
dephosphorylation cycle. The plant enzyme is activated
by an increase in stromal pH and [Mg
2H11001
], which occurs
on illumination of the plant (see Fig. 20–18). Bacteria
do not use triacylglycerols as energy stores. In E. coli,
the primary role of fatty acid synthesis is to provide pre-
cursors for membrane lipids; the regulation of this
process is complex, involving guanine nucleotides (such
as ppGpp) that coordinate cell growth with membrane
formation (see Figs 8–42, 28–24).
In addition to the moment-by-moment regulation of
enzymatic activity, these pathways are regulated at the
level of gene expression. For example, when animals in-
gest an excess of certain polyunsaturated fatty acids,
the expression of genes encoding a wide range of li-
pogenic enzymes in the liver is suppressed. The detailed
mechanism by which these genes are regulated is not
yet clear.
If fatty acid synthesis and H9252 oxidation were to pro-
ceed simultaneously, the two processes would consti-
tute a futile cycle, wasting energy. We noted earlier (see
Fig. 17–12) that H9252 oxidation is blocked by malonyl-CoA,
which inhibits carnitine acyltransferase I. Thus during
fatty acid synthesis, the production of the first inter-
mediate, malonyl-CoA, shuts down H9252 oxidation at the
level of a transport system in the mitochondrial inner
membrane. This control mechanism illustrates another
advantage of segregating synthetic and degradative
pathways in different cellular compartments.
Long-Chain Saturated Fatty Acids Are
Synthesized from Palmitate
Palmitate, the principal product of the fatty acid syn-
thase system in animal cells, is the precursor of other
long-chain fatty acids (Fig. 21–12). It may be length-
ened to form stearate (18:0) or even longer saturated
fatty acids by further additions of acetyl groups, through
the action of fatty acid elongation systems present
in the smooth endoplasmic reticulum and in mitochon-
dria. The more active elongation system of the ER ex-
tends the 16-carbon chain of palmitoyl-CoA by two car-
bons, forming stearoyl-CoA. Although different enzyme
systems are involved, and coenzyme A rather than ACP
is the acyl carrier in the reaction, the mechanism of elon-
gation in the ER is otherwise identical to that in palmi-
tate synthesis: donation of two carbons by malonyl-CoA,
followed by reduction, dehydration, and reduction to the
saturated 18-carbon product, stearoyl-CoA.
21.1 Biosynthesis of Fatty Acids and Eicosanoids 797
Malonyl-CoA
acetyl-CoA
carboxylase
Acetyl-CoA
Citrate
Palmitoyl-CoA
glucagon,
epinephrine
trigger phosphorylation/
inactivation
insulin triggers
activation
citrate
lyase
(a) (b)
400 ?
FIGURE 21–11 Regulation of fatty acid synthesis. (a) In the cells of
vertebrates, both allosteric regulation and hormone-dependent cova-
lent modification influence the flow of precursors into malonyl-CoA.
In plants, acetyl-CoA carboxylase is activated by the changes in [Mg
2H11001
]
and pH that accompany illumination (not shown here). (b) Filaments
of acetyl-CoA carboxylase (the active, dephosphorylated form) as seen
with the electron microscope.
Palmitate
16:0
elongation
Stearate
18:0
desaturation
Oleate
18:1(H9004
9
)
desaturation
(in plants
only)
Linoleate
18:2(H9004
9,12
)
desaturation
(in plants
only)
desaturation
H9251-Linolenate
18:3(H9004
9,12,15
)
Other polyunsaturated
fatty acids
H9253-Linolenate
18:3(H9004
6,9,12
)
Eicosatrienoate
20:3(H9004
8,11,14
)
elongation
Arachidonate
20:4(H9004
5,8,11,14
)
desaturation
desaturation
Palmitoleate
16:1(H9004
9
)
elongation
Longer saturated
fatty acids
FIGURE 21–12 Routes of synthesis of other fatty acids. Palmitate is
the precursor of stearate and longer-chain saturated fatty acids, as well
as the monounsaturated acids palmitoleate and oleate. Mammals can-
not convert oleate to linoleate or H9251-linolenate (shaded pink), which
are therefore required in the diet as essential fatty acids. Conversion
of linoleate to other polyunsaturated fatty acids and eicosanoids is out-
lined. Unsaturated fatty acids are symbolized by indicating the num-
ber of carbons and the number and position of the double bonds, as
in Table 10–1.
8885d_c21_787-832 2/26/04 9:35 AM Page 797 mac76 mac76:385_reb:
Desaturation of Fatty Acids Requires a
Mixed-Function Oxidase
Palmitate and stearate serve as precursors of the two
most common monounsaturated fatty acids of animal
tissues: palmitoleate, 16:1(H9004
9
), and oleate, 18:1(H9004
9
);
both of these fatty acids have a single cis double bond
between C-9 and C-10 (see Table 10–1). The double
bond is introduced into the fatty acid chain by an ox-
idative reaction catalyzed by fatty acyl–CoA desatu-
rase (Fig. 21–13), a mixed-function oxidase (Box
21–1). Two different substrates, the fatty acid and
NADH or NADPH, simultaneously undergo two-electron
oxidations. The path of electron flow includes a cyto-
chrome (cytochrome b
5
) and a flavoprotein (cyto-
chrome b
5
reductase), both of which, like fatty acyl–CoA
desaturase, are in the smooth ER. Bacteria have two
cytochrome b
5
reductases, one NADH-dependent and
the other NADPH-dependent; which of these is the
main electron donor in vivo is unclear. In plants, oleate
is produced by a stearoyl-ACP desaturase in the chloro-
plast stroma that uses reduced ferredoxin as the elec-
tron donor.
Mammalian hepatocytes can readily introduce dou-
ble bonds at the H9004
9
position of fatty acids but cannot in-
troduce additional double bonds between C-10 and the
Chapter 21 Lipid Biosynthesis798
BOX 21–1 THE WORLD OF BIOCHEMISTRY
Mixed-Function Oxidases, Oxygenases, and
Cytochrome P-450
In this chapter we encounter several enzymes that
carry out oxidation-reduction reactions in which
molecular oxygen is a participant. The reaction that
introduces a double bond into a fatty acyl chain
(see Fig. 21–13) is one such reaction.
The nomenclature for enzymes that catalyze re-
actions of this general type is often confusing to stu-
dents, as is the mechanism of the reactions. Oxidase
is the general name for enzymes that catalyze oxida-
tions in which molecular oxygen is the electron ac-
ceptor but oxygen atoms do not appear in the oxidized
product (however, there is an exception to this “rule,”
as we shall see!). The enzyme that creates a double
bond in fatty acyl–CoA during the oxidation of fatty
acids in peroxisomes (see Fig. 17–13) is an oxidase of
this type; a second example is the cytochrome oxidase
of the mitochondrial electron-transfer chain (see Fig.
19–14). In the first case, the transfer of two electrons
to H
2
O produces hydrogen peroxide, H
2
O
2
; in the sec-
ond, two electrons reduce
H5007
1
2
H5007
O
2
to H
2
O. Many, but not
all, oxidases are flavoproteins.
Oxygenases catalyze oxidative reactions in
which oxygen atoms are directly incorporated into
the substrate molecule, forming a new hydroxyl or
carboxyl group, for example. Dioxygenases cat-
alyze reactions in which both oxygen atoms of O
2
are incorporated into the organic substrate mole-
cule. An example of a dioxygenase is tryptophan 2,
3-dioxygenase, which catalyzes the opening of the
five-membered ring of tryptophan in the catabolism
of this amino acid. When this reaction takes place in
the presence of
18
O
2
, the isotopic oxygen atoms are
found in the two carbonyl groups of the product
(shown in red).
Monooxygenases, more abundant and more
complex in their action, catalyze reactions in which
only one of the two oxygen atoms of O
2
is incorpo-
rated into the organic substrate, the other being re-
duced to H
2
O. Monooxygenases require two sub-
strates to serve as reductants of the two oxygen atoms
of O
2
. The main substrate accepts one of the two oxy-
gen atoms, and a cosubstrate furnishes hydrogen
atoms to reduce the other oxygen atom to H
2
O. The
general reaction equation for monooxygenases is
AH H11001 BH
2
H11001 OOO 88n AOOH H11001 B H11001 H
2
O
where AH is the main substrate and BH
2
the cosub-
strate. Because most monooxygenases catalyze reac-
tions in which the main substrate becomes hydroxy-
lated, they are also called hydroxylases. They are
COO
H5008
CH
2
H
NH
3
CH
H11001
O
2
H
NH C
O
CH COO
H5008
NH
3
H11001
tryptophan
2,3-dioxygenase
N-Formylkynurenine
N
C
O
CH
2
Tryptophan
8885d_c21_787-832 2/26/04 9:35 AM Page 798 mac76 mac76:385_reb:
21.1 Biosynthesis of Fatty Acids and Eicosanoids 799
also sometimes called mixed-function oxidases or
mixed-function oxygenases, to indicate that they
oxidize two different substrates simultaneously. (Note
here the use of “oxidase”—a deviation from the gen-
eral meaning of this term noted above.)
There are different classes of monooxygenases,
depending on the nature of the cosubstrate. Some
use reduced flavin nucleotides (FMNH
2
or FADH
2
),
others use NADH or NADPH, and still others use H9251-
ketoglutarate as the cosubstrate. The enzyme that
hydroxylates the phenyl ring of phenylalanine to
form tyrosine is a monooxygenase for which tetrahy-
drobiopterin serves as cosubstrate (see Fig. 18–23).
This is the enzyme that is defective in the human ge-
netic disease phenylketonuria.
The most numerous and most complex monooxy-
genation reactions are those employing a type of heme
protein called cytochrome P-450. This cytochrome
is usually present in the smooth ER rather than the
mitochondria. Like mitochondrial cytochrome oxi-
dase, cytochrome P-450 can react with O
2
and bind
carbon monoxide, but it can be differentiated from cy-
tochrome oxidase because the carbon monoxide com-
plex of its reduced form absorbs light strongly at
450 nm—thus the name P-450.
Cytochrome P-450 catalyzes hydroxylation reac-
tions in which an organic substrate, RH, is hydroxy-
lated to ROOH, incorporating one oxygen atom of O
2
;
the other oxygen atom is reduced to H
2
O by reducing
equivalents that are furnished by NADH or NADPH
but are usually passed to cytochrome P-450 by an iron-
sulfur protein. Figure 1 shows a simplified outline of
the action of cytochrome P-450, which has interme-
diate steps not yet fully understood.
Cytochrome P-450 is actually a family of similar
proteins; several hundred members of this protein
family are known, each with a different substrate
specificity. In the adrenal cortex, for example, a spe-
cific cytochrome P-450 participates in the hydroxyla-
tion of steroids to yield the adrenocortical hormones
(see Fig. 21–47). Cytochrome P-450 is also important
in the hydroxylation of many different drugs, such as
barbiturates and other xenobiotics (substances for-
eign to the organism), particularly if they are hy-
drophobic and relatively insoluble. The environmen-
tal carcinogen benzo[a]pyrene (found in cigarette
smoke) undergoes cytochrome P-450–dependent
hydroxylation during detoxification. Hydroxylation of
xenobiotics makes them more soluble in water and
allows their excretion in the urine. Unfortunately, hy-
droxylation of some compounds converts them to toxic
substances, subverting the detoxification system.
Reactions described in this chapter that are cat-
alyzed by mixed-function oxidases are those involved
in fatty acyl–CoA desaturation (Fig. 21–13), leukotri-
ene synthesis (Fig. 21–16), plasmalogen synthesis
(Fig. 21–30), conversion of squalene to cholesterol
(Fig. 21–37), and steroid hormone synthesis (Fig.
21–47).
FIGURE 1
NADPH Oxidized Reduced RH
Reduced OxidizedNADP
H11001
O
2
H
2
O
ROH
cytochrome
P-450 reductase
(Fe–S)
cytochrome
P-450
2H
2
O
Monounsaturated
fatty acyl–CoA
O
CH
3
CH
2
CH
S-CoA
(CH
2
)
m
fatty acyl–
CoA desaturase
H11001
H11001 H11001O
2
Cyt b
5
reductase
(FADH
2
)
Cyt b
5
reductase
(FAD)
2 Cyt b
5
(Fe
2H11001
)
2 Cyt b
5
(Fe
3H11001
)
2H
H11001
NADP
H11001
(CH
2
)
n
CH
2
O
CH
3
C
S-CoA
(CH
2
)
m
(CH
2
)
n
CCH
Saturated
fatty acyl–CoA
NADPH
H11001 H
H11001
FIGURE 21–13 Electron transfer in the desaturation of fatty acids in vertebrates.
Blue arrows show the path of electrons as two substrates—a fatty acyl–CoA and
NADPH—undergo oxidation by molecular oxygen. These reactions take place on the
lumenal face of the smooth ER. A similar pathway, but with different electron carriers,
occurs in plants.
8885d_c21_787-832 2/26/04 9:35 AM Page 799 mac76 mac76:385_reb:
methyl-terminal end. Thus mammals cannot synthesize
linoleate, 18:2(H9004
9,12
), or H9251-linolenate, 18:3(H9004
9,12,15
).
Plants, however, can synthesize both; the desaturases
that introduce double bonds at the H9004
12
and H9004
15
positions
are located in the ER and the chloroplast. The ER en-
zymes act not on free fatty acids but on a phospholipid,
phosphatidylcholine, that contains at least one oleate
linked to the glycerol (Fig. 21–14). Both plants and bac-
teria must synthesize polyunsaturated fatty acids to
ensure membrane fluidity at reduced temperatures.
Because they are necessary precursors for the syn-
thesis of other products, linoleate and linolenate are es-
sential fatty acids for mammals; they must be ob-
tained from dietary plant material. Once ingested,
linoleate may be converted to certain other polyunsat-
urated acids, particularly H9253-linolenate, eicosatrienoate,
and arachidonate (eicosatetraenoate), all of which can
be made only from linoleate (Fig. 21–12). Arachidonate,
20:4(H9004
5,8,11,14
), is an essential precursor of regulatory
lipids, the eicosanoids. The 20-carbon fatty acids are
synthesized from linoleate (and linolenate) by fatty acid
elongation reactions analogous to those described on
page 797.
Eicosanoids Are Formed from 20-Carbon
Polyunsaturated Fatty Acids
Eicosanoids are a family of very potent biological sig-
naling molecules that act as short-range messengers, af-
fecting tissues near the cells that produce them. In re-
sponse to hormonal or other stimuli, phospholipase A
2
,
present in most types of mammalian cells, attacks mem-
brane phospholipids, releasing arachidonate from the
middle carbon of glycerol. Enzymes of the smooth ER
then convert arachidonate to prostaglandins, begin-
ning with the formation of prostaglandin H
2
(PGH
2
), the
immediate precursor of many other prostaglandins and
of thromboxanes (Fig. 21–15a). The two reactions that
lead to PGH
2
are catalyzed by a bifunctional enzyme,
cyclooxygenase (COX), also called prostaglandin
H
2
synthase. In the first of two steps, the cyclooxyge-
nase activity introduces molecular oxygen to convert
arachidonate to PGG
2
. The second step, catalyzed by
the peroxidase activity of COX, converts PGG
2
to PGH
2
.
Aspirin (acetylsalicylate; Fig. 21–15b) irre-
versibly inactivates the cyclooxygenase activity
of COX by acetylating a Ser residue and blocking the
enzyme’s active site, thus inhibiting the synthesis of
prostaglandins and thromboxanes. Ibuprofen, a widely
used nonsteroidal antiinflammatory drug (NSAID; Fig.
21–15c), inhibits the same enzyme. The recent discov-
ery that there are two isozymes of COX has led to the
development of more precisely targeted NSAIDs with
fewer undesirable side effects (Box 21–2).
Thromboxane synthase, present in blood
platelets (thrombocytes), converts PGH
2
to thrombox-
ane A
2
, from which other thromboxanes are derived
(Fig. 21–15a). Thromboxanes induce constriction of
blood vessels and platelet aggregation, early steps in
blood clotting. Low doses of aspirin, taken regularly, re-
duce the probability of heart attacks and strokes by re-
ducing thromboxane production. ■
Thromboxanes, like prostaglandins, contain a ring
of five or six atoms; the pathway from arachidonate to
these two classes of compounds is sometimes called the
“cyclic” pathway, to distinguish it from the “linear” path-
way that leads from arachidonate to the leukotrienes,
which are linear compounds (Fig. 21–16). Leukotriene
synthesis begins with the action of several lipoxygenases
that catalyze the incorporation of molecular oxygen into
arachidonate. These enzymes, found in leukocytes and
in heart, brain, lung, and spleen, are mixed-function ox-
idases that use cytochrome P-450 (Box 21–1). The var-
ious leukotrienes differ in the position of the peroxide
Chapter 21 Lipid Biosynthesis800
CH O C
O
2
OP
H11001
desaturase
1
CH
2
OC
O
O
O N(CH
3
)
3
H
2
C H
2
CH
O
H5008
C
Phosphatidylcholine containing
linolenate, 18:3(H9004
9,12,15
)
CH O C
O
2
OP
H11001
9
1
18
CH
2
OC
O
O
O N(CH
3
)
3
H
2
C H
2
CH
O
H5008
C
Phosphatidylcholine containing
oleate, 18:1(H9004
9
)
9
desaturase
CH O C
O
2
OP
H11001
1
12
CH
2
OC
O
O
O N(CH
3
)
3
H
2
C H
2
CH
O
H5008
C
Phosphatidylcholine containing
linoleate, 18:2(
H9004
9,12
)
9
18
12 1815
FIGURE 21–14 Action of plant desaturases. Desaturases in plants ox-
idize phosphatidylcholine-bound oleate to polyunsaturated fatty acids.
Some of the products are released from the phosphatidylcholine by
hydrolysis.
8885d_c21_787-832 2/26/04 9:35 AM Page 800 mac76 mac76:385_reb:
COO
H5008
O
OH
Phospholipid containing
arachidonate
phospholipase A
2
Lysophospholipid
Arachidonate,
20:4(H9004
5,8,11,14
)
aspirin, ibuprofen
peroxidase
activity
of COX
PGH
2
2O
2
COO
H5008
O
COO
H5008
O
OOH
PGG
2
O
Other
prostaglandins
Thromboxanes
(a)
cyclooxgenase
activity of COX
H11001
CH
3
Ser O C
COO
H5008
Ibuprofen
O
Acetylated,
inactivated
COX
H11001
CH
3
Ser
O
OH
C
COO
H5008
Aspirin
(acetylsalicylate)
O
COX
(c)
OH
CH
3
CH
2
COO
H5008
CH
3
Salicylate
(b)
CH
CH
CH
3
Naproxen
CH
CH
3
COO
H5008
O
CH
3
21.1 Biosynthesis of Fatty Acids and Eicosanoids 801
FIGURE 21–15 The “cyclic” pathway from arachidonate to pros-
taglandins and thromboxanes. (a) After arachidonate is released from
phospholipids by the action of phospholipase A
2
, the cyclooxygenase
and peroxidase activities of COX (also called prostaglandin H
2
syn-
thase) catalyze the production of PGH
2
, the precursor of other
prostaglandins and thromboxanes. (b) Aspirin inhibits the first reac-
tion by acetylating an essential Ser residue on the enzyme. (c) Ibupro-
fen and naproxen inhibit the same step, probably by mimicking the
structure of the substrate or an intermediate in the reaction.
114
5
85
COO
H5008
Arachidonate
LTD
4
HOO
5-Hydroperoxyeicosatetraenoate
(5-HPETE)
12-Hydroperoxyeicosatetraenoate
(12-HPETE)
OOH
Leukotriene A
4
(LTA
4
)
LTC
4
20
Other leukotrienes
12
COO
H5008
lipoxygenase
COO
H5008
11
O
2
lipoxygenase
multistep
multistep
O
2
FIGURE 21–16 The “linear” pathway from arachidonate to
leukotrienes.
8885d_c21_787-832 2/26/04 9:35 AM Page 801 mac76 mac76:385_reb:
group introduced by the lipoxygenases. This linear path-
way from arachidonate, unlike the cyclic pathway, is not
inhibited by aspirin or other NSAIDs.
Plants also derive important signaling molecules
from fatty acids. As in animals, a key step in the initia-
tion of signaling involves activation of a specific phos-
pholipase. In plants, the fatty acid substrate that is re-
leased is H9251-linolenate. A lipoxygenase then catalyzes the
first step in a pathway that converts linolenate to jas-
monate, a substance known to have signaling roles in
insect defense, resistance to fungal pathogens, and
pollen maturation. Jasmonate (see Fig. 12–28) also af-
fects seed germination, root growth, and fruit and seed
development.
Chapter 21 Lipid Biosynthesis802
BOX 21–2 BIOCHEMISTRY IN MEDICINE
Relief Is in (the Active) Site: Cyclooxygenase
Isozymes and the Search for a Better Aspirin
Each year, several thousand tons of aspirin (acetyl-
salicylate) are consumed around the world for the re-
lief of headaches, sore muscles, inflamed joints, and
fever. Because aspirin inhibits platelet aggregation and
blood clotting, it is also used in low doses to treat pa-
tients at risk of heart attacks. The medicinal proper-
ties of the compounds known as salicylates, including
aspirin, were first described by western science in
1763, when Edmund Stone of England noted that bark
of the willow tree Salix alba was effective against
fevers, aches, and pains. By the 1830s, German
chemists had purified the active components from wil-
low and from another plant rich in salicylates, the
meadowsweet, Spiraea ulmaria. However, salicylate
itself was bitter-tasting and its use had some un-
pleasant side effects, including severe stomach irrita-
tion in some cases. To address these problems, Felix
Hoffmann and Arthur Eichengrun synthesized acetyl-
salicylate at the Bayer company in Germany in 1897.
The new compound, with fewer side effects than sal-
icylate, was marketed in 1899 under the trade name
Aspirin (from a for acetyl and spir for Spirsaüre, the
German word for the acid prepared from Spiraea).
Within a few years, aspirin was in widespread use.
Aspirin (now a generic name) is one of a number
of nonsteroidal antiinflammatory drugs (NSAIDs);
others include ibuprofen and naproxen (see Fig.
21–15), all now sold over the counter. Unfortunately,
aspirin reduces but does not eliminate the side effects
of salicylates. In some patients, aspirin itself can pro-
duce stomach bleeding, kidney failure, and, in extreme
cases, death. New NSAIDs with the beneficial effects
of aspirin but without its side effects would be med-
ically valuable.
Aspirin and other NSAIDs inhibit the cyclooxyge-
nase activity of prostaglandin H
2
synthase (also called
COX, for cyclooxygenase), which adds molecular oxy-
gen to arachidonate to initiate prostaglandin synthe-
sis (see Fig. 21–15a). Prostaglandins regulate many
physiological processes, including platelet aggrega-
tion, uterine contractions, pain, inflammation, and the
secretion of mucins that protect the gastric mucosa
from acid and proteolytic enzymes in the stomach. The
stomach irritation that is a common side effect of as-
pirin use results from the drug’s interference with the
secretion of gastric mucin.
Mammals have two isozymes of prostaglandin H
2
synthase, COX-1 and COX-2. These have different
functions but closely similar amino acid sequences
(60% to 65% sequence identity) and similar reaction
mechanisms at both of their catalytic centers. COX-1
is responsible for the synthesis of the prostaglandins
that regulate the secretion of gastric mucin, and COX-
2 for the prostaglandins that mediate inflammation,
pain, and fever. Aspirin inhibits both isozymes about
equally, so a dose sufficient to reduce inflammation
also risks stomach irritation. Much research is aimed
at developing new NSAIDs that inhibit COX-2 specif-
ically, and several such drugs have become available.
The development of COX-2–specific inhibitors has
been helped immensely by knowledge of the detailed
three-dimensional structures of COX-1 and COX-2
(Fig. 1). Both proteins are homodimers. Each mono-
mer (M
r
70,000) has an amphipathic domain that
penetrates but does not span the ER; this anchors the
enzyme on the lumenal side of the ER (a very unusual
topology—generally the hydrophobic regions of inte-
gral membrane proteins span the entire bilayer). Both
catalytic sites are on the globular domain protruding
into the ER lumen.
COX-1 and COX-2 have virtually identical tertiary
and quaternary structures, but they differ subtly in a
long, thin hydrophobic channel extending from the
membrane interior to the lumenal surface. The chan-
nel includes both catalytic sites and is presumed to be
the binding site for the hydrophobic substrate, arachi-
donate. Both COX-1 and COX-2 have been crystallized
in the presence of several different bound NSAID
compounds, defining the NSAID-binding site (Fig. 1).
The bound drugs block the hydrophobic channel and
prevent arachidonate entry. The subtle differences be-
tween the channels of COX-1 and COX-2 have guided
8885d_c21_787-832 2/26/04 9:35 AM Page 802 mac76 mac76:385_reb:
21.1 Biosynthesis of Fatty Acids and Eicosanoids 803
SUMMARY 21.1 Biosynthesis of Fatty Acids
and Eicosanoids
■ Long-chain saturated fatty acids are
synthesized from acetyl-CoA by a cytosolic
complex of six enzyme activities plus acyl
carrier protein (ACP). The fatty acid synthase
complex, which in some organisms consists of
multifunctional polypeptides, contains two
types of OSH groups (one furnished by the
phosphopantetheine of ACP, the other by a
Cys residue of H9252-ketoacyl-ACP synthase)
that function as carriers of the fatty acyl
intermediates.
the design of NSAIDs that selectively fit COX-2 and
therefore inhibit COX-2 more effectively than COX-1.
Two of these drugs have been approved for use world-
wide: celecoxib (Celebrex) for osteoarthritis and
rheumatoid arthritis, and rofecoxib (Vioxx) for os-
teoarthritis and acute musculoskeletal pain (Fig. 2).
In clinical trials, these drugs have proven effective
while significantly reducing the stomach irritation and
other side effects of aspirin and other NSAIDs. The
use of precise structural information about an en-
zyme’s active site is a powerful tool in the develop-
ment of better, more specific drugs.
CF
3
S
O
H
2
N
N
N
H
3
C
Celecoxib
(Celebrex)
O
O
O
Rofecoxib
(Vioxx)
S
O
H
3
C
O
(b)
(a)
FIGURE 1 Structures of COX-1 and COX-2. (a) COX-1 with an
NSAID inhibitor (flurbiprofen, orange) bound (PDB ID 3PGH). The
enzyme consists of two identical monomers
(gray and blue) each with three domains: a mem-
brane anchor consisting of four amphipathic
helices; a second domain that somewhat resem-
bles a domain of the epidermal growth factor;
and the catalytic domain, which contains the
cyclooxygenase and peroxidase activities, as well
as the hydrophobic channel in which the sub-
strate (arachidonate) binds. The heme that is part
of the peroxidase active sites is shown in red;
Tyr
385
, a key residue in the cyclooxygenase site,
is turquoise. Other catalytically important residues include Arg
120
(dark blue), His
388
(green), and Ser
530
(yellow). Flurbiprofen blocks
access to the enzyme active site.
(b) A look at COX-1 and COX-2 side by side. The COX-1 en-
zyme from sheep is shown at left, with bound ibuprofen (PDB ID
1EQG). The COX-2 enzyme from mouse is shown at right, with a
similar inhibitor, Sc-558, bound (PDB ID 6COX). The inhibitors (red)
are partially buried within the structures in these representations,
which emphasize surface contours. COX-1 and COX-2 are very sim-
ilar in structure, but enzymologists exploit the small differences in
the structures of the active sites and the channels leading to them
to design inhibitors specific for one enzyme or the other.
FIGURE 2 Two COX-2–specific drugs that bind to COX-2 about
1,000 times better than to COX-1.
F
CH
CH
3
COO
H5008
Flurbiprofen
8885d_c21_787-832 2/26/04 9:35 AM Page 803 mac76 mac76:385_reb:
■ Malonyl-ACP, formed from acetyl-CoA (shuttled
out of mitochondria) and CO
2
, condenses with
an acetyl bound to the CysOSH to yield
acetoacetyl-ACP, with release of CO
2
. This is
followed by reduction to the D-H9252-hydroxy
derivative, dehydration to the trans-H9004
2
-
unsaturated acyl-ACP, and reduction to
butyryl-ACP. NADPH is the electron donor
for both reductions. Fatty acid synthesis
is regulated at the level of malonyl-CoA
formation.
■ Six more molecules of malonyl-ACP react
successively at the carboxyl end of the growing
fatty acid chain to form palmitoyl-ACP—the
end product of the fatty acid synthase reaction.
Free palmitate is released by hydrolysis.
■ Palmitate may be elongated to the 18-carbon
stearate. Palmitate and stearate can be
desaturated to yield palmitoleate and oleate,
respectively, by the action of mixed-function
oxidases.
■ Mammals cannot make linoleate and must
obtain it from plant sources; they convert
exogenous linoleate to arachidonate, the parent
compound of eicosanoids (prostaglandins,
thromboxanes, and leukotrienes), a family of
very potent signaling molecules.
21.2 Biosynthesis of Triacylglycerols
Most of the fatty acids synthesized or ingested by an or-
ganism have one of two fates: incorporation into tria-
cylglycerols for the storage of metabolic energy or in-
corporation into the phospholipid components of
membranes. The partitioning between these alternative
fates depends on the organism’s current needs. During
rapid growth, synthesis of new membranes requires the
production of membrane phospholipids; when an or-
ganism has a plentiful food supply but is not actively
growing, it shunts most of its fatty acids into storage
fats. Both pathways begin at the same point: the for-
mation of fatty acyl esters of glycerol. In this section we
examine the route to triacylglycerols and its regulation,
and the production of glycerol 3-phosphate in the
process of glyceroneogenesis.
Triacylglycerols and Glycerophospholipids Are
Synthesized from the Same Precursors
Animals can synthesize and store large quantities of tri-
acylglycerols, to be used later as fuel (see Box 17–1).
Humans can store only a few hundred grams of glyco-
gen in liver and muscle, barely enough to supply the
body’s energy needs for 12 hours. In contrast, the total
amount of stored triacylglycerol in a 70-kg man of
average build is about 15 kg, enough to support basal
energy needs for as long as 12 weeks (see Table 23–5).
Triacylglycerols have the highest energy content of all
stored nutrients—more than 38 kJ/g. Whenever carbo-
hydrate is ingested in excess of the organism’s capacity
to store glycogen, the excess is converted to triacyl-
glycerols and stored in adipose tissue. Plants also man-
ufacture triacylglycerols as an energy-rich fuel, mainly
stored in fruits, nuts, and seeds.
In animal tissues, triacylglycerols and glycerophos-
pholipids such as phosphatidylethanolamine share two
precursors (fatty acyl–CoA and L-glycerol 3-phosphate)
and several biosynthetic steps. The vast majority of
the glycerol 3-phosphate is derived from the glycolytic
intermediate dihydroxyacetone phosphate (DHAP)
by the action of the cytosolic NAD-linked glycerol
3-phosphate dehydrogenase; in liver and kidney, a
small amount of glycerol 3-phosphate is also formed
from glycerol by the action of glycerol kinase (Fig.
21–17). The other precursors of triacylglycerols are
fatty acyl–CoAs, formed from fatty acids by acyl-CoA
synthetases, the same enzymes responsible for the ac-
tivation of fatty acids for H9252 oxidation (see Fig. 17–5).
The first stage in the biosynthesis of triacylglycerols
is the acylation of the two free hydroxyl groups of L-
glycerol 3-phosphate by two molecules of fatty acyl–CoA
to yield diacylglycerol 3-phosphate, more commonly
called phosphatidic acid or phosphatidate (Fig. 21–17).
Phosphatidic acid is present in only trace amounts in
cells but is a central intermediate in lipid biosynthesis;
it can be converted either to a triacylglycerol or to a
glycerophospholipid. In the pathway to triacylglycerols,
phosphatidic acid is hydrolyzed by phosphatidic acid
phosphatase to form a 1,2-diacylglycerol (Fig. 21–18).
Diacylglycerols are then converted to triacylglycerols by
transesterification with a third fatty acyl–CoA.
Triacylglycerol Biosynthesis in Animals Is Regulated
by Hormones
In humans, the amount of body fat stays relatively con-
stant over long periods, although there may be minor
short-term changes as caloric intake fluctuates. Carbo-
hydrate, fat, or protein consumed in excess of energy
needs is stored in the form of triacylglycerols that can
be drawn upon for energy, enabling the body to with-
stand periods of fasting.
Biosynthesis and degradation of triacylglycerols
are regulated such that the favored path de-
pends on the metabolic resources and requirements of
the moment. The rate of triacylglycerol biosynthesis is
profoundly altered by the action of several hormones.
Insulin, for example, promotes the conversion of car-
bohydrate to triacylglycerols (Fig. 21–19). People with
severe diabetes mellitus, due to failure of insulin se-
cretion or action, not only are unable to use glucose
properly but also fail to synthesize fatty acids from
Chapter 21 Lipid Biosynthesis804
8885d_c21_787-832 2/26/04 9:35 AM Page 804 mac76 mac76:385_reb:
O
C
P
CHOH
HO
S-CoA
CH
2
OH
Glucose
glycolysis
Dihydroxyacetone
phosphate
glycerol 3-phosphate
dehydrogenase
glycerol
kinase
Glycerol
CoA-SH
L-Glycerol 3-phosphate
acyl transferase
Phosphatidic acid
O
H5008
CH
2
CH
2
OH
C
CH
2
O
O
P
O
O
H5008
O
H5008
CH
2
OH
C
CH
2
O P
O
O
H5008
O
H5008
H
H
ADP
CH
2
OH
R
2
C
O
ATP
R
2
AMP
R
1
S-CoA
COO
H5008
CoA-SH
acyl-CoA
synthetase
C
O
ATP
H11001
R
1
CoA-SH
R
1
CO
O
C
O
CH
2
O
O
O
H5008
acyl transferase
PP
AMP
R
2
COO
H5008
acyl-CoA
synthetase
ATP
H11001
CoA-SH
PP
i
H11001 H
H11001
NAD
H11001
NADH
carbohydrates or amino acids. If the diabetes is un-
treated, these individuals have increased rates of fat
oxidation and ketone body formation (Chapter 17) and
therefore lose weight. ■
An additional factor in the balance between biosyn-
thesis and degradation of triacylglycerols is that ap-
proximately 75% of all fatty acids released by lipolysis
are reesterified to form triacylglycerols rather than used
for fuel. This ratio persists even under starvation con-
ditions, when energy metabolism is shunted from the
use of carbohydrate to the oxidation of fatty acids. Some
of this fatty acid recycling takes place in adipose tissue,
with the reesterification occurring before release into
the bloodstream; some takes place via a systemic cycle
in which free fatty acids are transported to liver, recy-
cled to triacylglycerol, exported again to the blood
(transport of lipids in the blood is discussed in Section
21.4), and taken up again by adipose tissue after release
from triacylglycerol by extracellular lipoprotein lipase
21.2 Biosynthesis of Triacylglycerols 805
FIGURE 21–17 Biosynthesis of phosphatidic acid. A fatty acyl group
is activated by formation of the fatty acyl–CoA, then transferred to es-
ter linkage with L-glycerol 3-phosphate, formed in either of the two
ways shown. Phosphatidic acid is shown here with the correct stere-
ochemistry at C-2 of the glycerol molecule. To conserve space in sub-
sequent figures (and in Fig. 21–14), both fatty acyl groups of glyc-
erophospholipids, and all three acyl groups of triacylglycerols, are
shown projecting to the right.
O
CCH
PCH
2
R
2
OC
O
R
1
O
O
O
H5008
CH
2
O
Glycerophospholipid
attachment of
head group
(serine, choline,
ethanolamine, etc.)
phosphatidic acid
phosphatase
OC
CH
CH
2
R
3
OC
O
R
1
O
O
CH
2
O
Triacylglycerol
CCH
CH
2
R
3
OC
O
R
1
O
CH
2
O
1,2-Diacylglycerol
OH
O
Head
group
R
2
C
O
S-CoA
CoA-SH
acyl
transferase
O
CCH
PO
H5008
CH
2
R
2
OC
O
R
1
O
O
O
H5008
CH
2
O Phosphatidic acid
CR
2
FIGURE 21–18 Phosphatidic acid in lipid biosynthesis. Phosphatidic
acid is the precursor of both triacylglycerols and glycerophospholipids.
The mechanisms for head-group attachment in phospholipid synthe-
sis are described later in this section.
8885d_c21_787-832 2/26/04 9:35 AM Page 805 mac76 mac76:385_reb:
(Fig. 21–20; see also Fig. 17–1). Flux through this tri-
acylglycerol cycle between adipose tissue and liver
may be quite low when other fuels are available and the
release of fatty acids from adipose tissue is limited, but
as noted above, the proportion of released fatty acids
that are reesterified remains roughly constant at 75%
under all metabolic conditions. The level of free fatty
acids in the blood thus reflects both the rate of release
of fatty acids and the balance between the synthesis and
breakdown of triacylglycerols in adipose tissue and liver.
When the mobilization of fatty acids is required to
meet energy needs, release from adipose tissue is stim-
ulated by the hormones glucagon and epinephrine (see
Figs 17–3, 17–12). Simultaneously, these hormonal sig-
nals decrease the rate of glycolysis and increase the rate
of gluconeogenesis in the liver (providing glucose for
the brain, as further elaborated in Chapter 23). The re-
leased fatty acid is taken up by a number of tissues, in-
cluding muscle, where it is oxidized to provide energy.
Much of the fatty acid taken up by liver is not oxidized
but is recycled to triacylglycerol and returned to adi-
pose tissue.
The function of the apparently futile triacylglycerol
cycle (futile cycles are discussed in Chapter 15) is not
well understood. However, as we learn more about how
the triacylglycerol cycle is sustained via metabolism in
two separate organs and is coordinately regulated, some
possibilities emerge. For example, the excess capacity
in the triacylglycerol cycle (the fatty acid that is even-
tually reconverted to triacylglycerol rather than oxi-
dized as fuel) could represent an energy reserve in the
bloodstream during fasting, one that would be more rap-
idly mobilized in a “fight or flight” emergency than would
stored triacylglycerol.
The constant recycling of triacylglycerols in adipose
tissue even during starvation raises a second question:
what is the source of the glycerol 3-phosphate required
for this process? As noted above, glycolysis is sup-
pressed in these conditions by the action of glucagon
and epinephrine, so little DHAP is available, and glyc-
erol released during lipolysis cannot be converted di-
rectly to glycerol 3-phosphate in adipose tissue, because
these cells lack glycerol kinase (Fig. 21–17). So, how is
sufficient glycerol 3-phosphate produced? The answer
lies in a pathway discovered more than three decades
ago and given little attention until recently, a pathway
intimately linked to the triacylglycerol cycle and, in a
larger sense, to the balance between fatty acid and car-
bohydrate metabolism.
Adipose Tissue Generates Glycerol 3-phosphate
by Glyceroneogenesis
Glyceroneogenesis is a shortened version of gluco-
neogenesis, from pyruvate to DHAP (see Fig. 14–16),
followed by conversion of the DHAP to glycerol 3-
phosphate by cytosolic NAD-linked glycerol 3-phosphate
dehydrogenase (Fig. 21–21). Glycerol 3-phosphate is
subsequently used in triacylglycerol synthesis. Glycero-
Chapter 21 Lipid Biosynthesis806
Dietary
carbohydrates
Glucose
Dietary
proteins
Amino acids
Acetyl-CoA
Fatty acids
Triacylglycerols
Ketone bodies
(acetoacetate,
D- -hydroxybutyrate,
acetone)
insulin
increased
in diabetes
H9252
FIGURE 21–19 Regulation of triacylglycerol synthesis by in-
sulin. Insulin stimulates conversion of dietary carbohydrates
and proteins to fat. Individuals with diabetes mellitus lack insulin; in
uncontrolled disease, this results in diminished fatty acid synthesis,
and the acetyl-CoA arising from catabolism of carbohydrates and pro-
teins is shunted instead to ketone body production. People in severe
ketosis smell of acetone, so the condition is sometimes mistaken for
drunkenness (p. 909).
Adipose tissue
Glycerol
3-phosphate
Glycerol
3-phosphate
Fuel for
tissues
Glycerol
Triacylglycerol Triacylglycerol
Fatty
acid
Fatty
acid
Lipoprotein
lipase
LiverBlood
Glycerol
FIGURE 21–20 The triacylglycerol cycle. In mammals, triacylglycerol
molecules are broken down and resynthesized in a triacylglycerol cy-
cle during starvation. Some of the fatty acids released by lipolysis of
triacylglycerol in adipose tissue pass into the bloodstream, and the re-
mainder are used for resynthesis of triacylglycerol. Some of the fatty
acids released into the blood are used for energy (in muscle, for ex-
ample), and some are taken up by the liver and used in triacylglyc-
erol synthesis. The triacylglycerol formed in the liver is transported in
the blood back to adipose tissue, where the fatty acid is released by
extracellular lipoprotein lipase, taken up by adipocytes, and reesteri-
fied into triacylglycerol.
8885d_c21_787-832 2/26/04 9:35 AM Page 806 mac76 mac76:385_reb:
neogenesis was discovered in the 1960s by Lea Reshef,
Richard Hanson, and John Ballard, and simultaneously
by Eleazar Shafrir and his coworkers, who were in-
trigued by the presence of two gluconeogenic enzymes,
pyruvate carboxylase and phosphoenolpyruvate (PEP)
carboxykinase, in adipose tissue, where glucose is not
synthesized. After a long period of inattention, interest
in this pathway has been renewed by the demonstration
of a link between glyceroneogenesis and late-onset
(type 2) diabetes, as we shall see.
Glyceroneogenesis has multiple roles. In adipose tis-
sue, glyceroneogenesis coupled with reesterification of
free fatty acids controls the rate of fatty acid release to
the blood. In brown adipose tissue, the same pathway
may control the rate at which free fatty acids are deliv-
ered to mitochondria for use in thermogenesis (see Fig.
19–30). And in fasting humans, glyceroneogenesis in
the liver alone supports the synthesis of enough glyc-
erol 3-phosphate to account for up to 65% of fatty acids
reesterified to triacylglycerol.
Flux through the triacylglycerol cycle between liver
and adipose tissue is controlled to a large degree by the
activity of PEP carboxykinase, which limits the rate of
both gluconeogenesis and glyceroneogenesis. Gluco-
corticoid hormones such as cortisol (a biological steroid
derived from cholesterol; see Fig. 21–46) and dexa-
methasone (a synthetic glucocorticoid) regulate the
levels of PEP carboxykinase reciprocally in the liver and
adipose tissue. Acting through the glucocorticoid re-
ceptor, these steroid hormones increase the expression
of the gene encoding PEP carboxykinase in the liver,
thus increasing gluconeogenesis and glyceroneogenesis
(Fig. 21–22).
Stimulation of glyceroneogenesis leads to an in-
crease in the synthesis of triacylglycerol molecules in
the liver and their release into the blood. At the same
time, glucocorticoids suppress the expression of the
gene encoding PEP carboxykinase in adipose tissue.
This results in a decrease in glyceroneogenesis in adi-
pose tissue; recycling of fatty acids declines as a result,
and more free fatty acids are released into the blood.
Thus glyceroneogenesis is regulated reciprocally in the
liver and adipose tissue, affecting lipid metabolism in
opposite ways: a lower rate of glyceroneogenesis in adi-
pose tissue leads to more fatty acid release (rather than
recycling), whereas a higher rate in the liver leads to
more synthesis and export of triacylglycerols. The net
result is an increase in flux through the triacylglycerol
cycle. When the glucocorticoids are no longer present,
flux through the cycle declines as the expression of PEP
carboxykinase increases in adipose tissue and decreases
in the liver.
The recent attention given to glyceroneogenesis
has arisen in part from the connection between
this pathway and diabetes. High levels of free fatty acids
in the blood interfere with glucose utilization in muscle
and promote the insulin resistance that leads to type 2
diabetes. A new class of drugs called thiazolidine-
diones have been shown to reduce the levels of fatty
acids circulating in the blood and increase sensitivity to
21.2 Biosynthesis of Triacylglycerols 807
Pyruvate
pyruvate carboxylase
Oxaloacetate
PEP carboxykinase
Phosphoenolpyruvate
multistep
Dihydroxyacetone phosphate
glycerol 3-phosphate
dehydrogenase
CH
2
OH
CHOH
CH
2
O
H11002
O
H11002
O
O
P
Glycerol 3-phosphate
Triacylglycerol synthesis
FIGURE 21–21 Glyceroneogenesis. The pathway is essentially an ab-
breviated version of gluconeogenesis, from pyruvate to dihydroxyace-
tone phosphate (DHAP), followed by conversion of DHAP to glycerol
3-phosphate, which is used for the synthesis of triacylglycerol.
C
CH
2
OH
O
Cortisol
O
HC
3
HC
3
OH
HO
C
CH
2
OH
O
Dexamethasone
O
F
HC
3
HC
3
HC
3
OH
HO
CH
3
CH
3
CH
2
NN
N
H11002
O
S
Rosiglitazone (Avandia)
Thiazolidinediones
Pioglitazone (Actos)
N
O
O
N
H11002
S
O
O
O
8885d_c21_787-832 2/26/04 9:35 AM Page 807 mac76 mac76:385_reb:
insulin. Thiazolidinediones bind to and activate a nu-
clear hormone receptor called peroxisome proliferator-
activated receptor H9253 (PPARH9253), leading to the induction
in adipose tissue of PEP carboxykinase (Fig. 21–22); a
higher activity of PEP carboxykinase then leads to in-
creased synthesis of the precursors of glyceroneogene-
sis. The therapeutic effect of thiazolidinediones is thus
due, at least in part, to the increase in glyceroneogen-
esis, which in turn increases the resynthesis of triacyl-
glycerol in adipose tissue and reduces the release of free
fatty acid from adipose tissue into the blood. ■
SUMMARY 21.2 Biosynthesis of Triacylglycerols
■ Triacylglycerols are formed by reaction of two
molecules of fatty acyl–CoA with glycerol
3-phosphate to form phosphatidic acid; this
product is dephosphorylated to a
diacylglycerol, then acylated by a third
molecule of fatty acyl–CoA to yield a
triacylglycerol.
■ The synthesis and degradation of
triacylglycerol are hormonally regulated.
■ Mobilization and recycling of triacylglycerol
molecules results in a triacylglycerol cycle.
Triacylglycerols are resynthesized from free
fatty acids and glycerol 3-phosphate even
during starvation. The dihydroxyacetone
phosphate precursor of glycerol 3-phosphate is
derived from pyruvate via glyceroneogenesis.
21.3 Biosynthesis of
Membrane Phospholipids
In Chapter 10 we introduced two major classes of mem-
brane phospholipids: glycerophospholipids and sphin-
golipids. Many different phospholipid species can be
constructed by combining various fatty acids and polar
head groups with the glycerol or sphingosine backbone
(see Figs 10–8, 10–12). All the biosynthetic pathways
follow a few basic patterns. In general, the assembly of
phospholipids from simple precursors requires (1) syn-
thesis of the backbone molecule (glycerol or sphingo-
sine); (2) attachment of fatty acid(s) to the backbone
through an ester or amide linkage; (3) addition of a hy-
drophilic head group to the backbone through a phos-
phodiester linkage; and, in some cases, (4) alteration or
exchange of the head group to yield the final phospho-
lipid product.
In eukaryotic cells, phospholipid synthesis occurs
primarily on the surfaces of the smooth endoplasmic
reticulum and the mitochondrial inner membrane. Some
newly formed phospholipids remain at the site of syn-
thesis, but most are destined for other cellular locations.
Chapter 21 Lipid Biosynthesis808
FIGURE 21–22 Regulation of glyceroneogenesis. (a) Gluco-
corticoid hormones stimulate glyceroneogenesis and gluco-
neogenesis in the liver, while suppressing glyceroneogenesis in the
adipose tissue (by reciprocal regulation of the gene expressing PEP
carboxykinase (PEPCK) in the two tissues); this increases the flux
through the triacylglycerol cycle. The glycerol freed by the breakdown
of triacylglycerol in adipose tissue is released to the blood and trans-
ported to the liver, where it is primarily converted to glucose, although
some is converted to glycerol 3-phosphate by glycerol kinase.
(b) A class of drugs called thiazolidinediones are now used to
treat type 2 diabetes. In this disease, high levels of free fatty acids in
the blood interfere with glucose utilization in muscle and promote in-
sulin resistance. Thiazolidinediones activate a nuclear receptor called
peroxisome proliferator-activated receptor H9253 (PPARH9253), which induces
the activity of PEP carboxykinase. Therapeutically, thiazolidinediones
increase the rate of glyceroneogenesis, thus increasing the resynthe-
sis of triacylglycerol in adipose tissue and reducing the amount of free
fatty acid in the blood.
Adipose tissue
Glycerol
3-phosphate
Glycerol
3-phosphate
Fuel for
tissues
Glycerol
Triacylglycerol Triacylglycerol
Fatty
acid
Fatty
acid
LiverBlood
Glycerol
glycero-
neogenesis
(a)
(b)
Pyruvate
Pyruvate
DNA
PEPCKPEPCK
Glycerol
3-phosphate
Glycerol
3-phosphate
Fuel for
tissues
Glycerol
Triacylglycerol Triacylglycerol
Fatty
acid
Fatty
acid
Glycerol
glycero-
neogenesis
Pyruvate
DNA
PEPCK
Glucocorticoids
Thiazolidinediones
Lipoprotein
lipase
8885d_c21_808 2/27/04 12:40 PM Page 808 mac116 mac116:
The process by which water-insoluble phospholipids
move from the site of synthesis to the point of their
eventual function is not fully understood, but we con-
clude this section by discussing some mechanisms that
have emerged in recent years.
Cells Have Two Strategies for Attaching Phospholipid
Head Groups
The first steps of glycerophospholipid synthesis are
shared with the pathway to triacylglycerols (Fig. 21–17):
two fatty acyl groups are esterified to C-1 and C-2 of
L-glycerol 3-phosphate to form phosphatidic acid. Com-
monly but not invariably, the fatty acid at C-1 is satu-
rated and that at C-2 is unsaturated. A second route to
phosphatidic acid is the phosphorylation of a diacyl-
glycerol by a specific kinase.
The polar head group of glycerophospholipids is at-
tached through a phosphodiester bond, in which each
of two alcohol hydroxyls (one on the polar head group
and one on C-3 of glycerol) forms an ester with phos-
phoric acid (Fig. 21–23). In the biosynthetic process,
one of the hydroxyls is first activated by attachment of
a nucleotide, cytidine diphosphate (CDP). Cytidine
monophosphate (CMP) is then displaced in a nucle-
ophilic attack by the other hydroxyl (Fig. 21–24). The
CDP is attached either to the diacylglycerol, forming
the activated phosphatidic acid CDP-diacylglycerol
(strategy 1), or to the hydroxyl of the head group (strat-
egy 2). Eukaryotic cells employ both strategies, whereas
prokaryotes use only the first. The central importance
of cytidine nucleotides in lipid biosynthesis was discov-
ered by Eugene P. Kennedy in the early 1960s.
21.3 Biosynthesis of Membrane Phospholipids 809
Eugene P. Kennedy
O
CCH
P
CH
2
R
2
O C
O
R
1
O
O
O
H5008
CH
2
O
Glycerophospholipid
CCH
CH
2
O C
O
R
1
O
CH
2
ODiacylglycerol
O
O
Head
group
R
2
HHO OH
P
O
O
H5008
Head
group
HO
H
2
O H
2
O
phosphodiester
alcohol
Phosphoric
acid
alcohol
FIGURE 21–23 Head-group
attachment. The phospholipid
head group is attached to a
diacylglycerol by a phospho-
diester bond, formed when
phosphoric acid condenses
with two alcohols, eliminating
two molecules of H
2
O.
OH
Head
group
O
CCH
PO
H5008
CH
2
R
2
O C
O
R
1
O
O
CH
2
O
Glycerophospholipid
O
P
O
Rib
Cytosine
H5008
O O
Strategy 2
CMP
HO
Head
group
O
CCH
PO
H5008
CH
2
R
2
O C
O
R
1
O
O
CH
2
O
CDP-diacylglycerol
O
P
Rib
Cytosine
O
H5008
O
Strategy 1
Diacylglycerol
activated with CDP
Head group
activated with CDP
CMP
CCH
CH
2
R
2
O C
O
R
1
O
O
CH
2
O
P
O
O
H5008
O
O
Head
group
1,2-Diacylglycerol
FIGURE 21–24 Two general strategies for forming
the phosphodiester bond of phospholipids. In both
cases, CDP supplies the phosphate group of the
phosphodiester bond.
8885d_c21_809 2/27/04 12:41 PM Page 809 mac116 mac116:
Chapter 21 Lipid Biosynthesis810
OP
O
H11002
CDP-diacylglycerol
O Rib Cytosine
CR
1
O
C
O
R
2
O
CH
O
CCH
CH
2
R
2
C
O
R
1
O
O
CH
2
P
O
H11002
O
CTP
CMP
CCH
CH
2
R
2
OC
O
R
1
O
O
CH
2
O
P
O
H11002
OO
H11002
O
CH
2
O
C
CH
2
R
2
OO
O
O
C
O
OR
1
O
OO
O
CH
2
O
P
O
O
O
NH
3
O
CH
O
O
H11002
CH
2
H11001
Phosphatidylethanolamine
CHOH
CH
2
OP
O
H11002
O
Glycerol 3-
phosphate
cardiolipin
PP
i
C
CH CH
2
R
2
O
C
O
R
1
O
O
CH
2
O
PO
CO
2
O
CH
O
H11002
CH
2
Phosphatidylglycerol
3-phosphate
OH
CH
2
OP
O
H11002
O
O
H11002
PG 3-phosphate
synthase
Glycerol
Cardiolipin
Phosphatidylglycerol
decarboxylase
P
i
H
2
O
PG 3-phosphate
PS
O
C
CHCH
2
R
2
OO
O
O
C
O
OR
1
O
OO
O
CH
2
O
P
O
COO
H11002
O
O
O
NH
3
O
CH
O
O
H11002
CH
2
H11001
Phosphatidylserine
CH
2
O
C
CH CH
2
OH
R
2
O
C
O
R
1
O
CH
2
O
O
CH
CH
2
OH
CH
2
OP
O
H11002
O
Phosphatidylglycerol
CH
2
OOP
O
H11002
O
CH
2
CR
2
C
O
R
1
O
CH
2
O
O
CH
CH
2
O
phosphatase
synthase
(bacterial)
Serine
PS
synthase
CMP
O
O
FIGURE 21–25 Origin of the polar head groups of
phospholipids in E. coli. Initially, a head group (either
serine or glycerol 3-phosphate) is attached via a CDP-
diacylglycerol intermediate (strategy 1 in Fig. 21–24).
For phospholipids other than phosphatidylserine, the
head group is further modified, as shown here. In the
enzyme names, PG represents phosphatidylglycerol;
PS, phosphatidylserine.
8885d_c21_787-832 2/26/04 9:35 AM Page 810 mac76 mac76:385_reb:
21.3 Biosynthesis of Membrane Phospholipids 811
Phospholipid Synthesis in E. coli Employs
CDP-Diacylglycerol
The first strategy for head-group attachment is illus-
trated by the synthesis of phosphatidylserine, phos-
phatidylethanolamine, and phosphatidylglycerol in E.
coli. The diacylglycerol is activated by condensation of
phosphatidic acid with cytidine triphosphate (CTP) to
form CDP-diacylglycerol, with the elimination of pyro-
phosphate (Fig. 21–25). Displacement of CMP through
nucleophilic attack by the hydroxyl group of serine or
by the C-1 hydroxyl of glycerol 3-phosphate yields phos-
phatidylserine or phosphatidylglycerol 3-phosphate,
respectively. The latter is processed further by cleavage
of the phosphate monoester (with release of P
i
) to yield
phosphatidylglycerol.
Phosphatidylserine and phosphatidylglycerol can
serve as precursors of other membrane lipids in bacte-
ria (Fig. 21–25). Decarboxylation of the serine moiety
in phosphatidylserine, catalyzed by phosphatidylserine
decarboxylase, yields phosphatidylethanolamine. In
E. coli, condensation of two molecules of phosphati-
dylglycerol, with elimination of one glycerol, yields
cardiolipin, in which two diacylglycerols are joined
through a common head group.
Eukaryotes Synthesize Anionic Phospholipids from
CDP-Diacylglycerol
In eukaryotes, phosphatidylglycerol, cardiolipin, and the
phosphatidylinositols (all anionic phospholipids; see Fig.
10–8) are synthesized by the same strategy used for
phospholipid synthesis in bacteria. Phosphatidylglycerol
is made exactly as in bacteria. Cardiolipin synthesis in
eukaryotes differs slightly: phosphatidylglycerol con-
denses with CDP-diacylglycerol (Fig. 21–26), not an-
other molecule of phosphatidylglycerol as in E. coli
(Fig. 21–25).
Phosphatidylinositol is synthesized by condensation
of CDP-diacylglycerol with inositol (Fig. 21–26). Spe-
cific phosphatidylinositol kinases then convert phos-
phatidylinositol to its phosphorylated derivatives (see
Fig. 10–17). Phosphatidylinositol and its phosphory-
lated products in the plasma membrane play a central
role in signal transduction in eukaryotes (see Figs 12–8,
12–19).
OP
O
H11002
CDP-diacylglycerol
O Rib Cytosine
CR
1
O
C
O
R
2
O
CH
O
CH
2
O
P
O
H11002
O
Inositol
CMP
PI
synthase
CCH R
2
OC
O
R
1
O
CH
2
O
O
CH
2
CHOH
CH
2
O P
O
Phosphatidylglycerol
CMP
(eukaryotic)
Cardiolipin
C
PO
2H11002
C
O
R
1
O
O
CH
2
O
PO
H
O
H11002
3
.
CH
2
OO P
O
H11002
O
CH
2
CR
2
C
O
R
1
O
CH
2
O
O
CH
O
cardiolipin
synthase
O
CH
O
OH H
H
OH
H
HH OH
OH
Phosphatidylinositol
OH
R
2
OH
CH
2
CH
2
These
groups can
also be
esterified
with
O
H11002
FIGURE 21–26 Synthesis of cardiolipin and phosphatidylinositol in
eukaryotes. These glycerophospholipids are synthesized using strategy
1 in Figure 21-24. Phosphatidylglycerol is synthesized as in bacteria
(see Fig. 21–25). PI represents phosphatidylinositol.
8885d_c21_787-832 2/26/04 9:35 AM Page 811 mac76 mac76:385_reb:
Eukaryotic Pathways to Phosphatidylserine,
Phosphatidylethanolamine, and Phosphatidylcholine
Are Interrelated
Yeast, like bacteria, can produce phosphatidylserine by
condensation of CDP-diacylglycerol and serine, and can
synthesize phosphatidylethanolamine from phosphatidyl-
serine in the reaction catalyzed by phosphatidylserine
decarboxylase (Fig. 21–27). In mammalian cells, an al-
ternative route to phosphatidylserine is a head-group
exchange reaction, in which free serine displaces
ethanolamine. Phosphatidylethanolamine may also be
converted to phosphatidylcholine (lecithin) by the
addition of three methyl groups to its amino group; S-
adenosylmethionine is the methyl group donor (see
Fig. 18–18) for all three methylation reactions.
In mammals, phosphatidylserine is not synthesized
from CDP-diacylglycerol; instead, it is derived from
phosphatidylethanolamine via the head-group exchange
reaction (Fig. 21–27). Synthesis of phosphatidylethan-
olamine and phosphatidylcholine in mammals occurs by
strategy 2 of Figure 21–24: phosphorylation and activa-
tion of the head group, followed by condensation with
Chapter 21 Lipid Biosynthesis812
methyltransferase
Ethanolamine
C
CH
2
R
2
O
C
O
R
1
O
O
CH
2
O
PO
O
(CH
3
)
3
CH
O
H11002
CH
2
C
CH
CO
2
R
2
O
C
O
R
1
O
O
CH
2
O
P COO
H11002
O
O
NH
3
CH
O
H11002
CH
2
H11001
Phosphatidylcholine
Phosphatidylserine
phosphatidyl-
serine
decarboxylase
Serine
3 adoMet
CH
2
3 adoHcy
C
CH
2
R
2
O
C
O
R
1
O
O
CH
2
O
PO
O
NH
3
CH
O
H11002
CH
2
H11001
Phosphatidylethanolamine
CH
2
N
H11001
CH
2
phosphatidyl-
ethanolamine–
serine
transferase
FIGURE 21–27 The “salvage” pathway from phosphatidylserine to
phosphatidylethanolamine and phosphatidylcholine in yeast. Phos-
phatidylserine and phosphatidylethanolamine are interconverted by
a reversible head-group exchange reaction. In mammals, phos-
phatidylserine is derived from phosphatidylethanolamine by a re-
versal of this reaction; adoMet is S-adenosylmethionine; adoHcy, S-
adenosylhomocysteine.
CMP
CDP-choline–
diacylglycerol
phosphocholine
transferase
CDP-choline
CytosineOP
Diacylglycerol
Rib
CholineCH
2
HO CH
2
H11001
N(CH
3
)
3
O
O
H11002
ADP
choline
kinase
ATP
CH
2
O
H11002
O
P CH
2
H11001
N(CH
3
)
3
O
H11002
O
CH
2
OP CH
2
H11001
N(CH
3
)
3
O
H11002
O
O
PP
i
CTP-choline
cytidylyl
transferase
Phosphocholine
CTP
Phosphatidylcholine
CH
2
R
2
C
O
CH O
O
O
H11002
R
1
CCH
2
O
O
CH
2
POO CH
2
H11001
N(CH
3
)
3
FIGURE 21–28 Pathway for phosphatidylcholine synthesis from
choline in mammals. The same strategy shown here (strategy 2 in Fig.
21–24) is also used for salvaging ethanolamine in phosphatidyle-
thanolamine synthesis.
8885d_c21_787-832 2/26/04 9:35 AM Page 812 mac76 mac76:385_reb:
diacylglycerol. For example, choline is reused (“sal-
vaged”) by being phosphorylated then converted to
CDP-choline by condensation with CTP. A diacylgly-
cerol displaces CMP from CDP-choline, producing
phosphatidylcholine (Fig. 21–28). An analogous sal-
vage pathway converts ethanolamine obtained in the
diet to phosphatidylethanolamine. In the liver, phos-
phatidylcholine is also produced by methylation of
phosphatidylethanolamine (with S-adenosylmethionine,
as described above), but in all other tissues phos-
phatidylcholine is produced only by condensation of
diacylglycerol and CDP-choline. The pathways to
phosphatidylcholine and phosphatidylethanolamine in
various organisms are summarized in Figure 21–29.
Although the role of lipid composition in membrane
function is not entirely understood, changes in compo-
sition can produce dramatic effects. Researchers have
isolated fruit flies with mutations in the gene that en-
codes ethanolamine kinase (analogous to choline kinase;
Fig. 21–28). Lack of this enzyme eliminates one path-
way for phosphatidylethanolamine synthesis, thereby
reducing the amount of this lipid in cellular membranes.
Flies with this mutation—those with the genotype eas-
ily shocked—exhibit transient paralysis following elec-
trical stimulation or mechanical shock that would not
affect wild-type flies.
Plasmalogen Synthesis Requires Formation of an
Ether-Linked Fatty Alcohol
The biosynthetic pathway to ether lipids, including
plasmalogens and the platelet-activating factor
(see Fig. 10–9), involves the displacement of an esteri-
fied fatty acyl group by a long-chain alcohol to form the
ether linkage (Fig. 21–30). Head-group attachment fol-
lows, by mechanisms essentially like those used in for-
mation of the common ester-linked phospholipids. Fi-
nally, the characteristic double bond of plasmalogens
(shaded blue in Fig. 21–30) is introduced by the action
of a mixed-function oxidase similar to that responsible
for desaturation of fatty acids (Fig. 21–13). The perox-
isome is the primary site of plasmalogen synthesis.
Sphingolipid and Glycerophospholipid Synthesis
Share Precursors and Some Mechanisms
The biosynthesis of sphingolipids takes place in four
stages: (1) synthesis of the 18-carbon amine sphinga-
nine from palmitoyl-CoA and serine; (2) attachment of
a fatty acid in amide linkage to yield N-acylsphinga-
nine; (3) desaturation of the sphinganine moiety to
form N-acylsphingosine (ceramide); and (4) attach-
ment of a head group to produce a sphingolipid such
as a cerebroside or sphingomyelin (Fig. 21–31). The
pathway shares several features with the pathways
leading to glycerophospholipids: NADPH provides re-
ducing power, and fatty acids enter as their activated
CoA derivatives. In cerebroside formation, sugars enter
as their activated nucleotide derivatives. Head-group
attachment in sphingolipid synthesis has several novel
aspects. Phosphatidylcholine, rather than CDP-choline,
serves as the donor of phosphocholine in the synthesis
of sphingomyelin.
In glycolipids, the cerebrosides and gangliosides
(see Fig. 10–12), the head-group sugar is attached di-
rectly to the C-1 hydroxyl of sphingosine in glycosidic
linkage rather than through a phosphodiester bond. The
sugar donor is a UDP-sugar (UDP-glucose or UDP-
galactose).
21.3 Biosynthesis of Membrane Phospholipids 813
Choline
3 adoHcy
Mammals
3 adoMet
Serine
Ethanolamine
Phosphatidyl-
serine
Bacteria
and yeast
CDP-diacyl-
glycerol
Serine
CMP
Phosphatidyl-
choline
CDP-
ethanolamine
Ethanolamine
CMP CMP
CDP-choline
CO
2
decarboxylation
Diacylglycerol
Phosphatidyl-
ethanolamine
FIGURE 21–29 Summary of the pathways to phosphatidylcholine and
phosphatidylethanolamine. Conversion of phosphatidylethanolamine
to phosphatidylcholine in mammals takes place only in the liver.
8885d_c21_813 2/27/04 12:41 PM Page 813 mac116 mac116:
H11001
O
CH
2
OH
C
Dihydroxyacetone
phosphate
O
H11002
PO
O
H11002
O
OC
O
C
O
S-CoA
CoA-SH
R
2
COO
H11002
Fatty acyl–CoA
long-chain alcohol
R
1
CH
2
CH
2
R
1
O
PO
O
H11002
O
CH
2
C
R
1
CO
O
H11002
CH
2
fatty acyl
group
1-Acyldihydroxyacetone
3-phosphate
CH
2
OCH
2
R
2
O
S-CoA
CH
2
Saturated fatty alcohol
OHCH
2
R
2
C
PO
O
H11002
O
CH
2
O
H11002
R
2
PO
O
H11002
O
CH
2
CH
2
O
H11002
OCH
2
CH
2
CHOH
NAD
H11001
mixed-function
oxidase
1-Alkylglycerol 3-phosphate
A plasmalogen
CR
3
R
2
O
CH
2
O
O
CH
CH
2
CH
2
CH
2
O PO
H11002
O
O
H11002
Ethanolamine
head- group
attachment
C
CH
2
R
3
O
R
2
O
O
CH
O
PO
O
O
2
CH
O
H11002
CH
2
H11001
CH
2
CH
2
CH
NADP
H11001
R
3
C
O
S-CoA
CoA-SH
NH
3
2H
2
O
NADH H11001 H
H11001
C
CH
2
R
3
O
R
2
O
O
CH
2
O
PO
O
CH
O
H11002
CH
2
H11001
CH
2
CH
2
CH
2
NH
3
H11001 H
H11001
NADPH
CH
2
1-alkyldihydroxy-
acetone
3-phosphate
synthase
1-alkyldihydroxy-
acetone
3-phosphate reductase
1-alkylglycerol
3-phosphate
acyl transferase
1-Alkyl-2-acylglycerol
3-phosphate
1-Alkyldihydroxyacetone
3-phosphate
2NADP
H11001
2 CoA-SH
H11001
2H
H11001
NADPH
Polar Lipids Are Targeted to Specific
Cellular Membranes
After synthesis on the smooth ER, the polar lipids, in-
cluding the glycerophospholipids, sphingolipids, and
glycolipids, are inserted into specific cellular mem-
branes in specific proportions, by mechanisms not yet
understood. Membrane lipids are insoluble in water, so
they cannot simply diffuse from their point of synthesis
(the ER) to their point of insertion. Instead, they are
delivered in membrane vesicles that bud from the Golgi
complex then move to and fuse with the target mem-
brane (see Fig. 11–23). Cytosolic proteins also bind
phospholipids and sterols and transport them between
cellular membranes. These mechanisms contribute to
the establishment of the characteristic lipid composi-
tions of organelle membranes (see Fig. 11–2).
Chapter 21 Lipid Biosynthesis814
FIGURE 21–30 Synthesis of ether lipids and plasmalogens. The newly
formed ether linkage is shaded pink. The intermediate 1-alkyl-2-acyl-
glycerol 3-phosphate is the ether analog of phosphatidic acid.
Mechanisms for attaching head groups to ether lipids are essentially
the same as for their ester-linked analogs. The characteristic double
bond of plasmalogens (shaded blue) is introduced in a final step by a
mixed-function oxidase system similar to that shown in Figure 21–13.
8885d_c21_787-832 2/26/04 9:35 AM Page 814 mac76 mac76:385_reb:
HO
O
H11002
H
CH
3
C
O
O
C
CH
2
CoA-S
(CH
2
)
14
Palmitoyl-CoA
Serine
CoA-SH, CO
2
H9252-Ketosphinganine
NADP
H11001
Sphinganine
Fatty acyl–CoA
CoA-SH
N-acylsphinganine
O
CH
2
C
CH
2
OH
C
CH
HO
C
OH
N(CH
3
)
3
H
2
3
1
NADPH H11001 H
H11001
HO
H
3
NH
H11001
C
CH
2
CH
OH
NH C
O
R
mixed-function
oxidase
(animals)
CHCH CH
C
NH
R
UDP-Glc UDP
CH
2
HO
H
OH
NH C
O
CHCH CH
CR
CH
2
CerebrosideCeramide, containing
sphingosine
O Glc
Phosphatidylcholine
Diacylglycerol
head- group
attachment
Sphingomyelin
CH
2
HO
HNH C
O CHCH CH
CR
O
P
O
O
H11001
CH
2
CH
3
(CH
2
)
12
HH
3
N
H11001
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
14
CH
3
(CH
2
)
12
CH
3
(CH
2
)
12
SUMMARY 21.3 Biosynthesis of
Membrane Phospholipids
■ Diacylglycerols are the principal precursors of
glycerophospholipids.
■ In bacteria, phosphatidylserine is formed by the
condensation of serine with CDP-diacylglycerol;
decarboxylation of phosphatidylserine
produces phosphatidylethanolamine.
Phosphatidylglycerol is formed by condensation
of CDP-diacylglycerol with glycerol
3-phosphate, followed by removal of the
phosphate in monoester linkage.
■ Yeast pathways for the synthesis of
phosphatidylserine, phosphatidylethanolamine,
and phosphatidylglycerol are similar to those in
bacteria; phosphatidylcholine is formed by
methylation of phosphatidylethanolamine.
■ Mammalian cells have some pathways similar to
those in bacteria, but somewhat different
routes for synthesizing phosphatidylcholine and
phosphatidylethanolamine. The head-group
alcohol (choline or ethanolamine) is activated
as the CDP derivative, then condensed with
diacylglycerol. Phosphatidylserine is derived
only from phosphatidylethanolamine.
■ Synthesis of plasmalogens involves formation of
their characteristic double bond by a
mixed-function oxidase. The head groups of
sphingolipids are attached by unique
mechanisms.
■ Phospholipids travel to their intracellular
destinations via transport vesicles or specific
proteins.
21.3 Biosynthesis of Membrane Phospholipids 815
FIGURE 21–31 Biosynthesis of sphingolipids. Condensation of
palmitoyl-CoA and serine followed by reduction with NADPH
yields sphinganine, which is then acylated to N-acylsphinganine
(a ceramide). In animals, a double bond (shaded pink) is
created by a mixed-function oxidase, before the final addition
of a head group: phosphatidylcholine, to form sphingomyelin;
glucose, to form a cerebroside.
8885d_c21_787-832 2/26/04 9:35 AM Page 815 mac76 mac76:385_reb:
21.4 Biosynthesis of Cholesterol, Steroids,
and Isoprenoids
Cholesterol is doubtless the most publicized lipid, no-
torious because of the strong correlation between high
levels of cholesterol in the blood and the incidence of
human cardiovascular diseases. Less well advertised is
cholesterol’s crucial role as a component of cellular
membranes and as a precursor of steroid hormones and
bile acids. Cholesterol is an essential molecule in many
animals, including humans, but is not required in the
mammalian diet—all cells can synthesize it from simple
precursors.
The structure of this 27-carbon compound suggests
a complex biosynthetic pathway, but all of its carbon
atoms are provided by a single precursor—acetate (Fig.
21–32). The isoprene units that are the essential in-
termediates in the pathway from acetate to cholesterol
are also precursors to many other natural lipids, and the
mechanisms by which isoprene units are polymerized
are similar in all these pathways.
We begin with an account of the main steps in the
biosynthesis of cholesterol from acetate, then discuss
the transport of cholesterol in the blood, its uptake by
cells, the normal regulation of cholesterol synthesis, and
its regulation in those with defects in cholesterol uptake
or transport. We next consider other cellular compo-
nents derived from cholesterol, such as bile acids and
steroid hormones. Finally, an outline of the biosynthetic
pathways to some of the many compounds derived from
isoprene units, which share early steps with the path-
way to cholesterol, illustrates the extraordinary versa-
tility of isoprenoid condensations in biosynthesis.
CCH
2
CHCH
2
CH
3
Isoprene
Cholesterol Is Made from Acetyl-CoA in Four Stages
Cholesterol, like long-chain fatty acids, is made from
acetyl-CoA, but the assembly plan is quite different. In
early experiments, animals were fed acetate labeled
with
14
C in either the methyl carbon or the carboxyl car-
bon. The pattern of labeling in the cholesterol isolated
from the two groups of animals (Fig. 21–32) provided
the blueprint for working out the enzymatic steps in cho-
lesterol biosynthesis.
Synthesis takes place in four stages, as shown in
Figure 21–33: 1 condensation of three acetate units to
form a six-carbon intermediate, mevalonate; 2 con-
version of mevalonate to activated isoprene units; 3
polymerization of six 5-carbon isoprene units to form
the 30-carbon linear squalene; and 4 cyclization of
squalene to form the four rings of the steroid nucleus,
with a further series of changes (oxidations, removal or
migration of methyl groups) to produce cholesterol.
Chapter 21 Lipid Biosynthesis816
COO
H11002
CH
3
3
10
Acetate
AB
C
D
HO
Cholesterol
C
C
C
C
CC
C
C
C
C
C
C
C
C
C
C
CC
C
C
CC
C
C
C
C
C
9
2
4
5
6
7
1
25
19
11
8
20
21
12
18
13
14 15
16
17
22 24 26
23
27
FIGURE 21–32 Origin of the carbon atoms of cholesterol. This can
be deduced from tracer experiments with acetate labeled in the methyl
carbon (black) or the carboxyl carbon (red). The individual rings in
the fused-ring system are designated A through D.
FIGURE 21–33 Summary of cholesterol biosynthesis. The four stages
are discussed in the text. Isoprene units in squalene are set off by red
dashed lines.
COO
H5008
3 CH
3
C
CH
3
H5008
OOC
Acetate
O
H5008
CH
2
CH
2
HO
Mevalonate
C
CH
2
CH
2
OH
CH
3
4
1
CH
2
2
CH
2
O
PO P
O
O
H5008
O
O
H5008
isoprene
Activated isoprene
2
3
Squalene
Cholesterol
OH
8885d_c21_816 2/27/04 12:41 PM Page 816 mac116 mac116:
FIGURE 21–34 Formation of mevalonate from acetyl-CoA. The ori-
gin of C-1 and C-2 of mevalonate from acetyl-CoA is shown in pink.
CH
2
OH
2NADP
H11001
2 CH
3
C
O
CH
3
C
O
S-CoA
Acetyl-CoA
thiolase
CoA-SH
Acetoacetyl-CoA
HMG-CoA
reductase
CH
3
C
O
CH
2
C
O S-CoA
CoA-SH
CH
2
C
O
S-CoA
CH
2
CH
3
S-CoA
H9252-Hydroxy-H9252-methylglutaryl-CoA
(HMG-CoA)
COO
H5008
CoA-SH
H11001 2H
H11001
CH
2
C
CH
2
C
COO
H5008
OH
OHCH
3
Mevalonate
HMG-CoA
synthase
3
1
5
2
4
NADPH2
Stage 1 Synthesis of Mevalonate from Acetate The first
stage in cholesterol biosynthesis leads to the interme-
diate mevalonate (Fig. 21–34). Two molecules of
acetyl-CoA condense to form acetoacetyl-CoA, which
condenses with a third molecule of acetyl-CoA to yield
the six-carbon compound H9252-hydroxy-H9252-methylglu-
taryl-CoA (HMG-CoA). These first two reactions are
catalyzed by thiolase and HMG-CoA synthase, re-
spectively. The cytosolic HMG-CoA synthase in this
pathway is distinct from the mitochondrial isozyme that
catalyzes HMG-CoA synthesis in ketone body formation
(see Fig. 17–18).
The third reaction is the committed and rate-limiting
step: reduction of HMG-CoA to mevalonate, for which
each of two molecules of NADPH donates two electrons.
HMG-CoA reductase, an integral membrane protein of
the smooth ER, is the major point of regulation on the
pathway to cholesterol, as we shall see.
Stage 2 Conversion of Mevalonate to Two Activated Isoprenes
In the next stage of cholesterol synthesis, three phos-
phate groups are transferred from three ATP molecules
to mevalonate (Fig. 21–35). The phosphate attached to
the C-3 hydroxyl group of mevalonate in the interme-
diate 3-phospho-5-pyrophosphomevalonate is a good
21.4 Biosynthesis of Cholesterol, Steroids, and Isoprenoids 817
H5008
OOC
CH
3
C O
3-Phospho-5-
pyrophosphomevalonate
H9004
3
-Isopentenyl pyrophosphate
Dimethylallyl pyrophosphate
Activated
isoprenes
CH
2
5
CH
3
CH
2
O
H5008
OOC
CH
3
C
Mevalonate
ADP
ATP
CH
2
CH
2
CH
2
OH
OH
H5008
OOC
CH
3
C O
5-Phosphomevalonate
ADP
ATP
CH
2
O
H5008
CH
2
CH
2
OH
O
PO
H5008
H5008
OOC
CH
3
C
O
ADP
ATP
CH
2
O
H5008
CH
2
CH
2
OH
O
P
O
H5008
O
O
H5008
O
P
O
5-Pyrophosphomevalonate
O
H5008
O
PO
H5008
O
H5008
O
P
P
O
H5008
O
O
H5008
CO
2
, P
i
CH
3
C OCH
2
CH
2
CH
2
O
O
H5008
O
PP
O
H5008
O
O
H5008
CH
3
C OCH
2
CH
2
O
O
H5008
O
PP
O
H5008
O
O
H5008
CH
3142
mevalonate
5-phosphotransferase
phosphomevalonate
kinase
pyrophospho-
mevalonate
decarboxylase
pyrophospho-
mevalonate
decarboxylase
FIGURE 21–35 Conversion of mevalonate to activated isoprene
units. Six of these activated units combine to form squalene (see Fig.
21–36). The leaving groups of 3-phospho-5-pyrophosphomevalonate
are shaded pink. The bracketed intermediate is hypothetical.
8885d_c21_817 2/27/04 1:45 PM Page 817 Mac113 mac113:
leaving group; in the next step, both this phosphate and
the nearby carboxyl group leave, producing a double
bond in the five-carbon product, H9004
3
-isopentenyl
pyrophosphate. This is the first of the two activated
isoprenes central to cholesterol formation. Isomerization
of H9004
3
-isopentenyl pyrophosphate yields the second acti-
vated isoprene, dimethylallyl pyrophosphate. Synthe-
sis of isopentenyl pyrophosphate in the cytoplasm of
plant cells follows the pathway described here. However,
plant chloroplasts and many bacteria use a mevalonate-
independent pathway. This alternative pathway does not
occur in animals, so it is an attractive target for the
development of new antibiotics.
Stage 3 Condensation of Six Activated Isoprene Units to Form
Squalene Isopentenyl pyrophosphate and dimethylallyl
pyrophosphate now undergo a head-to-tail condensa-
tion, in which one pyrophosphate group is displaced and
a 10-carbon chain, geranyl pyrophosphate, is formed
(Fig. 21–36). (The “head” is the end to which pyrophos-
phate is joined.) Geranyl pyrophosphate undergoes an-
other head-to-tail condensation with isopentenyl pyro-
Chapter 21 Lipid Biosynthesis818
H9004
3
-Isopentenyl pyrophosphate
H5008
O O
O
H5008
O
PP
O
H5008
O
O
H11001
Dimethylallyl pyrophosphate
O O
O
H5008
O
PP
O
H5008
O
O
H5008
prenyl transferase
(head-to-tail)
O O
O
H5008
O
PP
O
H5008
O
O
H5008
Farnesyl pyrophosphate
O O
O
H5008
O
PP
O
H5008
O
O
H5008
PP
i
prenyl transferase
(head-to-tail
condensation)
Geranyl pyrophosphateO O
O
H5008
O
PP
O
H5008
O
O
H5008
H9004
3
-Isopentenyl pyrophosphate
O O
O
H5008
O
PP
O
H5008
O
O
H5008
Farnesyl pyrophosphate
NADPH H11001 H
H11001
NADP
H11001
2 PP
i
squalene synthase
(head-to-head)
Squalene
PP
i
FIGURE 21–36 Formation of squalene. This 30-carbon structure arises
through successive condensations of activated isoprene (five-carbon)
units.
8885d_c21_787-832 2/26/04 9:35 AM Page 818 mac76 mac76:385_reb:
phosphate, yielding the 15-carbon intermediate farne-
syl pyrophosphate. Finally, two molecules of farnesyl
pyrophosphate join head to head, with the elimination
of both pyrophosphate groups, to form squalene.
The common names of these intermediates derive
from the sources from which they were first isolated.
Geraniol, a component of rose oil, has the aroma of gera-
niums, and farnesol is an aromatic compound found in
the flowers of the Farnese acacia tree. Many natural
scents of plant origin are synthesized from isoprene
units. Squalene, first isolated from the liver of sharks
(genus Squalus), has 30 carbons, 24 in the main chain
and 6 in the form of methyl group branches.
Stage 4 Conversion of Squalene to the Four-Ring Steroid Nu-
cleus When the squalene molecule is represented as in
Figure 21–37, the relationship of its linear structure to
the cyclic structure of the sterols becomes apparent. All
21.4 Biosynthesis of Cholesterol, Steroids, and Isoprenoids 819
Squalene
Cholesterol
squalene
monooxygenase
O
2
C
2
H
5
NADP
H11001
Squalene 2,3-epoxide
multistep
(plants)
cyclase
Stigmasterol
O
cyclase
(animals)
Ergosterol
multistep
(fungi)
Lanosterol
multistep
NADPH H11001 H
H11001
HO
H
2
O
HO
HO
HO
HO
3
2
H11001
FIGURE 21–37 Ring closure converts
linear squalene to the condensed steroid
nucleus. The first step in this sequence is
catalyzed by a mixed-function oxidase
(a monooxygenase), for which the co-
substrate is NADPH. The product is an
epoxide, which in the next step is
cyclized to the steroid nucleus. The final
product of these reactions in animal cells
is cholesterol; in other organisms, slightly
different sterols are produced, as shown.
8885d_c21_787-832 2/26/04 9:35 AM Page 819 mac76 mac76:385_reb:
sterols have the four fused rings that form the steroid
nucleus, and all are alcohols, with a hydroxyl group at
C-3—thus the name “sterol.” The action of squalene
monooxygenase adds one oxygen atom from O
2
to
the end of the squalene chain, forming an epoxide.
This enzyme is another mixed-function oxidase (Box
21–1); NADPH reduces the other oxygen atom of O
2
to H
2
O. The double bonds of the product, squalene
2,3-epoxide, are positioned so that a remarkable con-
certed reaction can convert the linear squalene epox-
ide to a cyclic structure. In animal cells, this cycliza-
tion results in the formation of lanosterol, which
contains the four rings characteristic of the steroid
nucleus. Lanosterol is finally converted to cholesterol
in a series of about 20 reactions that include the
migration of some methyl groups and the removal of
others. Elucidation of this extraordinary biosynthetic
pathway, one of the most complex known, was ac-
complished by Konrad Bloch, Feodor Lynen, John
Cornforth, and George Popják in the late 1950s.
Cholesterol is the sterol characteristic of animal
cells; plants, fungi, and protists make other, closely re-
lated sterols instead. They use the same synthetic path-
way as far as squalene 2,3-epoxide, at which point the
pathways diverge slightly, yielding other sterols, such as
stigmasterol in many plants and ergosterol in fungi (Fig.
21–37).
Cholesterol Has Several Fates
Much of the cholesterol synthesis in vertebrates takes
place in the liver. A small fraction of the cholesterol
made there is incorporated into the membranes of he-
patocytes, but most of it is exported in one of three
forms: biliary cholesterol, bile acids, or cholesteryl es-
ters. Bile acids and their salts are relatively hydrophilic
cholesterol derivatives that are synthesized in the liver
and aid in lipid digestion (see Fig. 17–1). Cholesteryl
esters are formed in the liver through the action of
acyl-CoA–cholesterol acyl transferase (ACAT).
This enzyme catalyzes the transfer of a fatty acid from
coenzyme A to the hydroxyl group of cholesterol (Fig.
21–38), converting the cholesterol to a more hy-
drophobic form. Cholesteryl esters are transported in
secreted lipoprotein particles to other tissues that use
cholesterol, or they are stored in the liver.
All growing animal tissues need cholesterol for
membrane synthesis, and some organs (adrenal gland
and gonads, for example) use cholesterol as a precur-
sor for steroid hormone production (discussed below).
Cholesterol is also a precursor of vitamin D (see Fig.
10–20a).
Cholesterol and Other Lipids Are
Carried on Plasma Lipoproteins
Cholesterol and cholesteryl esters, like triacylglycerols
and phospholipids, are essentially insoluble in water,
yet must be moved from the tissue of origin to the tis-
sues in which they will be stored or consumed. They are
carried in the blood plasma as plasma lipoproteins,
Chapter 21 Lipid Biosynthesis820
Konrad Bloch, Feodor Lynen,
1912–2000 1911–1979
John Cornforth George Popják
Cholesteryl ester
O
Fatty acyl–CoA
CoA-SH
acyl-CoA–cholesterol
acyl transferase
(ACAT)
Cholesterol
HO
O
CR
FIGURE 21–38 Synthesis of cholesteryl esters. Esterification con-
verts cholesterol to an even more hydrophobic form for storage and
transport.
8885d_c21_787-832 2/26/04 9:35 AM Page 820 mac76 mac76:385_reb:
macromolecular complexes of specific carrier proteins,
apolipoproteins, with various combinations of phos-
pholipids, cholesterol, cholesteryl esters, and triacyl-
glycerols.
Apolipoproteins (“apo” designates the protein in its
lipid-free form) combine with lipids to form several
classes of lipoprotein particles, spherical complexes
with hydrophobic lipids in the core and hydrophilic
amino acid side chains at the surface (Fig. 21–39a). Dif-
ferent combinations of lipids and proteins produce par-
ticles of different densities, ranging from chylomicrons
to high-density lipoproteins. These particles can be sep-
arated by ultracentrifugation (Table 21–2) and visual-
ized by electron microscopy (Fig. 21–39b).
Each class of lipoprotein has a specific function, de-
termined by its point of synthesis, lipid composition, and
apolipoprotein content. At least nine different apolipo-
proteins are found in the lipoproteins of human plasma
(Table 21–3), distinguishable by their size, their reac-
tions with specific antibodies, and their characteristic
distribution in the lipoprotein classes. These protein
components act as signals, targeting lipoproteins to spe-
cific tissues or activating enzymes that act on the
lipoproteins.
Chylomicrons, discussed in Chapter 17 in con-
nection with the movement of dietary triacylglycerols
from the intestine to other tissues, are the largest of the
lipoproteins and the least dense, containing a high
21.4 Biosynthesis of Cholesterol, Steroids, and Isoprenoids 821
Phospholipid
monolayer
Triacylglycerols
Cholesteryl esters
Free (unesterified)
cholesterol
ApoB-100
(a)
Chylomicrons (H1100360,000) VLDL (H11003180,000)
LDL (H11003180,000)
HDL (H11003180,000)
(b)
FIGURE 21–39 Lipoproteins. (a) Structure of a low-density lipopro-
tein (LDL). Apolipoprotein B-100 (apoB-100) is one of the largest sin-
gle polypeptide chains known, with 4,636 amino acid residues (M
r
513,000). (b) Four classes of lipoproteins, visualized in the electron
microscope after negative staining. Clockwise from top left: chylomi-
crons, 50 to 200 nm in diameter; VLDL, 28 to 70 nm; HDL, 8 to
11 nm; and LDL, 20 to 25 nm. For properties of lipoproteins, see Table
21–2.
TABLE 21–2 Major Classes of Human Plasma Lipoproteins: Some Properties
Composition (wt %)
Lipoprotein Density (g/mL) Protein Phospholipids Free cholesterol Cholesteryl esters Triacylglycerols
Chylomicrons H110211.006 2 9 1 3 85
VLDL 0.95–1.006 10 18 7 12 50
LDL 1.006–1.063 23 20 8 37 10
HDL 1.063–1.210 55 24 2 15 4
Source: Modified from Kritchevsky, D. (1986) Atherosclerosis and nutrition. Nutr. Int. 2, 290–297.
8885d_c21_787-832 2/26/04 9:35 AM Page 821 mac76 mac76:385_reb:
proportion of triacylglycerols (see Fig. 17–2). Chylomi-
crons are synthesized in the ER of epithelial cells that
line the small intestine, then move through the lymphatic
system and enter the bloodstream via the left subcla-
vian vein. The apolipoproteins of chylomicrons include
apoB-48 (unique to this class of lipoproteins), apoE, and
apoC-II (Table 21–3). ApoC-II activates lipoprotein lipase
in the capillaries of adipose, heart, skeletal muscle, and
lactating mammary tissues, allowing the release of free
fatty acids to these tissues. Chylomicrons thus carry di-
etary fatty acids to tissues where they will be consumed
or stored as fuel (Fig. 21–40). The remnants of chylo-
microns (depleted of most of their triacylglycerols but
still containing cholesterol, apoE, and apoB-48) move
through the bloodstream to the liver. Receptors in the
liver bind to the apoE in the chylomicron remnants and
mediate their uptake by endocytosis. In the liver, the
remnants release their cholesterol and are degraded in
lysosomes.
When the diet contains more fatty acids than are
needed immediately as fuel, they are converted to tria-
cylglycerols in the liver and packaged with specific
apolipoproteins into very-low-density lipoprotein
(VLDL). Excess carbohydrate in the diet can also be
converted to triacylglycerols in the liver and exported
as VLDLs (Fig. 21–40a). In addition to triacylglycerols,
VLDLs contain some cholesterol and cholesteryl esters,
as well as apoB-100, apoC-I, apoC-II, apoC-III, and apo-
E (Table 21–3). These lipoproteins are transported in
the blood from the liver to muscle and adipose tissue,
where activation of lipoprotein lipase by apoC-II causes
the release of free fatty acids from the VLDL triacyl-
glycerols. Adipocytes take up these fatty acids, recon-
vert them to triacylglycerols, and store the products in
intracellular lipid droplets; myocytes, in contrast, pri-
marily oxidize the fatty acids to supply energy. Most
VLDL remnants are removed from the circulation by
hepatocytes. The uptake, like that for chylomicrons, is
Chapter 21 Lipid Biosynthesis822
Intestine
Chylomicrons
Free fatty acids
Capillary
lipoprotein lipase
Mammary, muscle, or adipose tissue
Blood plasma
after fast
(a) (b)
HDL precursors
(from liver and
intestine)
Extrahepatic
tissues
Reverse
cholesterol
transport
HDL
LDL
Liver
VLDL
Chylomicron
remnants
VLDL
remnants
(IDL)
Blood plasma
after meal
FIGURE 21–40 Lipoproteins and lipid transport. (a) Lipids are
transported in the bloodstream as lipoproteins, which exist as sev-
eral variants that have different functions, different protein and lipid
compositions (see Tables 21–2, 21–3), and thus different densities.
Dietary lipids are packaged into chylomicrons; much of their tria-
cylglycerol content is released by lipoprotein lipase to adipose and
muscle tissues during transport through capillaries. Chylomicron
remnants (containing largely protein and cholesterol) are taken up
by the liver. Endogenous lipids and cholesterol from the liver are
delivered to adipose and muscle tissue by VLDL. Extraction of lipid
from VLDL (along with loss of some apolipoproteins) gradually con-
verts some of it to LDL, which delivers cholesterol to extrahepatic
tissues or returns to the liver. The liver takes up LDL, VLDL rem-
nants, and chylomicron remnants by receptor-mediated endocyto-
sis. Excess cholesterol in extrahepatic tissues is transported back to
the liver as HDL. In the liver, some cholesterol is converted to bile
salts.
(b) Blood plasma samples collected after a fast (left) and after a
high-fat meal (right). Chylomicrons produced after a fatty meal give
the plasma a milky appearance.
8885d_c21_787-832 2/26/04 9:35 AM Page 822 mac76 mac76:385_reb:
receptor-mediated and depends on the presence of
apoE in the VLDL remnants (Box 21–3 describes a link
between apoE and Alzheimer’s disease).
The loss of triacylglycerol converts some VLDL to
VLDL remnants (also called intermediate density lipo-
protein, IDL); further removal of triacylglycerol from
VLDL produces low-density lipoprotein (LDL)
(Table 21–2). Very rich in cholesterol and cholesteryl
esters and containing apoB-100 as their major apoli-
poprotein, LDLs carry cholesterol to extrahepatic tis-
sues that have specific plasma membrane receptors that
recognize apoB-100. These receptors mediate the up-
take of cholesterol and cholesteryl esters in a process
described below.
The fourth major lipoprotein type, high-density
lipoprotein (HDL), originates in the liver and small
intestine as small, protein-rich particles that contain rel-
atively little cholesterol and no cholesteryl esters (Fig.
21–40). HDLs contain apoA-I, apoC-I, apoC-II, and other
apolipoproteins (Table 21–3), as well as the enzyme
lecithin-cholesterol acyl transferase (LCAT),
which catalyzes the formation of cholesteryl esters from
lecithin (phosphatidylcholine) and cholesterol (Fig.
21–41). LCAT on the surface of nascent (newly form-
ing) HDL particles converts the cholesterol and phos-
phatidylcholine of chylomicron and VLDL remnants to
cholesteryl esters, which begin to form a core, trans-
forming the disk-shaped nascent HDL to a mature,
spherical HDL particle. This cholesterol-rich lipoprotein
then returns to the liver, where the cholesterol is un-
loaded; some of this cholesterol is converted to bile salts.
21.4 Biosynthesis of Cholesterol, Steroids, and Isoprenoids 823
TABLE 21–3 Apolipoproteins of the Human Plasma Lipoproteins
Apolipoprotein Molecular weight Lipoprotein association Function (if known)
ApoA-I 28,331 HDL Activates LCAT; interacts with ABC transporter
ApoA-II 17,380 HDL
ApoA-IV 44,000 Chylomicrons, HDL
ApoB-48 240,000 Chylomicrons
ApoB-100 513,000 VLDL, LDL Binds to LDL receptor
ApoC-I 7,000 VLDL, HDL
ApoC-II 8,837 Chylomicrons, VLDL, HDL Activates lipoprotein lipase
ApoC-III 8,751 Chylomicrons, VLDL, HDL Inhibits lipoprotein lipase
ApoD 32,500 HDL
ApoE 34,145 Chylomicrons, VLDL, HDL Triggers clearance of VLDL and chylomicron
remnants
Source: Modified from Vance, D.E. & Vance, J.E. (eds) (1985) Biochemistry of Lipids and Membranes. The Benjamin/Cummings Publishing
Company, Menlo Park, CA.
Cholesteryl ester
lecithin-cholesterol
acyl transferase
(LCAT)
Cholesterol
CH
2
P
CH
O
O
CH
2
O
H5008
R
1
O
O
C
O
CH
2
CH
2
N(CH
3
)
3
H11001
H11001
Phosphatidylcholine (lecithin)
HO
CH
2
P
R
2
CH
O O
CH
2
O
H5008
C
R
1
O
O
C
O
O
O
CH
2
CH
2
N(CH
3
)
3
H11001
H11001
OH
Lysolecithin
O
O
CR
2
FIGURE 21–41 Reaction catalyzed by lecithin-cholesterol acyl trans-
ferase (LCAT). This enzyme is present on the surface of HDL and is
stimulated by the HDL component apoA-I. Cholesteryl esters accu-
mulate within nascent HDLs, converting them to mature HDLs.
8885d_c21_787-832 2/26/04 9:35 AM Page 823 mac76 mac76:385_reb:
HDL may be taken up in the liver by receptor-
mediated endocytosis, but at least some of the choles-
terol in HDL is delivered to other tissues by a novel
mechanism. HDL can bind to plasma membrane recep-
tor proteins called SR-BI in hepatic and steroidogenic
tissues such as the adrenal gland. These receptors me-
diate not endocytosis but a partial and selective trans-
fer of cholesterol and other lipids in HDL into the cell.
Depleted HDL then dissociates to recirculate in the
bloodstream and extract more lipids from chylomicron
and VLDL remnants. Depleted HDL can also pick up
cholesterol stored in extrahepatic tissues and carry it to
the liver, in reverse cholesterol transport pathways
(Fig. 21–40). In one reverse transport path, interaction
of nascent HDL with SR-BI receptors in cholesterol-rich
cells triggers passive movement of cholesterol from the
cell surface into HDL, which then carries it back to the
liver. In a second pathway, apoA-I in depleted HDL in-
teracts with an active transporter, the ABC1 protein, in
a cholesterol-rich cell. The apoA-I (and presumably the
HDL) is taken up by endocytosis, then resecreted with
a load of cholesterol, which it transports to the liver.
The ABC1 protein is a member of a large family of
multidrug transporters, sometimes called ABC trans-
porters because they all have ATP-binding cassettes;
they also have two transmembrane domains with six
transmembrane helices (Chapter 11). These proteins
actively transport a variety of ions, amino acids, vita-
mins, steroid hormones, and bile salts across plasma
membranes. The CFTR protein that is defective in cys-
tic fibrosis (see Box 11–3) is another member of this
ABC family of multidrug transporters.
Cholesteryl Esters Enter Cells
by Receptor-Mediated Endocytosis
Each LDL particle in the bloodstream contains apoB-
100, which is recognized by specific surface receptor
proteins, LDL receptors, on cells that need to take up
cholesterol. The binding of LDL to an LDL receptor ini-
tiates endocytosis, which conveys the LDL and its re-
ceptor into the cell within an endosome (Fig. 21–42).
The endosome eventually fuses with a lysosome, which
contains enzymes that hydrolyze the cholesteryl esters,
releasing cholesterol and fatty acid into the cytosol. The
apoB-100 of LDL is also degraded to amino acids that
are released to the cytosol, but the LDL receptor es-
capes degradation and is returned to the cell surface, to
function again in LDL uptake. ApoB-100 is also present
in VLDL, but its receptor-binding domain is not avail-
able for binding to the LDL receptor; conversion of
VLDL to LDL exposes the receptor-binding domain of
apoB-100. This pathway for the transport of cholesterol
in blood and its receptor-mediated endocytosis by
target tissues was elucidated by Michael Brown and
Joseph Goldstein.
Chapter 21 Lipid Biosynthesis824
BOX 21–3 BIOCHEMISTRY IN MEDICINE
ApoE Alleles Predict Incidence of
Alzheimer’s Disease
In the human population there are three common vari-
ants, or alleles, of the gene encoding apolipoprotein
E. The most common, accounting for about 78% of
human apoE alleles, is APOE3; alleles APOE4 and
APOE2 account for 15% and 7%, respectively. The
APOE4 allele is particularly common in humans with
Alzheimer’s disease, and the link is highly predictive.
Individuals who inherit APOE4 have an increased risk
of late-onset Alzheimer’s disease. Those who are ho-
mozygous for APOE4 have a 16-fold increased risk of
developing the disease; for those who do, the mean
age of onset is just under 70 years. For people who
inherit two copies of APOE3, by contrast, the mean
age of onset of Alzheimer’s disease exceeds 90 years.
The molecular basis for the association between
apoE4 and Alzheimer’s disease is not yet known. Spec-
ulation has focused on a possible role for apoE in sta-
bilizing the cytoskeletal structure of neurons. The
apoE2 and apoE3 proteins bind to a number of pro-
teins associated with neuronal microtubules, whereas
apoE4 does not. This may accelerate the death of neu-
rons. Whatever the mechanism proves to be, these ob-
servations promise to expand our understanding of
the biological functions of apolipoproteins.
Michael Brown and Joseph Goldstein
8885d_c21_787-832 2/26/04 9:35 AM Page 824 mac76 mac76:385_reb:
Cholesterol that enters cells by this path may be in-
corporated into membranes or reesterified by ACAT
(Fig. 21–38) for storage within cytosolic lipid droplets.
Accumulation of excess intracellular cholesterol is pre-
vented by reducing the rate of cholesterol synthesis
when sufficient cholesterol is available from LDL in the
blood.
The LDL receptor also binds to apoE and plays a
significant role in the hepatic uptake of chylomicrons
and VLDL remnants. However, if LDL receptors are un-
available (as, for example, in a mouse strain that lacks
the gene for the LDL receptor), VLDL remnants and
chylomicrons are still taken up by the liver even though
LDL is not. This indicates the presence of a back-up sys-
tem for receptor-mediated endocytosis of VLDL rem-
nants and chylomicrons. One back-up receptor is
lipoprotein receptor-related protein (LRP), which binds
to apoE as well as to a number of other ligands.
Cholesterol Biosynthesis Is Regulated
at Several Levels
Cholesterol synthesis is a complex and energy-
expensive process, so it is clearly advantageous to an
organism to regulate the biosynthesis of cholesterol to
complement dietary intake. In mammals, cholesterol
production is regulated by intracellular cholesterol con-
centration and by the hormones glucagon and insulin.
The rate-limiting step in the pathway to cholesterol (and
a major site of regulation) is the conversion of HMG-
CoA to mevalonate (Fig. 21–34), the reaction catalyzed
by HMG-CoA reductase.
Regulation in response to cholesterol levels is me-
diated by an elegant system of transcriptional regulation
of the gene encoding HMG-CoA reductase. This gene,
along with more than 20 other genes encoding enzymes
that mediate the uptake and synthesis of cholesterol and
21.4 Biosynthesis of Cholesterol, Steroids, and Isoprenoids 825
LDL particle
ApoB-100
Cholesteryl ester
LDL receptor
Golgi
complex
LDL receptor
synthesis
Endoplasmic
reticulum
Nucleus
Cholesterol
Cholesteryl
ester droplet
Fatty
acids
Amino
acids
Lysosome
Endosome
receptor-mediated
endocytosis
FIGURE 21–42 Uptake of cholesterol by receptor-mediated endocytosis.
8885d_c21_787-832 2/26/04 9:35 AM Page 825 mac76 mac76:385_reb:
unsaturated fatty acids, is controlled by a small family
of proteins called sterol regulatory element-binding
proteins (SREBPs). When newly synthesized, these pro-
teins are embedded in the ER. Only the soluble amino-
terminal domain of an SREBP functions as a transcrip-
tional activator, using mechanisms discussed in Chapter
28. However, this domain has no access to the nucleus
and cannot participate in gene activation while it re-
mains part of the SREBP molecule. To activate tran-
scription of the HMG-CoA reductase gene and other
genes, the transcriptionally active domain is separated
from the rest of the SREBP by proteolytic cleavage.
When cholesterol levels are high, SREBPs are inactive,
secured to the ER in a complex with another protein
called SREBP cleavage-activating protein (SCAP) (Fig.
21–43). It is SCAP that binds cholesterol and a number
of other sterols, thus acting as a sterol sensor. When
sterol levels are high, the SCAP-SREBP complex prob-
ably interacts with another protein that retains the en-
tire complex in the ER. When the level of sterols in the
cell declines, a conformational change in SCAP causes
release of the SCAP-SREBP complex from the ER-
retention activity, and the complex migrates within vesi-
cles to the Golgi complex. In the Golgi complex, SREBP
is cleaved twice by two different proteases, the second
cleavage releasing the amino-terminal domain into the
cytosol. This domain travels to the nucleus and activates
transcription of its target genes. The amino-terminal
domain of SREBP has a short half-life and is rapidly
degraded by proteasomes (see Fig. 27–42). When sterol
levels increase sufficiently, the proteolytic release of
SREBP amino-terminal domains is again blocked, and
proteasome degradation of the existing active domains
results in a rapid shut-down of the gene targets.
Several other mechanisms also regulate cholesterol
synthesis (Fig. 21–44). Hormonal control is mediated
by covalent modification of HMG-CoA reductase itself.
The enzyme exists in phosphorylated (inactive) and
dephosphorylated (active) forms. Glucagon stimulates
phosphorylation (inactivation), and insulin promotes
dephosphorylation, activating the enzyme and favoring
cholesterol synthesis. High intracellular concentrations
of cholesterol activate ACAT, which increases esterifi-
cation of cholesterol for storage. Finally, a high cellular
cholesterol level diminishes transcription of the gene
that encodes the LDL receptor, reducing production of
the receptor and thus the uptake of cholesterol from
the blood.
Chapter 21 Lipid Biosynthesis826
Cytosol
Golgi
complex
Golgi
complex
SCAP
migration to
Golgi complex
SREBP
Endoplasmic
reticulum
Sterol (binds
SCAP, prevents
release of SREBP)
Cleavage
by first
protease
Cleavage
by second
protease
Transcription
of target genes
is activated
released
domain of
SREBP
migrates
to nucleus
N
C
N
N
N
C
C
C
N N
DNA
Nucleus
C
FIGURE 21–43 SREBP activation. Sterol regulatory element-binding
proteins (SREBPs, shown in green) are embedded in the ER when first
synthesized, in a complex with the protein SREBP cleavage-activating
protein (SCAP, red). (N and C represent the amino and carboxyl ter-
mini of the proteins.) When bound to SCAP, SREBPs are inactive. When
sterol levels decline, the complex migrates to the Golgi complex, and
SREBP is cleaved by two different proteases in succession. The liber-
ated amino-terminal domain of SREBP migrates to the nucleus, where
it activates transcription of sterol-regulated genes.
insulin
Acetyl-CoA
-Hydroxy- -methyl-
glutaryl-CoA
Mevalonate
Cholesterol
(intracellular)
HMG-CoA
reductase
glucagon
X
receptor-
mediated
endocytosis
ACAT
Cholesteryl
esters
LDL-cholesterol
(extracellular)
H9252H9252
stimulates
proteolysis
of HMG-CoA
reductase
multistep
multistep
FIGURE 21–44 Regulation of cholesterol formation balances syn-
thesis with dietary uptake. Glucagon promotes phosphorylation (in-
activation) of HMG-CoA reductase; insulin promotes dephosphoryla-
tion (activation). X represents unidentified metabolites of cholesterol
that stimulate proteolysis of HMG-CoA reductase.
8885d_c21_826 2/27/04 1:45 PM Page 826 Mac113 mac113:
FIGURE 21–46 Some steroid hormones derived from cholesterol. The
structures of some of these compounds are shown in Figure 10–19.
Cholesterol
Corticosterone
(mineralocorticoid)
Affects protein and
carbohydrate metabolism;
suppresses immune
response, inflammation,
and allergic responses.
Pregnenolone
Progesterone
Aldosterone
(mineralocorticoid)
Testosterone
Cortisol
(glucocorticoid)
Estradiol
Male and female sex
hormones. Influence
secondary sexual char-
acteristics; regulate
female reproductive
cycle.
Regulate reabsorption
of Na
H11001
, Cl
H5008
, HCO
H5008
in
the kidney.
3
Unregulated cholesterol production can lead to
serious human disease. When the sum of choles-
terol synthesized and cholesterol obtained in the diet
exceeds the amount required for the synthesis of mem-
branes, bile salts, and steroids, pathological accumula-
tions of cholesterol in blood vessels (atherosclerotic
plaques) can develop, resulting in obstruction of blood
vessels (atherosclerosis). Heart failure due to oc-
cluded coronary arteries is a leading cause of death in
industrialized societies. Atherosclerosis is linked to high
levels of cholesterol in the blood, and particularly to high
levels of LDL-bound cholesterol; there is a negative cor-
relation between HDL levels and arterial disease.
In familial hypercholesterolemia, a human genetic
disorder, blood levels of cholesterol are extremely high
and severe atherosclerosis develops in childhood. These
individuals have a defective LDL receptor and lack
receptor-mediated uptake of cholesterol carried by LDL.
Consequently, cholesterol is not cleared from the blood;
it accumulates and contributes to the formation of ath-
erosclerotic plaques. Endogenous cholesterol synthesis
continues despite the excessive cholesterol in the blood,
because extracellular cholesterol cannot enter the cell
to regulate intracellular synthesis (Fig. 21–44). Two
products derived from fungi, lovastatin and com-
pactin, are used to treat patients with familial hyper-
cholesterolemia. Both these compounds, and several
synthetic analogs, resemble mevalonate (Fig. 21–45)
and are competitive inhibitors of HMG-CoA reductase,
thus inhibiting cholesterol synthesis. Lovastatin treat-
ment lowers serum cholesterol by as much as 30% in
individuals having one defective copy of the gene for the
LDL receptor. When combined with an edible resin that
binds bile acids and prevents their reabsorption from
the intestine, the drug is even more effective.
In familial HDL deficiency, HDL levels are very low;
they are almost undetectable in Tangier disease. Both
genetic disorders are the result of mutations in the
ABC1 protein. Cholesterol-depleted HDL cannot take
up cholesterol from cells that lack ABC1 protein, and
cholesterol-poor HDL is rapidly removed from the blood
and destroyed. Both familial HDL deficiency and Tang-
ier disease are very rare (worldwide, fewer than 100
families with Tangier disease are known), but the exis-
tence of these diseases establishes a role for ABC1 pro-
tein in the regulation of plasma HDL levels. Because low
plasma HDL levels correlate with a high incidence of
coronary artery disease, the ABC1 protein may prove a
useful target for drugs to control HDL levels. ■
Steroid Hormones Are Formed by Side-Chain
Cleavage and Oxidation of Cholesterol
Humans derive all their steroid hormones from choles-
terol (Fig. 21–46). Two classes of steroid hormones
are synthesized in the cortex of the adrenal gland:
mineralocorticoids, which control the reabsorption of
inorganic ions (Na
H11001
, Cl
H11002
, and HCO
3
H11002
) by the kidney,
and glucocorticoids, which
help regulate gluconeogene-
sis and reduce the inflamma-
tory response. Sex hormones
are produced in male and fe-
male gonads and the pla-
centa. They include proges-
terone, which regulates the
female reproductive cycle,
and androgens (such as testosterone) and estrogens
(such as estradiol), which influence the development of
21.4 Biosynthesis of Cholesterol, Steroids, and Isoprenoids 827
FIGURE 21–45 Inhibitors of HMG-CoA reductase. A comparison of
the structures of mevalonate and four pharmaceutical compounds that
inhibit HMG-CoA reductase.
Mevalonate
CH
3
CH
3
R
2
R
1
H11005 H
R
1
H11005 CH
3
R
1
H11005 H
R
1
H11005 H
R
2
H11005 H
R
2
H11005 CH
3
R
2
H11005 OH
R
2
H11005 CH
3
Compactin
Simvastatin (Zocor)
Pravastatin (Pravachol)
Lovastatin (Mevacor)
R
1
COO
H5008
OH
HO
H
3
C
COO
H5008
OH
O
O
HO
C
O
CH
3
Progesterone
O
8885d_c21_787-832 2/26/04 9:35 AM Page 827 mac76 mac76:385_reb:
secondary sexual characteristics in males and females,
respectively. Steroid hormones are effective at very low
concentrations and are therefore synthesized in rela-
tively small quantities. In comparison with the bile salts,
their production consumes relatively little cholesterol.
Synthesis of steroid hormones requires removal of
some or all of the carbons in the “side chain” on C-17
of the D ring of cholesterol. Side-chain removal takes
place in the mitochondria of steroidogenic tissues. Re-
moval involves the hydroxylation of two adjacent car-
bons in the side chain (C-20 and C-22) followed by
cleavage of the bond between them (Fig. 21–47). For-
mation of the various hormones also involves the intro-
duction of oxygen atoms. All the hydroxylation and oxy-
genation reactions in steroid biosynthesis are catalyzed
by mixed-function oxidases (Box 21–1) that use
NADPH, O
2
, and mitochondrial cytochrome P-450.
Intermediates in Cholesterol Biosynthesis Have Many
Alternative Fates
In addition to its role as an intermediate in cholesterol
biosynthesis, isopentenyl pyrophosphate is the acti-
vated precursor of a huge array of biomolecules with di-
verse biological roles (Fig. 21–48). They include vita-
mins A, E, and K; plant pigments such as carotene and
the phytol chain of chlorophyll; natural rubber; many
essential oils (such as the fragrant principles of lemon
oil, eucalyptus, and musk); insect juvenile hormone,
which controls metamorphosis; dolichols, which serve
as lipid-soluble carriers in complex polysaccharide
synthesis; and ubiquinone and plastoquinone, electron
carriers in mitochondria and chloroplasts. Collectively,
these molecules are called isoprenoids. More than
Chapter 21 Lipid Biosynthesis828
Bile acids
Vitamin DSteroid
hormones
Cholesterol
Rubber
Vitamin A
Quinone
electron
carriers:
ubiquinone,
plastoquinone
Dolichols
Isoprene
Phytol chain
of chlorophyll
Vitamin E
Vitamin K
Carotenoids
Plant hormones
abscisic acid
and gibberellic
acid
H9004
3
-Isopentenyl
pyrophosphate
CH
2
C
O
CH
3
CH
2
CH
2
P
O
PO
O
H11002
O
H11002
OO
H11002
FIGURE 21–48 Overview of isoprenoid biosynthesis. The structures
of most of the end products shown here are given in Chapter 10.
FIGURE 21–47 Side-chain cleavage in the synthesis of steroid hor-
mones. Cytochrome P-450 acts as electron carrier in this mixed-
function oxidase system that oxidizes adjacent carbons. The process
also requires the electron-transferring proteins adrenodoxin and
adrenodoxin reductase. This system for cleaving side chains is found
in mitochondria of the adrenal cortex, where active steroid pro-
duction occurs. Pregnenolone is the precursor of all other steroid
hormones (see Fig. 21–46).
17
22
20
Cholesterol
Isocaproaldehyde
Pregnenolone
mixed-function
oxidase
cyt P-450
adrenodoxin
(Fe–S)
adrenodoxin
reductase
(flavoprotein)
2O
2
2H
2
O
OH
desmolase
O
2
H11001
2H
H11001
NADPH
H11001 H
2
ONADP
H11001
2NADP
H11001
H11001 O
2
H
C
OCH
3
C
NADPH H11001 H
H11001
HO
HO
HO
HO
20,22-Dihydroxycholesterol
8885d_c21_787-832 2/26/04 9:35 AM Page 828 mac76 mac76:385_reb:
20,000 different isoprenoid molecules have been dis-
covered in nature, and hundreds of new ones are re-
ported each year.
Prenylation (covalent attachment of an isoprenoid;
see Fig. 27–30) is a common mechanism by which pro-
teins are anchored to the inner surface of cellular mem-
branes in mammals (see Fig. 11–14). In some of these
proteins the attached lipid is the 15-carbon farnesyl
group; others have the 20-carbon geranylgeranyl group.
Different enzymes attach the two types of lipids. It is
possible that prenylation reactions target proteins to dif-
ferent membranes, depending on which lipid is at-
tached. Protein prenylation is another important role for
the isoprene derivatives of the pathway to cholesterol.
SUMMARY 21.4 Biosynthesis of Cholesterol,
Steroids, and Isoprenoids
■ Cholesterol is formed from acetyl-CoA in a
complex series of reactions, through the
intermediates H9252-hydroxy-H9252-methylglutaryl-CoA,
mevalonate, and two activated isoprenes,
dimethylallyl pyrophosphate and isopentenyl
pyrophosphate. Condensation of isoprene units
produces the noncyclic squalene, which is
cyclized to yield the steroid ring system and
side chain.
■ Cholesterol synthesis is under hormonal control
and is also inhibited by elevated concentrations
of intracellular cholesterol, which acts through
covalent modification and transcriptional
regulation mechanisms.
■ Cholesterol and cholesteryl esters are carried
in the blood as plasma lipoproteins. VLDL
carries cholesterol, cholesteryl esters, and
triacylglycerols from the liver to other tissues,
where the triacylglycerols are degraded by
lipoprotein lipase, converting VLDL to LDL.
The LDL, rich in cholesterol and its esters, is
taken up by receptor-mediated endocytosis, in
which the apolipoprotein B-100 of LDL is
recognized by receptors in the plasma
membrane. HDL removes cholesterol from the
blood, carrying it to the liver. Dietary
conditions or genetic defects in cholesterol
metabolism may lead to atherosclerosis and
heart disease.
■ The steroid hormones (glucocorticoids,
mineralocorticoids, and sex hormones) are
produced from cholesterol by alteration of the
side chain and introduction of oxygen atoms
into the steroid ring system. In addition to
cholesterol, a wide variety of isoprenoid
compounds are derived from mevalonate
through condensations of isopentenyl
pyrophosphate and dimethylallyl
pyrophosphate.
■ Prenylation of certain proteins targets them for
association with cellular membranes and is
essential for their biological activity.
Chapter 21 Key Terms 829
Key Terms
acetyl-CoA carboxylase 787
fatty acid synthase 789
acyl carrier protein (ACP) 790
fatty acyl-CoA desaturase 798
mixed-function oxidases 799
mixed-function oxygenases 799
cytochrome P-450 799
essential fatty acids 800
prostaglandins 800
cyclooxygenase (COX) 800
prostaglandin H
2
synthase 800
thromboxane synthase 800
thromboxanes 800
leukotrienes 800
glycerol 3-phosphate
dehydrogenase 804
triacylglycerol cycle 806
glyceroneogenesis 806
thiazolidinediones 807
phosphatidylserine 811
phosphatidylglycerol 811
phosphatidylethanolamine 811
cardiolipin 811
phosphatidylcholine 812
plasmalogen 813
platelet-activating factor 813
cerebroside 813
sphingomyelin 813
gangliosides 813
isoprene 816
mevalonate 817
H9252-hydroxy-H9252-methylglutaryl-CoA
(HMG-CoA) 817
thiolase 817
HMG-CoA synthase 817
HMG-CoA reductase 817
bile acids 820
cholesteryl esters 820
apolipoproteins 821
chylomicron 821
very-low-density lipoprotein
(VLDL) 822
low-density lipoprotein (LDL) 823
high-density lipoprotein (HDL) 823
reverse cholesterol transport 824
LDL receptors 824
receptor-mediated endocytosis 824
atherosclerosis 827
lovastatin 827
mineralocorticoids 827
glucocorticoids 827
progesterone 827
androgens 827
estrogens 827
Terms in bold are defined in the glossary.
8885d_c21_787-832 2/26/04 9:35 AM Page 829 mac76 mac76:385_reb:
Chapter 21 Lipid Biosynthesis830
Further Reading
The general references in Chapters 10 and 17 are also useful.
General
Bell, S.J., Bradley, D., Forse, R.A., & Bistrian, B.R. (1997)
The new dietary fats in health and disease. J. Am. Dietetic Assoc.
97, 280–286.
Gotto, A.M., Jr. (ed.) (1987) Plasma Lipoproteins, New
Comprehensive Biochemistry, Vol. 14 (Neuberger, A. & van
Deenen, L.L.M., series eds), Elsevier Biomedical Press,
Amsterdam.
Twelve reviews cover the structure, synthesis, and metabolism
of lipoproteins, regulation of cholesterol synthesis, and the
enzymes LCAT and lipoprotein lipase.
Hajjar, D.P. & Nicholson, A.C. (1995) Atherosclerosis. Am. Sci.
83, 460–467.
A good description of the molecular basis of this disease and
prospects for therapy.
Hawthorne, J.N. & Ansell, G.B. (eds) (1982) Phospholipids,
New Comprehensive Biochemistry, Vol. 4 (Neuberger, A. & van
Deenen, L.L.M., series eds), Elsevier Biomedical Press,
Amsterdam.
This volume has excellent reviews of biosynthetic pathways to
glycerophospholipids and sphingolipids, phospholipid transfer
proteins, and bilayer assembly.
Ohlrogge, J. & Browse, J. (1995) Lipid biosynthesis. Plant Cell
7, 957–970.
A good summary of pathways for lipid biosynthesis in plants.
Vance, D.E. & Vance, J.E. (eds) (1996) Biochemistry of
Lipids, Lipoproteins, and Membranes, New Comprehensive
Biochemistry, Vol. 31, Elsevier Science Publishing Co., Inc.,
New York.
Excellent reviews of lipid structure, biosynthesis, and function.
Biosynthesis of Fatty Acids and Eicosanoids
Capdevila, J.H., Falck, J.R., & Estabrook, R.W. (1992)
Cytochrome P450 and the arachidonate cascade. FASEB J. 6,
731–736.
This issue contains 20 articles on the structure and function of
various types of cytochrome P-450.
Creelman, R.A. & Mullet, J.E. (1997) Biosynthesis and action of
jasmonates in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol.
48, 355–381.
DeWitt, D.L. (1999) Cox-2-selective inhibitors: the new super
aspirins. Mol. Pharmacol. 55, 625–631.
A short, clear review of the topic discussed in Box 21–2.
Drazen, J.M., Israel, E., & O’Byrne, P.M. (1999) Drug therapy:
treatment of asthma with drugs modifying the leukotriene pathway.
New Engl. J. Med. 340, 197–206.
Lands, W.E.M. (1991) Biosynthesis of prostaglandins. Annu.
Rev. Nutr. 11, 41–60.
Discussion of the nutritional requirement for unsaturated fatty
acids and recent biochemical work on pathways from
arachidonate to prostaglandins; advanced level.
Munday, M.R. (2002) Regulation of mammalian acetyl-CoA
carboxylase. Biochem. Soc. Trans. 30, 1059–1064.
Reshef, L., Olswang, Y., Cassuto, H., Blum, B., Croniger,
C.M., Kalhan, S.C., Tilghman, S.M., & Hanson, R.M. (2003)
Glyceroneogenesis and the triglyceride/fatty acid cycle. J. Biol.
Chem. 278, 30,413–30,416.
Smith, S. (1994) The animal fatty acid synthase: one gene, one
polypeptide, seven enzymes. FASEB J. 8, 1248–1259.
Smith, W.L., Garavito, R.M., & DeWitt, D.L. (1996)
Prostaglandin endoperoxide H synthases (cyclooxygenases)-1 and -2.
J. Biol. Chem. 271, 33,157–33,160.
A concise review of the properties and roles of COX-1 and
COX-2.
Biosynthesis of Membrane Phospholipids
Bishop, W.R. & Bell, R.M. (1988) Assembly of phospholipids into
cellular membranes: biosynthesis, transmembrane movement and
intracellular translocation. Annu. Rev. Cell Biol. 4, 579–610.
Advanced review of the enzymology and cell biology of
phospholipid synthesis and targeting.
Dowhan, W. (1997) Molecular basis for membrane phospholipid
diversity: why are there so many lipids? Annu. Rev. Biochem. 66,
199–232.
Kennedy, E.P. (1962) The metabolism and function of complex
lipids. Harvey Lect. 57, 143–171.
A classic description of the role of cytidine nucleotides in
phospholipid synthesis.
Pavlidis, P., Ramaswami, M., & Tanouye, M.A. (1994) The
Drosophila easily shocked gene: a mutation in a phospholipid
synthetic pathway causes seizure, neuronal failure, and paralysis.
Cell 79, 23–33.
Description of the fascinating effects of changing the
composition of membrane lipids in fruit flies.
Raetz, C.R.H. & Dowhan, W. (1990) Biosynthesis and function
of phospholipids in Escherichia coli. J. Biol. Chem. 265,
1235–1238.
A brief review of bacterial biosynthesis of phospholipids and
lipopolysaccharides.
Biosynthesis of Cholesterol, Steroids, and Isoprenoids
Bittman, R. (ed.) (1997) Subcellular Biochemistry, Vol. 28:
Cholesterol: Its Functions and Metabolism in Biology and
Medicine, Plenum Press, New York.
Bloch, K. (1965) The biological synthesis of cholesterol. Science
150, 19–28.
The author’s Nobel address; a classic description of cholesterol
synthesis in animals.
Brown, M.S., & Goldstein, J.L. (1999) A proteolytic pathway
that controls the cholesterol content of membranes, cells, and
blood. Proc. Natl. Acad. Sci. USA 96, 11,041–11,048.
Chang, T.Y., Chang, C.C.Y., & Cheng, D. (1997) Acyl–coenzyme
A: cholesterol acyltransferase. Annu. Rev. Biochem. 66, 613–638.
Edwards, P.A. & Ericsson, J. (1999) Sterols and isoprenoids:
signaling molecules derived from the cholesterol biosynthetic
pathway. Annu. Rev. Biochem. 68, 157–185.
8885d_c21_787-832 2/26/04 9:35 AM Page 830 mac76 mac76:385_reb:
Chapter 21 Problems 831
Gimpl, G., Burger, K., & Fahrenholz, F. (2002) A closer look at
the cholesterol sensor. Trends Biochem. Sci. 27, 596–599.
Goldstein, J.L. & Brown, M.S. (1990) Regulation of the
mevalonate pathway. Nature 343, 425–430.
Description of the allosteric and covalent regulation of the
enzymes of the mevalonate pathway; includes a short
discussion of the prenylation of Ras and other proteins.
Knopp, R.H. (1999) Drug therapy: drug treatment of lipid
disorders. New Engl. J. Med. 341, 498–511.
Review of the use of HMG-CoA inhibitors and bile acid–binding
resins to reduce serum cholesterol.
Krieger, M. (1999) Charting the fate of the “good cholesterol”:
identification and characterization of the high-density lipoprotein
receptor SR-BI. Annu. Rev. Biochem. 68, 523–558.
McGarvey, D.J. & Croteau, R. (1995) Terpenoid metabolism.
Plant Cell 7, 1015–1026.
A description of the amazing diversity of isoprenoids in plants.
Olson, R.E. (1998) Discovery of the lipoproteins, their role in fat
transport and their significance as risk factors. J. Nutr. 128
(2 Suppl.), 439S–443S.
Brief, clear, historical background on studies of lipoprotein
function.
Russell, D.W. (2003) The enzymes, regulation, and genetics of
bile acid synthesis. Annu. Rev. Biochem. 72, 137–174.
Young, S.G. & Fielding, C.J. (1999) The ABCs of cholesterol
efflux. Nat. Genet. 22, 316–318.
A brief review of three papers in this journal issue that
establish mutations in ABC1 as the cause of Tangier disease
and familial HDL deficiency.
1. Pathway of Carbon in Fatty Acid Synthesis Using
your knowledge of fatty acid biosynthesis, provide an expla-
nation for the following experimental observations:
(a) Addition of uniformly labeled [
14
C]acetyl-CoA to a
soluble liver fraction yields palmitate uniformly labeled with
14
C.
(b) However, addition of a trace of uniformly labeled
[
14
C]acetyl-CoA in the presence of an excess of unlabeled
malonyl-CoA to a soluble liver fraction yields palmitate
labeled with
14
C only in C-15 and C-16.
2. Synthesis of Fatty Acids from Glucose After a per-
son has ingested large amounts of sucrose, the glucose and
fructose that exceed caloric requirements are transformed to
fatty acids for triacylglycerol synthesis. This fatty acid syn-
thesis consumes acetyl-CoA, ATP, and NADPH. How are these
substances produced from glucose?
3. Net Equation of Fatty Acid Synthesis Write the net
equation for the biosynthesis of palmitate in rat liver, start-
ing from mitochondrial acetyl-CoA and cytosolic NADPH,
ATP, and CO
2
.
4. Pathway of Hydrogen in Fatty Acid Synthesis Con-
sider a preparation that contains all the enzymes and cofac-
tors necessary for fatty acid biosynthesis from added acetyl-
CoA and malonyl-CoA.
(a) If [2-
2
H]acetyl-CoA (labeled with deuterium, the
heavy isotope of hydrogen)
and an excess of unlabeled malonyl-CoA are added as sub-
strates, how many deuterium atoms are incorporated into
every molecule of palmitate? What are their locations? Explain.
(b) If unlabeled acetyl-CoA and [2-
2
H]malonyl-CoA
are added as substrates, how many deuterium atoms are in-
corporated into every molecule of palmitate? What are their
locations? Explain.
5. Energetics of H9252-Ketoacyl-ACP Synthase In the
condensation reaction catalyzed by H9252-ketoacyl-ACP synthase
(see Fig. 21–5), a four-carbon unit is synthesized by the com-
bination of a two-carbon unit and a three-carbon unit, with
the release of CO
2
. What is the thermodynamic advantage of
this process over one that simply combines two two-carbon
units?
6. Modulation of Acetyl-CoA Carboxylase Acetyl-
CoA carboxylase is the principal regulation point in the
biosynthesis of fatty acids. Some of the properties of the en-
zyme are described below.
(a) Addition of citrate or isocitrate raises the V
max
of the
enzyme as much as 10-fold.
(b) The enzyme exists in two interconvertible forms that
differ markedly in their activities:
Protomer (inactive) filamentous polymer (active)
Citrate and isocitrate bind preferentially to the filamentous
form, and palmitoyl-CoA binds preferentially to the protomer.
Explain how these properties are consistent with the reg-
ulatory role of acetyl-CoA carboxylase in the biosynthesis of
fatty acids.
7. Shuttling of Acetyl Groups across the Mitochon-
drial Inner Membrane The acetyl group of acetyl-CoA,
produced by the oxidative decarboxylation of pyruvate in the
z
y
H5008
OOC C
2
H
2
H
C
O
S-CoA
2
H C
2
H
2
H
C
O
S-CoA
Problems
8885d_c21_787-832 2/26/04 9:35 AM Page 831 mac76 mac76:385_reb:
Chapter 21 Lipid Biosynthesis832
mitochondrion, is transferred to the cytosol by the acetyl
group shuttle outlined in Figure 21-10.
(a) Write the overall equation for the transfer of one
acetyl group from the mitochondrion to the cytosol.
(b) What is the cost of this process in ATPs per acetyl
group?
(c) In Chapter 17 we encountered an acyl group shuttle
in the transfer of fatty acyl–CoA from the cytosol to the mito-
chondrion in preparation for H9252 oxidation (see Fig. 17–6). One
result of that shuttle was separation of the mitochondrial and
cytosolic pools of CoA. Does the acetyl group shuttle also
accomplish this? Explain.
8. Oxygen Requirement for Desaturases The biosyn-
thesis of palmitoleate (see Fig. 21–12), a common unsatu-
rated fatty acid with a cis double bond in the H9004
9
position, uses
palmitate as a precursor. Can this be carried out under strictly
anaerobic conditions? Explain.
9. Energy Cost of Triacylglycerol Synthesis Use a net
equation for the biosynthesis of tripalmitoylglycerol (tri-
palmitin) from glycerol and palmitate to show how many ATPs
are required per molecule of tripalmitin formed.
10. Turnover of Triacylglycerols in Adipose Tissue
When [
14
C]glucose is added to the balanced diet of adult rats,
there is no increase in the total amount of stored triacyl-
glycerols, but the triacylglycerols become labeled with
14
C.
Explain.
11. Energy Cost of Phosphatidylcholine Synthesis
Write the sequence of steps and the net reaction for the
biosynthesis of phosphatidylcholine by the salvage pathway
from oleate, palmitate, dihydroxyacetone phosphate, and
choline. Starting from these precursors, what is the cost (in
number of ATPs) of the synthesis of phosphatidylcholine by
the salvage pathway?
12. Salvage Pathway for Synthesis of Phosphatidyl-
choline A young rat maintained on a diet deficient in me-
thionine fails to thrive unless choline is included in the diet.
Explain.
13. Synthesis of Isopentenyl Pyrophosphate If 2-
[
14
C]acetyl-CoA is added to a rat liver homogenate that is syn-
thesizing cholesterol, where will the
14
C label appear in H9004
3
-
isopentenyl pyrophosphate, the activated form of an isoprene
unit?
14. Activated Donors in Lipid Synthesis In the biosyn-
thesis of complex lipids, components are assembled by trans-
fer of the appropriate group from an activated donor. For ex-
ample, the activated donor of acetyl groups is acetyl-CoA.
For each of the following groups, give the form of the acti-
vated donor: (a) phosphate; (b) D-glucosyl; (c) phospho-
ethanolamine; (d) D-galactosyl; (e) fatty acyl; (f) methyl;
(g) the two-carbon group in fatty acid biosynthesis; (h) H9004
3
-
isopentenyl.
15. Importance of Fats in the Diet When young rats are
placed on a totally fat-free diet, they grow poorly, develop a
scaly dermatitis, lose hair, and soon die—symptoms that can
be prevented if linoleate or plant material is included in the
diet. What makes linoleate an essential fatty acid? Why can
plant material be substituted?
16. Regulation of Cholesterol Biosynthesis Cholesterol
in humans can be obtained from the diet or synthesized de novo.
An adult human on a low-cholesterol diet typically synthesizes
600 mg of cholesterol per day in the liver. If the amount of cho-
lesterol in the diet is large, de novo synthesis of cholesterol is
drastically reduced. How is this regulation brought about?
8885d_c21_787-832 2/26/04 9:35 AM Page 832 mac76 mac76:385_reb:
chapter
N
itrogen ranks behind only carbon, hydrogen, and
oxygen in its contribution to the mass of living sys-
tems. Most of this nitrogen is bound up in amino acids
and nucleotides. In this chapter we address all aspects
of the metabolism of these nitrogen-containing com-
pounds except amino acid catabolism, which is covered
in Chapter 18.
Discussing the biosynthetic pathways for amino
acids and nucleotides together is a sound approach, not
only because both classes of molecules contain nitrogen
(which arises from common biological sources) but be-
cause the two sets of pathways are extensively inter-
twined, with several key intermediates in common. Cer-
tain amino acids or parts of amino acids are incorporated
into the structure of purines and pyrimidines, and in one
case part of a purine ring is incorporated into an amino
acid (histidine). The two sets of pathways also share
much common chemistry, in particular a preponderance
of reactions involving the transfer of nitrogen or one-
carbon groups.
The pathways described here can be intimidating to
the beginning biochemistry student. Their complexity
arises not so much from the chemistry itself, which in
many cases is well understood, but from the sheer num-
ber of steps and variety of intermediates. These path-
ways are best approached by maintaining a focus on
metabolic principles we have already discussed, on key
intermediates and precursors, and on common classes
of reactions. Even a cursory look at the chemistry can
be rewarding, for some of the most unusual chemical
transformations in biological systems occur in these
pathways; for instance, we find prominent examples of
the rare biological use of the metals molybdenum, sele-
nium, and vanadium. The effort also offers a practical
dividend, especially for students of human or veterinary
medicine. Many genetic diseases of humans and animals
have been traced to an absence of one or more enzymes
of amino acid and nucleotide metabolism, and many
pharmaceuticals in common use to combat infectious
diseases are inhibitors of enzymes in these pathways—
as are a number of the most important agents in cancer
chemotherapy.
Regulation is crucial in the biosynthesis of the
nitrogen-containing compounds. Because each amino
acid and each nucleotide is required in relatively small
amounts, the metabolic flow through most of these path-
ways is not nearly as great as the biosynthetic flow lead-
ing to carbohydrate or fat in animal tissues. Because the
different amino acids and nucleotides must be made in
BIOSYNTHESIS OF AMINO ACIDS,
NUCLEOTIDES, AND RELATED
MOLECULES
22.1 Overview of Nitrogen Metabolism 834
22.2 Biosynthesis of Amino Acids 841
22.3 Molecules Derived from Amino Acids 854
22.4 Biosynthesis and Degradation of Nucleotides 862
Time passes rapidly when you are having fun. The thrill of
seeing people get well who might otherwise have died of
disease . . . cannot be described in words. The Nobel Prize
was only the icing on the cake.
—Gertrude Elion, quoted in an article in Science, 2002
22
833
8885d_c22_833-880 2/6/04 8:35 AM Page 833 mac76 mac76:385_reb:
the correct ratios and at the right time for protein and
nucleic acid synthesis, their biosynthetic pathways must
be accurately regulated and coordinated with each other.
And because amino acids and nucleotides are charged
molecules, their levels must be regulated to maintain
electrochemical balance in the cell. As discussed in ear-
lier chapters, pathways can be controlled by changes in
either the activity or the amounts of specific enzymes.
The pathways we encounter in this chapter provide some
of the best-understood examples of the regulation of
enzyme activity. Control of the amounts of different
enzymes in a cell (that is, of their synthesis and degra-
dation) is a topic covered in Chapter 28.
22.1 Overview of Nitrogen Metabolism
The biosynthetic pathways leading to amino acids and
nucleotides share a requirement for nitrogen. Because
soluble, biologically useful nitrogen compounds are gen-
erally scarce in natural environments, most organisms
maintain strict economy in their use of ammonia, amino
acids, and nucleotides. Indeed, as we shall see, free
amino acids, purines, and pyrimidines formed during
metabolic turnover of proteins and nucleic acids are of-
ten salvaged and reused. We first examine the pathways
by which nitrogen from the environment is introduced
into biological systems.
The Nitrogen Cycle Maintains a Pool of Biologically
Available Nitrogen
The most important source of nitrogen is air, which is
four-fifths molecular nitrogen (N
2
). However, relatively
few species can convert atmospheric nitrogen into forms
useful to living organisms. In the biosphere, the meta-
bolic processes of different species function interde-
pendently to salvage and reuse biologically available ni-
trogen in a vast nitrogen cycle (Fig. 22–1). The first
step in the cycle is fixation (reduction) of atmospheric
nitrogen by nitrogen-fixing bacteria to yield ammonia
(NH
3
or NH
4
H11001
). Although ammonia can be used by most
living organisms, soil bacteria that derive their energy
by oxidizing ammonia to nitrite (NO
2
H11002
) and ultimately
nitrate (NO
3
H11002
) are so abundant and active that nearly all
ammonia reaching the soil is oxidized to nitrate. This
process is known as nitrification. Plants and many bac-
teria can take up and readily reduce nitrate and nitrite
through the action of nitrate and nitrite reductases. The
ammonia so formed is incorporated into amino acids by
plants. Animals then use plants as a source of amino
acids, both nonessential and essential, to build their pro-
teins. When organisms die, microbial degradation of
their proteins returns ammonia to the soil, where nitri-
fying bacteria again convert it to nitrite and nitrate. A
balance is maintained between fixed nitrogen and at-
mospheric nitrogen by bacteria that convert nitrate to
N
2
under anaerobic conditions, a process called deni-
trification (Fig. 22–1). These soil bacteria use NO
3
H11002
rather than O
2
as the ultimate electron acceptor in a se-
ries of reactions that (like oxidative phosphorylation)
generates a transmembrane proton gradient, which is
used to synthesize ATP.
Now let’s examine the process of nitrogen fixation,
the first step in the nitrogen cycle.
Nitrogen Is Fixed by Enzymes
of the Nitrogenase Complex
Only certain prokaryotes can fix atmospheric nitrogen.
These include the cyanobacteria of soils and fresh and
salt waters, other kinds of free-living soil bacteria such
as Azotobacter species, and the nitrogen-fixing bacte-
ria that live as symbionts in the root nodules of legu-
minous plants. The first important product of nitrogen
fixation is ammonia, which can be used by all organisms
either directly or after its conversion to other soluble
compounds such as nitrites, nitrates, or amino acids.
The reduction of nitrogen to ammonia is an exer-
gonic reaction:
N
2
H11001 3H
2
88n 2NH
3
H9004GH11032H11034 H11005 H1100233.5 kJ/mol
reduction
by some
anaerobic
bacteria,
most
plants
N
2
Ammonia
NH
4
Nitrite
NO
2
nitrogen fixation
by some bacteria
(e.g., Klebsiella,
Azotobacter, Rhizobium)
denitrification
nitrification
by soil bacteria
(e.g., Nitrobacter)
nitrification
by soil bacteria
(e.g., Nitrosomonas)
synthesis in
plants and
microorganisms
degradation
by animals and
microorganisms
Amino acids
and other
reduced
nitrogen-carbon
compounds
Nitrate
NO
3
H11002
H11002
H11001
FIGURE 22–1 The nitrogen cycle. The total amount of nitrogen fixed
annually in the biosphere exceeds 10
11
kg.
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules834
8885d_c22_833-880 2/6/04 8:35 AM Page 834 mac76 mac76:385_reb:
The NmN triple bond, however, is very stable, with a
bond energy of 930 kJ/mol. Nitrogen fixation therefore
has an extremely high activation energy, and atmos-
pheric nitrogen is almost chemically inert under normal
conditions. Ammonia is produced industrially by the
Haber process (named for its inventor, Fritz Haber),
which requires temperatures of 400 to 500 H11034C and ni-
trogen and hydrogen at pressures of tens of thousands
of kilopascals (several hundred atmospheres) to provide
the necessary activation energy. Biological nitrogen fix-
ation, however, must occur at biological temperatures
and at 0.8 atm of nitrogen, and the high activation bar-
rier is overcome by other means. This is accomplished,
at least in part, by the binding and hydrolysis of ATP.
The overall reaction can be written
N
2
H11001 10H
H11001
H11001 8e
H11002
H11001 16ATP 88n 2NH
4
H11001
H11001 16ADP H11001 16P
i
H11001 H
2
Biological nitrogen fixation is carried out by a highly
conserved complex of proteins called the nitrogenase
complex (Fig. 22–2), the crucial components of which
are dinitrogenase reductase and dinitrogenase
(Fig. 22–3). Dinitrogenase reductase (M
r
60,000) is a
dimer of two identical subunits. It contains a single 4Fe-
4S redox center (see Fig. 19–5), bound between the
subunits, and can be oxidized and reduced by one elec-
tron. It also has two binding sites for ATP/ADP (one site
on each subunit). Dinitrogenase (M
r
240,000), a
tetramer with two copies of two different subunits, con-
tains both iron and molybdenum; its redox centers have
a total of 2 Mo, 32 Fe, and 30 S per tetramer. About half
of the iron and sulfur is present as two bridged pairs of
4Fe-4S centers called P clusters; the remainder is pres-
ent as part of a novel iron-molybdenum cofactor. A form
of nitrogenase that contains vanadium rather than
molybdenum has been discovered, and some bacterial
species can produce both types of nitrogenase systems.
The vanadium-containing enzyme may be the primary
nitrogen-fixing system under some environmental con-
ditions, but it is not yet as well characterized as the
molybdenum-dependent enzyme.
Nitrogen fixation is carried out by a highly reduced
form of dinitrogenase and requires eight electrons: six
for the reduction of N
2
and two to produce one mole-
cule of H
2
as an obligate part of the reaction mecha-
nism. Dinitrogenase is reduced by the transfer of elec-
trons from dinitrogenase reductase (Fig. 22–2). The
dinitrogenase tetramer has two binding sites for the re-
ductase. The required eight electrons are transferred
from reductase to dinitrogenase one at a time: a reduced
reductase molecule binds to the dinitrogenase and
transfers a single electron, then the oxidized reductase
dissociates from dinitrogenase, in a repeating cycle.
Each turn of the cycle requires the hydrolysis of two
ATP molecules by the dimeric reductase. The immedi-
ate source of electrons to reduce dinitrogenase reduc-
tase varies, with reduced ferredoxin (p. 733; see also
Fig. 19–5), reduced flavodoxin, and perhaps other
sources playing a role. In at least one species, the ulti-
mate source of electrons to reduce ferredoxin is pyru-
vate (Fig. 22–2).
The role of ATP in this process is somewhat un-
usual. As you will recall, ATP can contribute not only
chemical energy, through the hydrolysis of one or more
of its phosphoanhydride bonds, but also binding en-
ergy (pp. 196, 301), through noncovalent interactions
that lower the activation energy. In the reaction car-
ried out by dinitrogenase reductase, both ATP binding
22.1 Overview of Nitrogen Metabolism 835
8 Ferredoxin or
8 flavodoxin
(oxidized)
8 Dinitrogenase
reductase (oxidized)
+ 16ATP
Dinitrogenase
(oxidized)
Dinitrogenase
(reduced)
8 Dinitrogenase
reductase
(reduced)
8 Dinitrogenase
reductase
(oxidized)
8 Ferredoxin or
8 flavodoxin
(reduced)
N
2
2NH
4
8 Dinitrogenase
reductase (reduced)
+ 16ATP
2H
+
H
2
8e
H11002
8e
H11002
8e
H11002
16ADP
+ 16P
i
16 ATP
8e
H11002
4CoA +
4 pyruvate
4CO
2
+
4 acetyl-CoA
+
FIGURE 22–2 Nitrogen fixation by the nitrogenase complex. Elec-
trons are transferred from pyruvate to dinitrogenase via ferredoxin (or
flavodoxin) and dinitrogenase reductase. Dinitrogenase reductase re-
duces dinitrogenase one electron at a time, with at least six electrons
required to fix one molecule of N
2
. An additional two electrons are
used to reduce 2 H
H11001
to H
2
in a process that obligatorily accompanies
nitrogen fixation in anaerobes, making a total of eight electrons re-
quired per N
2
molecule. The subunit structures and metal cofactors of
the dinitrogenase reductase and dinitrogenase proteins are described
in the text and in Figure 22–3.
8885d_c22_833-880 2/6/04 8:35 AM Page 835 mac76 mac76:385_reb:
and ATP hydrolysis bring about protein conformational
changes that help overcome the high activation energy
of nitrogen fixation. The binding of two ATP molecules
to the reductase shifts the reduction potential (EH11032H11034) of
this protein from H11002300 to H11002420 mV, an enhancement
of its reducing power that is required to transfer elec-
trons to dinitrogenase. The ATP molecules are then hy-
drolyzed just before the actual transfer of one electron
to dinitrogenase.
Another important characteristic of the nitrogenase
complex is an extreme lability in the presence of oxy-
gen. The reductase is inactivated in air, with a half-life
of 30 seconds; dinitrogenase has a half-life of 10 min-
utes in air. Free-living bacteria that fix nitrogen cope
with this problem in a variety of ways. Some live only
anaerobically or repress nitrogenase synthesis when
oxygen is present. Some aerobic species, such as Azo-
tobacter vinelandii, partially uncouple electron trans-
fer from ATP synthesis so that oxygen is burned off as
rapidly as it enters the cell (see Box 19–1). When fix-
ing nitrogen, cultures of these bacteria actually increase
in temperature as a result of their efforts to rid them-
selves of oxygen.
The symbiotic relationship between leguminous
plants and the nitrogen-fixing bacteria in their root
nodules (Fig. 22–4) takes care of both the energy re-
quirements and the oxygen lability of the nitrogenase
complex. The energy required for nitrogen fixation
was probably the evolutionary driving force for this
plant-bacteria association. The bacteria in root nod-
ules have access to a large reservoir of energy in the
form of abundant carbohydrate and citric acid cycle
intermediates made available by the plant. This may
allow the bacteria to fix hundreds of times more ni-
trogen than their free-living cousins can fix under con-
ditions generally encountered in soils. To solve the
oxygen-toxicity problem, the bacteria in root nodules
are bathed in a solution of the oxygen-binding heme
protein leghemoglobin, produced by the plant (al-
though the heme may be contributed by the bacteria).
Leghemoglobin binds all available oxygen so that it
cannot interfere with nitrogen fixation, and efficiently
delivers the oxygen to the bacterial electron-transfer
system. The benefit to the plant, of course, is a ready
supply of reduced nitrogen. The efficiency of the sym-
biosis between plants and bacteria is evident in the
enrichment of soil nitrogen brought about by legumi-
nous plants. This enrichment is the basis of crop ro-
tation methods, in which plantings of nonleguminous
plants (such as maize) that extract fixed nitrogen from
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules836
(a)
(b) (c)
FIGURE 22–3 Enzymes and cofactors of the nitrogenase complex.
(PDB ID 1N2C) (a) In this ribbon diagram, the dinitrogenase subunits
are shown in gray and pink, the dinitrogenase reductase subunits in
blue and green. The bound ADP is red. Note the 4Fe-4S complex (Fe
atoms orange, S atoms yellow) and the iron-molybdenum cofactor (Mo
black, homocitrate light gray). The P clusters (bridged pairs of 4Fe-4S
complexes) are also shown. (b) The dinitrogenase complex cofactors
without the protein (colors as in (a)). (c) The iron-molybdenum co-
factor contains 1 Mo (black), 7 Fe (orange), 9 S (yellow), and one mol-
ecule of homocitrate (gray).
8885d_c22_833-880 2/6/04 8:35 AM Page 836 mac76 mac76:385_reb:
the soil are alternated every few years with plantings
of legumes such as alfalfa, peas, or clover.
Nitrogen fixation is the subject of intense study, be-
cause of its immense practical importance. Industrial
production of ammonia for use in fertilizers requires a
large and expensive input of energy, and this has
spurred a drive to develop recombinant or transgenic
organisms that can fix nitrogen. Recombinant DNA
techniques (Chapter 9) are being used to transfer the
DNA that encodes the enzymes of nitrogen fixation into
non-nitrogen-fixing bacteria and plants. Success in
these efforts will depend on overcoming the problem of
oxygen toxicity in any cell that produces nitrogenase.
Ammonia Is Incorporated into Biomolecules
through Glutamate and Glutamine
Reduced nitrogen in the form of NH
4
H11001
is assimilated into
amino acids and then into other nitrogen-containing bio-
molecules. Two amino acids, glutamate and glutamine,
provide the critical entry point. Recall that these same two
amino acids play central roles in the catabolism of am-
monia and amino groups in amino acid oxidation (Chap-
ter 18). Glutamate is the source of amino groups for most
other amino acids, through transamination reactions
(the reverse of the reaction shown in Fig. 18–4). The
amide nitrogen of glutamine is a source of amino groups
in a wide range of biosynthetic processes. In most types
of cells, and in extracellular fluids in higher organisms,
one or both of these amino acids are present at higher
concentrations—sometimes an order of magnitude or
more higher—than other amino acids. An Escherichia
coli cell requires so much glutamate that this amino acid
is one of the primary solutes in the cytosol. Its concen-
tration is regulated not only in response to the cell’s ni-
trogen requirements but also to maintain an osmotic bal-
ance between the cytosol and the external medium.
The biosynthetic pathways to glutamate and gluta-
mine are simple, and all or some of the steps occur in
most organisms. The most important pathway for the
22.1 Overview of Nitrogen Metabolism 837
FIGURE 22–4 Nitrogen-fixing nodules. (a) Root nodules of bird’s-foot
trefoil, a legume. The flower of this common plant is shown in the in-
set. (b) Artificially colorized electron micrograph of a thin section
through a pea root nodule. Symbiotic nitrogen-fixing bacteria, or bac-
teroids (red), live inside the nodule cells, surrounded by the peribac-
teroid membrane (blue). Bacteroids produce the nitrogenase complex
that converts atmospheric nitrogen (N
2
) to ammonium (NH
4
H11001
); with-
out the bacteroids, the plant is unable to utilize N
2
. The infected root
cells provide some factors essential for nitrogen fixation, including
leghemoglobin; this heme protein has a very high binding affinity for
oxygen, which strongly inhibits nitrogenase. (The cell nucleus is shown
in yellow/green. Not visible in this micrograph are other organelles of
the infected root cell that are normally found in plant cells.)
(a) (b)
2mH9262
8885d_c22_833-880 2/6/04 8:35 AM Page 837 mac76 mac76:385_reb:
assimilation of NH
4
H11001
into glutamate requires two reactions.
First, glutamine synthetase catalyzes the reaction of
glutamate and NH
4
H11001
to yield glutamine. This reaction
takes place in two steps, with enzyme-bound H9253-glutamyl
phosphate as an intermediate (see Fig. 18–8):
(1) Glutamate H11001 ATP 88n H9253-glutamyl phosphate H11001 ADP
(2) H9253-Glutamyl phosphate H11001 NH
4
H11001
88n glutamine H11001 P
i
H11001 H
H11001
Sum: Glutamate H11001 NH
4
H11001
H11001 ATP 88n
glutamine H11001 ADP H11001 Pi H11001 H
H11001
(22–1)
Glutamine synthetase is found in all organisms. In ad-
dition to its importance for NH
4
H11001
assimilation in bacte-
ria, it has a central role in amino acid metabolism in
mammals, converting toxic free NH
4
H11001
to glutamine for
transport in the blood (Chapter 18).
In bacteria and plants, glutamate is produced from
glutamine in a reaction catalyzed by glutamate syn-
thase. H9251-Ketoglutarate, an intermediate of the citric
acid cycle, undergoes reductive amination with gluta-
mine as nitrogen donor:
H9251-Ketoglutarate H11001 glutamine H11001 NADPH H11001 H
H11001
88n
2 glutamate H11001 NADP
H11001
(22–2)
The net reaction of glutamine synthetase and glutamate
synthase (Eqns 22–1 and 22–2) is
H9251-Ketoglutarate H11001 NH
4
H11001
H11001 NADPH H11001 ATP 88n
L-glutamate H11001 NADP
H11001
H11001 ADP H11001 P
i
Glutamate synthase is not present in animals, which, in-
stead, maintain high levels of glutamate by processes
such as the transamination of H9251-ketoglutarate during
amino acid catabolism.
Glutamate can also be formed in yet another, albeit
minor, pathway: the reaction of H9251-ketoglutarate and
NH
4
H11001
to form glutamate in one step. This is catalyzed by
L-glutamate dehydrogenase, an enzyme present in all or-
ganisms. Reducing power is furnished by NADPH:
H9251-Ketoglutarate H11001 NH
4
H11001
H11001 NADPH 88n
L-glutamate H11001 NADP
H11001
H11001 H
2
O
We encountered this reaction in the catabolism of amino
acids (see Fig. 18–7). In eukaryotic cells, L-glutamate
dehydrogenase is located in the mitochondrial matrix.
The reaction equilibrium favors reactants, and the K
m
for NH
4
H11001
(~1 mM) is so high that the reaction probably
makes only a modest contribution to NH
4
H11001
assimilation
into amino acids and other metabolites. (Recall that the
glutamate dehydrogenase reaction, in reverse (see Fig.
18–10), is one source of NH
4
H11001
destined for the urea cy-
cle.) Concentrations of NH
4
H11001
high enough for the gluta-
mate dehydrogenase reaction to make a significant con-
tribution to glutamate levels generally occur only when
NH
3
is added to the soil or when organisms are grown
in a laboratory in the presence of high NH
3
concentra-
tions. In general, soil bacteria and plants rely on the two-
enzyme pathway outlined above (Eqns 22–1, 22–2).
Glutamine Synthetase Is a Primary Regulatory Point
in Nitrogen Metabolism
The activity of glutamine synthetase is regulated in vir-
tually all organisms—not surprising, given its central
metabolic role as an entry point for reduced nitrogen.
In enteric bacteria such as E. coli, the regulation is un-
usually complex. The enzyme has 12 identical subunits
of M
r
50,000 (Fig. 22–5) and is regulated both alloster-
ically and by covalent modification. Alanine, glycine, and
at least six end products of glutamine metabolism are
allosteric inhibitors of the enzyme (Fig. 22–6). Each in-
hibitor alone produces only partial inhibition, but the ef-
fects of multiple inhibitors are more than additive, and
all eight together virtually shut down the enzyme. This
control mechanism provides a constant adjustment of
glutamine levels to match immediate metabolic re-
quirements.
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules838
FIGURE 22–5 Subunit structure of glutamine synthetase as deter-
mined by x-ray diffraction. (PDB ID 2GLS) (a) Side view. The 12 sub-
units are identical; they are differently colored to illustrate packing
and placement. (b) Top view, showing active sites (green).
(b)
(a)
8885d_c22_833-880 2/6/04 8:35 AM Page 838 mac76 mac76:385_reb:
Superimposed on the allosteric regulation is inhibi-
tion by adenylylation of (addition of AMP to) Tyr
397
, lo-
cated near the enzyme’s active site (Fig. 22–7). This co-
valent modification increases sensitivity to the allosteric
inhibitors, and activity decreases as more subunits are
adenylylated. Both adenylylation and deadenylylation
are promoted by adenylyltransferase (AT in Fig.
22–7), part of a complex enzymatic cascade that re-
sponds to levels of glutamine, H9251-ketoglutarate, ATP, and
P
i
. The activity of adenylyltransferase is modulated by
binding to a regulatory protein called P
II
, and the activ-
ity of P
II
, in turn, is regulated by covalent modification
(uridylylation), again at a Tyr residue. The adenylyl-
transferase complex with uridylylated P
II
(P
II
-UMP)
stimulates deadenylylation, whereas the same complex
22.1 Overview of Nitrogen Metabolism 839
Glutamate
ATP
ADP + P
i
NH
3
Glutamine CTP
Histidine
Glucosamine 6-phosphate
AMP
Tryptophan
Carbamoyl phosphate
Glycine
Alanine
glutamine
synthetase
FIGURE 22–6 Allosteric regulation of glutamine synthetase. The en-
zyme undergoes cumulative regulation by six end products of gluta-
mine metabolism. Alanine and glycine probably serve as indicators of
the general status of amino acid metabolism in the cell.
Enzyme OPCH
2
O
H11002
H
OH
Adenine
(a)
(b)
O
O
H
OH
H
H
O
AMP
Glutamine
synthetase
(inactive)
Glutamine
synthetase
(active)
Glutamate
ADP
+ P
i
NH
3
Glutamine
deadenylylation
adenylylation
PP
i
AT
P
II
P
i
ADP
AT
P
II UMP
P
II
P
II
AT
UMP
H
2
OPP
i
uridylylation
-KetoglutarateH9251
UT
P
i
Gln
ATP
AT
ATP
ATP
UTP
UMP
H17011 FIGURE 22–7 Second level of regulation of glutamine synthetase:
covalent modifications. (a) An adenylylated Tyr residue. (b) Cascade
leading to adenylylation (inactivation) of glutamine synthetase. AT rep-
resents adenylyltransferase; UT, uridylyltransferase. Details of this cas-
cade are discussed in the text.
8885d_c22_833-880 2/6/04 8:35 AM Page 839 mac76 mac76:385_reb:
with deuridylylated P
II
stimulates adenylylation of glut-
amine synthetase. Both uridylylation and deuridylyla-
tion of P
II
are brought about by a single enzyme, uridy-
lyltransferase. Uridylylation is inhibited by binding of
glutamine and P
i
to uridylyltransferase and is stimulated
by binding of H9251-ketoglutarate and ATP to P
II
.
The regulation does not stop there. The uridylylated
P
II
also mediates the activation of transcription of the
gene encoding glutamine synthetase, thus increasing
the cellular concentration of the enzyme; the deuridy-
lylated P
II
brings about a decrease in transcription of
the same gene. This mechanism involves an interaction
of P
II
with additional proteins involved in gene regula-
tion, of a type described in Chapter 28. The net result
of this elaborate system of controls is a decrease in glu-
tamine synthetase activity when glutamine levels are
high, and an increase in activity when glutamine levels
are low and H9251-ketoglutarate and ATP (substrates for the
synthetase reaction) are available. The multiple layers
of regulation permit a sensitive response in which glut-
amine synthesis is tailored to cellular needs.
Several Classes of Reactions Play Special Roles in
the Biosynthesis of Amino Acids and Nucleotides
The pathways described in this chapter include a vari-
ety of interesting chemical rearrangements. Several of
these recur and deserve special note before we progress
to the pathways themselves. These are (1) transamina-
tion reactions and other rearrangements promoted by
enzymes containing pyridoxal phosphate; (2) transfer
of one-carbon groups, with either tetrahydrofolate (usu-
ally at the OCHO and OCH
2
OH oxidation levels) or S-
adenosylmethionine (at the OCH
3
oxidation level) as
cofactor; and (3) transfer of amino groups derived from
the amide nitrogen of glutamine. Pyridoxal phosphate
(PLP), tetrahydrofolate (H
4
folate), and S-adenosylme-
thionine (adoMet) were described in some detail in
Chapter 18 (see Figs 18–6, 18–17, and 18–18). Here we
focus on amino group transfer involving the amide ni-
trogen of glutamine.
More than a dozen known biosynthetic reactions use
glutamine as the major physiological source of amino
groups, and most of these occur in the pathways out-
lined in this chapter. As a class, the enzymes catalyzing
these reactions are called glutamine amidotrans-
ferases. All have two structural domains: one binding
glutamine, the other binding the second substrate, which
serves as amino group acceptor (Fig. 22–8). A conserved
Cys residue in the glutamine-binding domain is believed
to act as a nucleophile, cleaving the amide bond of glu-
tamine and forming a covalent glutamyl-enzyme inter-
mediate. The NH
3
produced in this reaction is not re-
leased, but instead is transferred through an “ammonia
channel” to a second active site, where it reacts with
the second substrate to form the aminated product. The
covalent intermediate is hydrolyzed to the free enzyme
and glutamate. If the second substrate must be acti-
vated, the usual method is the use of ATP to generate
an acyl phosphate intermediate (ROOX in Fig. 22–8,
with X as a phosphoryl group). The enzyme glutaminase
acts in a similar fashion but uses H
2
O as the second sub-
strate, yielding NH
4
H11001
and glutamate (see Fig. 18–8).
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules840
H
3
N
Cys SH
CH
2
CH
2
H
+
NH
3
channel
Glutamine
amidotransferase
Glutamine-
binding domain
NH
3
-
acceptor domain
Glutamine
Acceptor
H
2
O
Glutamate
Activated
substrate
NH
2
CH
:
C
COO
–
OH orR
R
1
R
2
OC
+
O
Glutamyl-
enzyme
intermediate
1
2
H
3
N
Cys S
CH
2
CH
2
CH
C
COO
–
+
O
NH
3
NH
2
OX
or
R
R
1
R
R
2
+
+
OC
OX
or
H
R
1
R
2
NHCH
2
O
MECHANISM FIGURE 22–8 Proposed mechanism for glutamine
amidotransferases. Each enzyme has two domains. The glutamine-
binding domain contains structural elements conserved among many
of these enzymes, including a Cys residue required for activity. The
NH
3
-acceptor (second-substrate) domain varies. 1 The H9253-amido
nitrogen of glutamine (red) is released as NH
3
in a reaction that prob-
ably involves a covalent glutamyl-enzyme intermediate. The NH
3
trav-
els through a channel to the second active site, where 2 it reacts
with any of several acceptors. Two types of amino acceptors are shown.
X represents an activating group, typically a phosphoryl group derived
from ATP, that facilitates displacement of a hydroxyl group from
ROOH by NH
3
.
8885d_c22_833-880 2/6/04 8:35 AM Page 840 mac76 mac76:385_reb:
SUMMARY 22.1 Overview of Nitrogen Metabolism
■ The molecular nitrogen that makes up 80% of
the earth’s atmosphere is unavailable to most
living organisms until it is reduced. This fixation
of atmospheric N
2
takes place in certain free-
living bacteria and in symbiotic bacteria in the
root nodules of leguminous plants.
■ The nitrogen cycle entails formation of
ammonia by bacterial fixation of N
2
,
nitrification of ammonia to nitrate by soil
organisms, conversion of nitrate to ammonia by
higher plants, synthesis of amino acids from
ammonia by all organisms, and conversion of
nitrate to N
2
by denitrifying soil bacteria.
■ Fixation of N
2
as NH
3
is carried out by the
nitrogenase complex, in a reaction that
requires ATP. The nitrogenase complex is
highly labile in the presence of O
2
.
■ In living systems, reduced nitrogen is
incorporated first into amino acids and then
into a variety of other biomolecules, including
nucleotides. The key entry point is the amino
acid glutamate. Glutamate and glutamine are
the nitrogen donors in a wide variety of
biosynthetic reactions. Glutamine synthetase,
which catalyzes the formation of glutamine
from glutamate, is a main regulatory enzyme of
nitrogen metabolism.
■ The amino acid and nucleotide biosynthetic
pathways make repeated use of the biological
cofactors pyridoxal phosphate, tetrahydrofolate,
and S-adenosylmethionine. Pyridoxal phos-
phate is required for transamination reactions
involving glutamate and for other amino acid
transformations. One-carbon transfers require
S-adenosylmethionine and tetrahydrofolate.
Glutamine amidotransferases catalyze reactions
that incorporate nitrogen derived from
glutamine.
22.2 Biosynthesis of Amino Acids
All amino acids are derived from intermediates in gly-
colysis, the citric acid cycle, or the pentose phosphate
pathway (Fig. 22–9). Nitrogen enters these pathways by
way of glutamate and glutamine. Some pathways are
simple, others are not. Ten of the amino acids are just
one or several steps removed from the common metabo-
lite from which they are derived. The biosynthetic path-
ways for others, such as the aromatic amino acids, are
more complex.
Organisms vary greatly in their ability to synthesize
the 20 common amino acids. Whereas most bacteria and
plants can synthesize all 20, mammals can synthesize
only about half of them—generally those with simple
pathways. These are the nonessential amino acids,
not needed in the diet (see Table 18–1). The remain-
der, the essential amino acids, must be obtained from
food. Unless otherwise indicated, the pathways for the
20 common amino acids presented below are those op-
erative in bacteria.
22.2 Biosynthesis of Amino Acids 841
Histidine
Glutamine
Proline
Arginine
Asparagine
Methionine
Threonine
Lysine
Pyruvate
Tryptophan
Phenylalanine
Tyrosine
Phosphoenolpyruvate
Glycine
Cysteine
Serine
4 steps
4 steps
Alanine
Valine
Leucine
Isoleucine
Erythrose 4-
phosphate
3-Phosphoglycerate
Glucose 6-phosphate
Glucose
GlutamateAspartate
Oxaloacetate
Ribose 5-
phosphate
-KetoglutarateH9251
Citrate
FIGURE 22–9 Overview of amino acid biosynthesis. The carbon
skeleton precursors derive from three sources: glycolysis (pink), the
citric acid cycle (blue), and the pentose phosphate pathway (purple).
8885d_c22_833-880 2/6/04 8:35 AM Page 841 mac76 mac76:385_reb:
A useful way to organize these biosynthetic path-
ways is to group them into six families corresponding
to their metabolic precursors (Table 22–1), and we use
this approach to structure the detailed descriptions
that follow. In addition to these six precursors, there
is a notable intermediate in several pathways of amino
acid and nucleotide synthesis—5-phosphoribosyl-1-
pyrophosphate (PRPP):
PRPP is synthesized from ribose 5-phosphate derived
from the pentose phosphate pathway (see Fig. 14–21),
in a reaction catalyzed by ribose phosphate pyro-
phosphokinase:
Ribose 5-phosphate H11001 ATP 88n
5-phosphoribosyl-1-pyrophosphate H11001 AMP
This enzyme is allosterically regulated by many of the
biomolecules for which PRPP is a precursor.
H9251-Ketoglutarate Gives Rise to Glutamate, Glutamine,
Proline, and Arginine
We have already described the biosynthesis of gluta-
mate and glutamine. Proline is a cyclized derivative
of glutamate (Fig. 22–10). In the first step of proline
synthesis, ATP reacts with the H9253-carboxyl group of glu-
tamate to form an acyl phosphate, which is reduced
by NADPH or NADH to glutamate H9253-semialdehyde. This
intermediate undergoes rapid spontaneous cyclization
and is then reduced further to yield proline.
Arginine is synthesized from glutamate via or-
nithine and the urea cycle in animals (Chapter 18). In
principle, ornithine could also be synthesized from glu-
tamate H9253-semialdehyde by transamination, but the spon-
taneous cyclization of the semialdehyde in the proline
pathway precludes a sufficient supply of this intermedi-
ate for ornithine synthesis. Bacteria have a de novo
biosynthetic pathway for ornithine (and thus arginine)
that parallels some steps of the proline pathway but in-
cludes two additional steps that avoid the problem of the
spontaneous cyclization of glutamate H9253-semialdehyde
(Fig. 22–10). In the first step, the H9251-amino group of glu-
tamate is blocked by an acetylation requiring acetyl-CoA;
Glutamate
Glutamine Proline Arginine
H9251-Ketoglutarate
O
P
OH
CH
2
O
H11002
H11002
O
H
OH
H
H
H
O
O
O
O
PPO
O
O
H11002
O
H11002
O
H11002
then, after the transamination step, the acetyl group is
removed to yield ornithine.
The pathways to proline and arginine are somewhat
different in mammals. Proline can be synthesized by the
pathway shown in Figure 22–10, but it is also formed
from arginine obtained from dietary or tissue protein.
Arginase, a urea cycle enzyme, converts arginine to orni-
thine and urea (see Figs 18–10, 18–26). The ornithine is
converted to glutamate H9253-semialdehyde by the enzyme
ornithine H9254-aminotransferase (Fig. 22–11). The semi-
aldehyde cyclizes to H9004
1
-pyrroline-5-carboxylate, which
is then converted to proline (Fig. 22–10). The pathway
for arginine synthesis shown in Figure 22–10 is absent
in mammals. When arginine from dietary intake or pro-
tein turnover is insufficient for protein synthesis, the
ornithine H9254-aminotransferase reaction operates in the
direction of ornithine formation. Ornithine is then con-
verted to citrulline and arginine in the urea cycle.
Serine, Glycine, and Cysteine Are Derived
from 3-Phosphoglycerate
The major pathway for the formation of serine is the
same in all organisms (Fig. 22–12). In the first step, the
hydroxyl group of 3-phosphoglycerate is oxidized by a
3-Phosphoglycerate
Serine
CysteineGlycine
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules842
TABLE 22–1
H9251-Ketoglutarate
Glutamate
Glutamine
Proline
Arginine
3-Phosphoglycerate
Serine
Glycine
Cysteine
Oxaloacetate
Aspartate
Asparagine
Methionine*
Threonine*
Lysine*
Pyruvate
Alanine
Valine*
Leucine*
Isoleucine*
Phosphoenolpyruvate and
erythrose 4-phosphate
Tryptophan*
Phenylalanine*
Tyrosine
?
Ribose 5-phosphate
Histidine*
Amino Acid Biosynthetic Families,
Grouped by Metabolic Precursor
*Essential amino acids.
?
Derived from phenylalanine in mammals.
8885d_c22_833-880 2/6/04 8:35 AM Page 842 mac76 mac76:385_reb:
ADP
A
O
D
B
M
O
C
H11002
O
CH
2
COO
H11002
CH
S-CoA CoA-SH
OO
O
O
CH
3
O
O
C
CH
2
acetylglutamate synthase
Glutamate
N-acetylglutamate
dehydrogenase
A
O
NH
3
COO
H11002
CHOOCH
2
O
B
O
CH
3
C
A
O
D
M
CH
2
COO
H11002
CH
H11001
O
O
H
2
C
O
O
C
CH
2
O
pyrroline carboxylate
reductase
NAD(P)
H11001
A
O
D
M
H
CH
2
COO
H11002
CHOO
NH
3
O
CCH
2
O
N-Acetylglutamate
B
O
HN OCH
3
CO
N-acetylglutamate
kinase
glutamate kinase
N-Acetyl-H9253-glutamyl
phosphate
H9253-Glutamyl
phosphate
NAD(P)H
NAD(P)
H11001
H11001 H
H11001
B
O
HN OCH
3
C
A
O
D
M
O
CH
2
COO
H11002
CHOO
O
O
CCH
2
O
O
P
A
O
D
M
O
CH
2
COO
H11002
CH
H11001
OO
NH
3
O
O
CCH
2
O
Glutamate H9253-semialdehyde
(P5C)
nonenzymatic
N-Acetylornithine
H
2
O
CH
3
COO
H11002
N-acetylornithinase
H9004
1
-Pyrroline-5-carboxylate
COO
H11002
H11002
O
OCH
2
H11001
NH
3
H
C
N
H
2
CH
COO
H11002
H11001
O
H
2
C CH
2
HC
N
H
O CH
glutamate
dehydrogenase
P
i
H
f
H
B
O
HN OCH
3
C
A
O
D
M
H
CH
2
COO
H11002
CHOO
O
CCH
2
O
O
N-Acetylglutamate
H9253-semialdehyde
Glutamate
H9251-Ketoglutarate
aminotransferase
OCH
2
H
2
N
H11001
OHN
A
O CH
2
COO
H11002
CHOOCH
2
O
H11001
OCH
2
H
3
N
H11001
CH
2
Ornithine
A
O
NH
3
COO
H11002
CHOOCH
2
O
H11001
OCH
2
H
3
N
H11001
CH
2
G
JH11001
H
2
N
N
H
OC
Proline
Arginine
ATP
AMP H11001 PP
i
L-Citrulline
argininosuccinate
synthetase
ATP H11001 aspartate
P
i
ornithine
carbamoyl-
transferase
Carbamoyl phosphate
P
NAD(P)H H11001 H
H11001
ADP
NAD(P)
H11001
NAD(P)H H11001 H
H11001
P
i
ATP
Fumarate
Argininosuccinate
argininosuccinase
Urea cycle
FIGURE 22–10 Biosynthesis of proline and arginine from glutamate
in bacteria. All five carbon atoms of proline arise from glutamate. In
many organisms, glutamate dehydrogenase is unusual in that it uses
either NADH or NADPH as a cofactor. The same may be true of other
enzymes in these pathways. The H9253-semialdehyde in the proline path-
way undergoes a rapid, reversible cyclization to H9004
1
-pyrroline-5-
carboxylate (P5C), with the equilibrium favoring P5C formation.
Cyclization is averted in the ornithine/arginine pathway by acetylation
of the H9251-amino group of glutamate in the first step and removal of the
acetyl group after the transamination. Although some bacteria lack
arginase and thus the complete urea cycle, they can synthesize argi-
nine from ornithine in steps that parallel the mammalian urea cycle,
with citrulline and argininosuccinate as intermediates (see Fig. 18–10).
Here, and in subsequent figures in this chapter, the reaction ar-
rows indicate the linear path to the final products, without consider-
ing the reversibility of individual steps. For example, the second step
of the pathway leading to arginine, catalyzed by N-acetylglutamate
dehydrogenase, is chemically similar to the glyceraldehyde 3-phos-
phate dehydrogenase reaction in glycolysis (see Fig. 14–7) and is read-
ily reversible.
843
8885d_c22_843 2/6/04 1:08 PM Page 843 mac76 mac76:385_reb:
dehydrogenase (using NAD
H11001
) to yield 3-phosphohy-
droxypyruvate. Transamination from glutamate yields 3-
phosphoserine, which is hydrolyzed to free serine by
phosphoserine phosphatase.
Serine (three carbons) is the precursor of glycine
(two carbons) through removal of a carbon atom by
serine hydroxymethyltransferase (Fig. 22–12).
Tetrahydrofolate accepts the H9252 carbon (C-3) of serine,
which forms a methylene bridge between N-5 and N-10
to yield N
5
,N
10
-methylenetetrahydrofolate (see Fig.
18–17). The overall reaction, which is reversible, also
requires pyridoxal phosphate. In the liver of verte-
brates, glycine can be made by another route: the re-
verse of the reaction shown in Figure 18–20c, cat-
alyzed by glycine synthase (also called glycine
cleavage enzyme):
CO
2
H11001 NH
4
H11001
H11001 N
5
,N
10
-methylenetetrahydrofolate H11001
NADH H11001 H
H11001
88n
glycine H11001 tetrahydrofolate H11001 NAD
H11001
Plants and bacteria produce the reduced sulfur re-
quired for the synthesis of cysteine (and methionine,
described later) from environmental sulfates; the path-
way is shown on the right side of Figure 22–13. Sulfate
is activated in two steps to produce 3-phosphoadeno-
sine 5H11032-phosphosulfate (PAPS), which undergoes an
eight-electron reduction to sulfide. The sulfide is then
used in formation of cysteine from serine in a two-step
pathway. Mammals synthesize cysteine from two amino
acids: methionine furnishes the sulfur atom and serine
furnishes the carbon skeleton. Methionine is first con-
verted to S-adenosylmethionine (see Fig. 18–18), which
can lose its methyl group to any of a number of accep-
tors to form S-adenosylhomocysteine (adoHcy). This
demethylated product is hydrolyzed to free homocys-
teine, which undergoes a reaction with serine, catalyzed
by cystathionine H9252-synthase, to yield cystathionine
(Fig. 22–14). Finally, cystathionine H9253-lyase, a PLP-
requiring enzyme, catalyzes removal of ammonia and
cleavage of cystathionine to yield free cysteine.
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules844
Ornithine
COO
H11002
CHH
3
N
CH
2
CH
2
CH
2
H11001
NH
3
H11001
ornithine
H9254-aminotransferase
-KetoglutarateH9251
Glutamate
COO
H11002
CHH
3
N
CH
2
CH
2
C
H
O
H11001
Glutamate
H9253-semialdehyde
H9004
1
-Pyrroline-5-
carboxylate
(P5C)
H
2
C
H
2
O
H
2
O
CH
2
CH CH COO
H11002
N
H
H11001
FIGURE 22–11 Ornithine H9254-aminotransferase
reaction: a step in the mammalian pathway
to proline. This enzyme is found in the
mitochondrial matrix of most tissues.
Although the equilibrium favors P5C
formation, the reverse reaction is the only
mammalian pathway for synthesis of ornithine
(and thus arginine) when arginine levels are
insufficient for protein synthesis.
FIGURE 22–12 Biosynthesis of serine from 3-phosphoglycerate and
of glycine from serine in all organisms. Glycine is also made from
CO
2
and NH
4
H11001
by the action of glycine synthase, with N
5
,N
10
-methy-
lenetetrahydrofolate as methyl group donor (see text).
COO
H11002
H
C
OH
O
H11001
CH
2
H
H
3
N
3-Phosphoglycerate
3-Phosphohydroxypyruvate
3-Phosphoserine
Glycine
phosphoglycerate
dehydrogenase
phosphoserine
aminotransferase
phosphoserine
phosphatase
serine
hydroxymethyl-
transferase
PLP
H
4
folate
N
5
,N
10
-Methylene H
4
folate
A
O
A
C
A
H O
OOH
H
O
O
COO
H11002
C
A
A
P
O
O
COO
H11002
C
CH
2
NAD
H11001
NADH H11001 H
H11001
A
A
P
OO
O
O
COO
H11002
C
H11001
H
3
N
Glutamate
H9251-Ketoglutarate
A
A
HOO
COO
H11002
C
H11001
CH
2
OH
H
3
N
P
i
Serine
A
A
HOO
H
2
O
H
2
O
PO
PO
8885d_c22_833-880 2/6/04 8:35 AM Page 844 mac76 mac76:385_reb:
Three Nonessential and Six Essential Amino Acids
Are Synthesized from Oxaloacetate and Pyruvate
Alanine and aspartate are synthesized from pyruvate
and oxaloacetate, respectively, by transamination from
glutamate. Asparagine is synthesized by amidation of
aspartate, with glutamine donating the NH
4
H11001
. These are
nonessential amino acids, and their simple biosynthetic
pathways occur in all organisms.
Methionine, threonine, lysine, isoleucine, valine, and
leucine are essential amino acids. Their biosynthetic
pathways are complex and interconnected (Fig. 22–15).
Aspartate
Asparagine Methionine ThreonineLysine
Oxaloacetate
Isoleucine
Pyruvate
Alanine Valine Leucine
22.2 Biosynthesis of Amino Acids 845
ATP SO
4
2H11002
H11001
H9007
H11001
PP
i
ADP
ATP
CH
3
COO
H11002
S
2H11002
H11001 H
H11001
S
2H11002
3H11032-Phosphoadenosine 5H11032-phosphate (PAP)
NADPH
NADP
H11001
ATP sulfurylase
3NADP
H11001
3NADPH
sulfide reductase
PAPS reductase
H11002
OSO
O
O
H11002
P O O
HH
H
OH OH
H
CH
2
O
O
H11002
Adenine
Adenosine
5H11032-phosphosulfate (APS)
H11002
OSO
O
O
H11002
P O O
HH
H
O
SO
3
2H11002
P
OH
H
CH
2
H11002
O O
O
O
H11002
O
H11002
Adenine
3H11032-Phosphoadenosine
5H11032-phosphosulfate (PAPS)
Sulfite
Sulfide
COO
H11002
CH
2
OH
CHH
3
N
H11001
Serine
COO
H11002
CH
2
SH
CHH
3
N
H11001
Cysteine
COO
H11002
CH
2
O
CHH
3
N
CH
3
CO
H11001
O-Acetylserine
CoA-SH
H
3
CC
serine
acetyltransferase
O-acetylserine
(thiol) lyase
S-CoA
O
FIGURE 22–13 Biosynthesis of cysteine from serine in
bacteria and plants. The origin of reduced sulfur is shown
in the pathway on the right.
H11002
OOC
HOCH
2
H11001
NH
3
H
2
O
COO
H11002
CH
2
SH
HSC
O
CH
CHCH
2
CH
2
S
H11001
H11001
NH
3
Homocysteine Serine
NH
4
H
2
O
H11002
OOC
H11001
NH
3
COO
H11002
CH
2
CH CHCH
2
Cystathionine
PLP
CH
3
H11002
OOC
H11001
NH
3
COO
H11002
CHCH
2
CH
2
H11001
H9251-Ketobutyrate Cysteine
cystathionine H9252-synthase
cystathionine H9253-lyase
H11001
H11001
NH
3
PLP
FIGURE 22–14 Biosynthesis of cysteine from homocysteine and ser-
ine in mammals. The homocysteine is formed from methionine, as de-
scribed in the text.
8885d_c22_833-880 2/6/04 8:35 AM Page 845 mac76 mac76:385_reb:
O
C
OH
CH
NH
2
CH
2
H11002
OOC
NH
3
H11001
NADP
H11001
NADPH
H11001 H
H11001
O
H11002
O
NH
3
H11001
Threonine
O
CHCH
3
COO
H11002
P
NH
3
H11001
ATP
Aspartyl-H9252-phosphate
O
CH
2
C
H
C CH
2
CH COO
H11002
NH
3
H11001
H
COO
H11002H11002
OOC
Aspartate H9252-semialdehyde
NADP
NADPH H11001 H
H11001
H
COO
H11002H11002
OOC N
Dihydropicolinate
H
H
10
8
Pyruvate
Pyruvate H11001 NH
3
COO
H11002H11002
OOC N
H9004
1
-Piperidine-2,6-
dicarboxylate
Lysine
NH
H
3
N
15
COO
H11002H11002
OOC O
N-Succinyl-2-amino-
6-keto-L-pimelate
NH
Succinate
H9251-Ketoglutarate Glutamate
NH
3
H11001
11
PLP
COO
H11002
COO
H11002
10 H
2
O
C
H
H
H11001
N-Succinyl-L, L-
H9251, H9280-diamino-
pimelate
(CH
2
)
3
COO
H11002
NH
3
H11001
L, L-H9251, H9280-Diamino-
pimelate
H
3
N
COO
H11002
C
C
H
H
H11001
COO
H11002
meso-H9251, H9280-Diamino-
pimelate
H
3
N
COO
H11002
C
C
H
H
H11001
H
3
N
H11001
H
H11001
CH
2
CO
2
13
16
CAspartate
CCHCH
2
COO
H11002
NH
3
H11001
4
ADP
CHCH
2
COO
H11002
NH
3
H11001
CHCH
H
2
O
PLP
CH
2
P
Methionine
ATP
1
ADP
S
CH COO
H11002
NH
3
H11001
CH
2
OH
CH
2
COO
H11002
NH
3
H11001
CH
2
HomoserineCH
12
Succinyl-CoA H11001
CoA
Succinate
6
Succinyl-CoA
CoA
CHCH
2
COO
H11002
Succinate
CH
2
O-Succinylhomoserine
O
H
2
O
H
COO
H11002
NH
3
H11001
C
CHCH
2
9
CH
2
CystathionineSH
2
C COO
H11002
NH
3
H11001
Cysteine
7
Succinate
CHCH
2
CH
2
HS COO
H11002
NH
3
H11001
Homocysteine
CH
2
3
COO
H11002
O CH
2
Phosphohomoserine
CH
3
OH
NH
4
H11001
H11001
2
14
Succinate
H
2
O
N
5
-Methyl H
4
folate
H
4
folate
NH
3
H11001
5
P
i
H
2
O
PLP
P
i
NADPH H11001 H
H11001
NADP
H11001
O
(CH
2
)
3
(CH
2
)
3
17
PLP
PLP
PLP
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules846
FIGURE 22–15 Biosynthesis of six essential amino acids from oxalo-
acetate and pyruvate in bacteria: methionine, threonine, lysine, iso-
leucine, valine, and leucine. Here, and in other multistep pathways,
the enzymes are listed in the key. Note that L,L-H9251,H9280-diaminopimelate,
the product of step 14 , is symmetric. The carbons derived from pyru-
vate (and the amino group derived from glutamate) are not traced
beyond this point, because subsequent reactions may place them at
either end of the lysine molecule.
8885d_c22_833-880 2/6/04 8:35 AM Page 846 mac76 mac76:385_reb:
25
22
Glutamate
CH
3
COO
H11002
CO
2
OH
NH
3
H11001
H9251-Ketoglutarate
CH C
Pyruvate
18
CH
3
H11002
Leucine
O
C
CH
3
C
CH
2
COO
H11002
CH
CO
2
NADH H11001 H
H11001
NAD
H11001
19
OH
H9251-Acetolactate
CH
3
CCH
3
COO
H11002
O
C
OH
H9252-Isopropylmalate
CH
3
CH
3
O
H9251-Isopropylmalate
CH
3
COO
H11002
21
19
Glutamate
HO
CH
3
COO
H11002
H
2
O
NAD(P)H H11001 H
H11001
NAD(P)
H11001
CH
3
O
NH
3
H11001
H9251-Ketoglutarate
C
CH
2
C
O
COO
H11002
CH
2
C
CH
3
H9251-Ketobutyrate
18
Isoleucine
CH
2
CH
CH
3
CH
3
COO
H11002
CH
19
OH
H9251-Aceto-H9251-
hydroxybutyrate
CH
2
C
CH
3
CH
3
COO
H11002
O
C
H
H9251, H9252-Dihydroxy-
H9252-methylvalerate
CH
2
C
CH
3
CH
3
COO
H11002
OH
C
20
O
H9251-Keto-H9252-
methylvalerate
CH
2
C
CH
3
CH
3
COO
H11002
H
C
C
O
COO
H11002
CH
3
CH
3
TPP
OH
C
PLP
24
Pyruvate
PLP
21
19
Glutamate
CH
3
COO
H11002
NAD(P)H H11001 H
H11001
NAD(P)
H11001
O
NH
3
H11001
H9251-Ketoglutarate
C
CH
3
C
Valine
CH
3
CHCH
3
COO
H11002
CH
OH
H9251, H9252-Dihydroxy-
isovalerate
CH
3
CCH
3
COO
H11002
OH
C
O
H9251-Keto-
isovalerate
CH
3
CCH
3
COO
H11002
H
C
20
H
2
O
PLP
CH
3
CH
CH
3
CH
2
CH
CH
CH CH
COO
H11002
COO
H11002
CH
2
CH
3
COO
H11002
H9251-Ketoisocaproate
23
CoA
Acetyl -CoA
COO
H11002
OH
H
OH
TPP
18
aspartokinase
aspartate H9252-semialdehyde dehydrogenase
homoserine dehydrogenase
homoserine kinase
threonine synthase
homoserine acyltransferase
cystathionine H9253-synthase
cystathionine H9252-lyase
methionine synthase
dihydropicolinate synthase
H9004
1
-piperidine-2,6-dicarboxylate dehydrogenase
N-succinyl-2-amino-6-ketopimelate synthase
succinyl diaminopimelate aminotransferase
succinyl diaminopimelate desuccinylase
diaminopimelate epimerase
diaminopimelate decarboxylase
threonine dehydratase (serine dehydratase)
acetolactate synthase
acetohydroxy acid isomeroreductase
dihydroxy acid dehydratase
valine aminotransferase
H9251-isopropylmalate synthase
isopropylmalate isomerase
H9252-isopropylmalate dehydrogenase
leucine aminotransferase
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
22.2 Biosynthesis of Amino Acids 847
8885d_c22_833-880 2/6/04 8:35 AM Page 847 mac76 mac76:385_reb:
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules848
O
C
CH
2
H
2
O
COO
H11002
CH
P
CH
2
O
3
O
C
CH
2
CHOH
COO
H11002
H11001
O
C
H
P
CHOH
CH
2
O
7
H
H
HO
OH
OH
C
C
H
NAD
H11001
P
i
2
HO
C
H
OH
COO
H11002
H
HO
Phosphoenolpyruvate
(PEP)
Erythrose 4-phosphate
2-Keto-3-deoxy-D-
arabinoheptulosonate
7-phosphate
3-Dehydroquinate
H
2
O
P
i
1
H
OH
H
COO
H11002
HO H
NADP
H11001
NADPH H11001 H
H11001
4
HO
H
OH
H
COO
H11002
HO H
ADP
5
H
OH
H
COO
H11002
HO H
PEP
P
i
6
HH
COO
H11002
HO H
P
i
O
O
COO
H11002
H
2
C
C
O
CH
2
O C COO
H11002
Shikimate
Shikimate
3-phosphate
5-Enolpyruvylshikimate
3-phosphate
Chorismate
3-Dehydroshikimate
O
O
HO
3-Dehydroshikimate
ATP
P
P
P
COO
H11002
2-keto-3-deoxy-D-arabinoheptulosonate 7-phosphate
synthase
dehydroquinate synthase
3-dehydroquinate dehydratase
shikimate dehydrogenase
shikimate kinase
5-enolpyruvylshikimate 3-phosphate synthase
chorismate synthase
1
2
3
4
5
6
7
to valine and isoleucine in pathways that begin with con-
densation of two carbons of pyruvate (in the form of
hydroxyethyl thiamine pyrophosphate; see Fig. 14–13)
with another molecule of pyruvate (valine path) or with
H9251-ketobutyrate (isoleucine path). The H9251-ketobutyrate is
derived from threonine in a reaction that requires pyri-
doxal phosphate (Fig. 22–15, step 17 ). An intermedi-
ate in the valine pathway, H9251-ketoisovalerate, is the start-
ing point for a four-step branch pathway leading to
leucine (steps 22 to 25 ).
FIGURE 22–16 Biosynthesis of chorismate, an intermediate in
the synthesis of aromatic amino acids in bacteria and plants.
All carbons are derived from either erythrose 4-phosphate
(light purple) or phosphoenolpyruvate (pink). Note that the
NAD
H11001
required as a cofactor in step 2 is released
unchanged; it may be transiently reduced to NADH during the
reaction, with formation of an oxidized reaction intermediate.
Step 6 is competitively inhibited by glyphosate
(
H11002
COOOCH
2
ONHOCH
2
OPO
3
2H11002
), the active ingredient in
the widely used herbicide Roundup. The herbicide is relatively
nontoxic to mammals, which lack this biosynthetic pathway.
The chemical names quinate, shikimate, and chorismate are
derived from the names of plants in which these intermediates
have been found to accumulate.
In some cases, the pathways in bacteria, fungi, and
plants differ significantly. The bacterial pathways are
outlined in Figure 22–15.
Aspartate gives rise to methionine, threonine,
and lysine. Branch points occur at aspartate H9252-semi-
aldehyde, an intermediate in all three pathways, and at
homoserine, a precursor of threonine and methionine.
Threonine, in turn, is one of the precursors of isoleucine.
The valine and isoleucine pathways share four en-
zymes (Fig. 22–15, steps 18 to 21). Pyruvate gives rise
8885d_c22_848 2/6/04 1:08 PM Page 848 mac76 mac76:385_reb:
22.2 Biosynthesis of Amino Acids 849
OH
PLP
O
B
COO
H11002
A
H11001
H
2
O
Serine
PP
i
!
'
HO
H
H
COO
H11002
O
C
PRPP
O
O
Glutamine
Glutamate
Pyruvate
COO
H11002
CH
2
H
2
O H11001 CO
2
5
1
NH
2
H
HH
H
N
H
CH
2
O
O
O
O P
COO
H11002
HO
CH
CH
O
HO
OCOC
O
O
CH
2
O
O
P
A
A
B
H
A
A
A
OH
OH
COO
H11002
H
N
O POOCH
2
O
O
CH
A
OH
OH
N
H
NH
3
Glyceraldehyde 3-phosphate
N
H
CH
2
O
CH
COO
H11002
Tryptophan
Indole-3-glycerol phosphate
Enol-1-o-carboxyphenylamino-1-
deoxyribulose phosphate
N-(5H11032-Phosphoribosyl)-
anthranilate
Anthranilate
Chorismate
2
C
G
H
3
4
e
merase
anthranilate synthase
anthranilate phosphoribosyltransferase
N-(5'-phosphoribosyl)-anthranilate isomerase
indole-3-glycerol phosphate synthase
tryptophan synthase
1
2
3
4
5
FIGURE 22–17 Biosynthesis of tryptophan from chorismate in bac-
teria and plants. In E. coli, enzymes catalyzing steps 1 and 2 are
subunits of a single complex.
Chorismate Is a Key Intermediate in the Synthesis
of Tryptophan, Phenylalanine, and Tyrosine
Aromatic rings are not readily available in the environ-
ment, even though the benzene ring is very stable. The
branched pathway to tryptophan, phenylalanine, and ty-
rosine, occurring in bacteria, fungi, and plants, is the
main biological route of aromatic ring formation. It pro-
ceeds through ring closure of an aliphatic precursor fol-
lowed by stepwise addition of double bonds. The first
four steps produce shikimate, a seven-carbon molecule
derived from erythrose 4-phosphate and phospho-
enolpyruvate (Fig. 22–16). Shikimate is converted to
chorismate in three steps that include the addition of
three more carbons from another molecule of phospho-
enolpyruvate. Chorismate is the first branch point of the
pathway, with one branch leading to tryptophan, the
other to phenylalanine and tyrosine.
In the tryptophan branch (Fig. 22–17), chorismate
is converted to anthranilate in a reaction in which glu-
tamine donates the nitrogen that will become part of the
indole ring. Anthranilate then condenses with PRPP.
The indole ring of tryptophan is derived from the ring
carbons and amino group of anthranilate plus two car-
bons derived from PRPP. The final reaction in the se-
quence is catalyzed by tryptophan synthase. This en-
zyme has an H9251
2
H9252
2
subunit structure and can be
dissociated into two H9251 subunits and a H9252
2
subunit that
catalyze different parts of the overall reaction:
Indole-3-glycerol phosphate 88888n
H9251 subunit
indole H11001 glyceraldehyde 3-phosphate
Indole H11001 serine 888888n tryptophan H11001 H
2
O
H9252
2
subunit
Phosphoenolpyruvate
Erythrose 4-phosphate
Phenylalanine Tyrosine Tryptophan
Tyrosine
H11001
8885d_c22_833-880 2/6/04 8:35 AM Page 849 mac76 mac76:385_reb:
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules850
CH
3
O
COO
H11002
P
H11001
NH
H11001
NH
H11001
NH
H
Glyceraldehyde
3-phosphate
N
H
CH CH CH
2
P
Indole-3-glycerol phosphate
O
H
C
OH
NH
3
CH
CH
Indole
H11001
N
H
COO
H11002
C
PLP
CH
2
P
O
O
C
H
O
H
H
2
O
CH
2
PLP
H9251
tryptophan
synthase
H9251 subunits
H9251
Quinonoid intermediate
CH
3
COO
H11002
P HO
C
H11001
H9252
N
H
CH
2
C
H
B
O
H
Indole
H
2
C
H9252
Aldimine with tryptophan
CH
2
H11001
N
H
N
H
H
N
H
CH
3
P HO
C
N
H
CH
2
O
H
H9252
PLP-aminoacrylate
adduct
H11001
COO
H11002
N
H
CH
2
C
H9252
H11001
H
OH Serine
NH
3
CH
H11001
COO
H11002
H9252
CH
2
H
2
O
N
H
Tryptophan
CH
2
tryptophan
synthase
H9252 subunits
Enzyme
OH
OH
1
2
3
5
4
B HB
H9252H9252
MECHANISM FIGURE 22–18 Tryptophan synthase reaction. This en-
zyme catalyzes a multistep reaction with several types of chemical re-
arrangements. 1 An aldol cleavage produces indole and glyceralde-
hyde 3-phosphate; this reaction does not require PLP. 2 Dehydration
of serine forms a PLP-aminoacrylate intermediate. In steps 3 and 4
this condenses with indole, and 5 the product is hydrolyzed to re-
lease tryptophan. These PLP-facilitated transformations occur at the H9252
carbon (C-3) of the amino acid, as opposed to the H9251-carbon reactions
described in Figure 18–6. The H9252 carbon of serine is attached to the in-
dole ring system. Tryptophan Synthase Mechanism
The second part of the reaction requires pyridoxal phos-
phate (Fig. 22–18). Indole formed in the first part is not
released by the enzyme, but instead moves through a
channel from the H9251-subunit active site to the H9252-subunit
active site, where it condenses with a Schiff base inter-
mediate derived from serine and PLP. Intermediate
channeling of this type may be a feature of the entire
pathway from chorismate to tryptophan. Enzyme active
sites catalyzing different steps (sometimes not sequen-
tial steps) of the pathway to tryptophan are found on
single polypeptides in some species of fungi and bacte-
ria, but are separate proteins in others. In addition, the
activity of some of these enzymes requires a noncova-
lent association with other enzymes of the pathway.
These observations suggest that all the pathway en-
zymes are components of a large, multienzyme complex
in both prokaryotes and eukaryotes. Such complexes are
generally not preserved intact when the enzymes are
isolated using traditional biochemical methods, but ev-
idence for the existence of multienzyme complexes is
accumulating for this and a number of other metabolic
pathways (p. 605).
8885d_c22_833-880 2/6/04 8:35 AM Page 850 mac76 mac76:385_reb:
(see Figs 18–23, 18–24). Tyrosine is considered a con-
ditionally essential amino acid, or as nonessential inso-
far as it can be synthesized from the essential amino
acid phenylalanine.
Histidine Biosynthesis Uses Precursors
of Purine Biosynthesis
The pathway to histidine in all plants and bacteria dif-
fers in several respects from other amino acid biosyn-
thetic pathways. Histidine is derived from three pre-
cursors (Fig. 22–20): PRPP contributes five carbons, the
purine ring of ATP contributes a nitrogen and a carbon,
and glutamine supplies the second ring nitrogen. The
key steps are condensation of ATP and PRPP, in which
N-1 of the purine ring is linked to the activated C-1 of
the ribose of PRPP (step 1 in Fig. 22–20); purine ring
opening that ultimately leaves N-1 and C-2 of adenine
linked to the ribose (step 3 ); and formation of the im-
idazole ring, a reaction in which glutamine donates a ni-
trogen (step 5 ). The use of ATP as a metabolite rather
than a high-energy cofactor is unusual— but not waste-
ful, because it dovetails with the purine biosynthetic
pathway. The remnant of ATP that is released after the
transfer of N-1 and C-2 is 5-aminoimidazole-4-carbox-
amide ribonucleotide (AICAR), an intermediate of
purine biosynthesis (see Fig. 22–33) that is rapidly re-
cycled to ATP.
Amino Acid Biosynthesis Is under
Allosteric Regulation
The most responsive regulation of amino acid synthesis
takes place through feedback inhibition of the first re-
action in a sequence by the end product of the pathway.
This first reaction is usually irreversible and catalyzed
by an allosteric enzyme. As an example, Figure 22–21
shows the allosteric regulation of isoleucine synthesis
from threonine (detailed in Fig. 22–15). The end prod-
uct, isoleucine, is an allosteric inhibitor of the first
reaction in the sequence. In bacteria, such allosteric
modulation of amino acid synthesis occurs as a minute-
to-minute response.
Allosteric regulation can be considerably more com-
plex. An example is the remarkable set of allosteric con-
trols exerted on glutamine synthetase of E. coli (Fig.
22–6). Six products derived from glutamine serve as
negative feedback modulators of the enzyme, and the
overall effects of these and other modulators are more
than additive. Such regulation is called concerted in-
hibition.
Ribose 5-phosphate
Histidine
22.2 Biosynthesis of Amino Acids 851
FIGURE 22–19 Biosynthesis of phenylalanine and tyrosine from cho-
rismate in bacteria and plants. Conversion of chorismate to prephen-
ate is a rare biological example of a Claisen rearrangement.
Phenylpyruvate
O
HO H
COO
H11002
amino-
transferase
2
CH
2
H
CO
2
H11001 OH
H11002
O
COO
H11002
Prephenate
H11002
OOC
1
COO
H11002
H9251-Ketoglutarate
C
HO H Chorismate
NADH H11001 H
H11001
NAD
H11001
C
CH
2
COO
H11002
O
C
CH
COO
H11002
O
C
4-Hydroxyphenyl-
pyruvate
3
CH
2
Phenylalanine
CO
2
COO
H11002
CH
2
CH
2
Glutamate
OH
Tyrosine
amino-
transferase
COO
H11002
H9251-Ketoglutarate
CHCH
2
OH
Glutamate
NH
3
H11001H11001
NH
3
chorismate mutase
prephenate dehydrogenase
prephenate dehydratase
1
2
3
In plants and bacteria, phenylalanine and tyro-
sine are synthesized from chorismate in pathways much
less complex than the tryptophan pathway. The com-
mon intermediate is prephenate (Fig. 22–19). The final
step in both cases is transamination with glutamate.
Animals can produce tyrosine directly from phenyl-
alanine through hydroxylation at C-4 of the phenyl
group by phenylalanine hydroxylase; this enzyme
also participates in the degradation of phenylalanine
8885d_c22_851 2/6/04 1:08 PM Page 851 mac76 mac76:385_reb:
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules852
H
C
O
O
HC
HHC
OH
C
H
O
P
O
H11001
5-Phosphoribosyl-
1-pyrophosphate (PRPP)
N
HC
N
N
NH
2
N
P
ATP
5
PP
i
P P
P
H
C
O
CH
2
O
HC
H
H
C
C
O
P
H
PRib P P
5-Aminoimidazole-
4-carboxamide
ribonucleotide (AICAR)
NH
2
3
H
C
O
CH
2
O
H
C
HHC
OH
C
O
P
PRibNN
HN
C
H
N
N
H
N
1
-5H11032-Phosphoribosyl-ATP
H
2
O
7
H
C
O
CH
2
O
H
C
HHC
OH
C
O
P
PRibNN
HN
C
H
N
N
H
N
1
-5H11032-Phosphoribosyl-AMP
RibNN
H
H
2
O
C
O
C
O
H
2
N
PRibNN
CN
N
H
H
C
C
H
HH C
HO
CH
2
O
P
N
1
-5H11032-Phosphoribosylformimino-
5-aminoimidazole-4-
carboxamide ribonucleotide
H
2
N
RibNN
NC
O
P
H
2
N
To purine biosynthesis
C
N
OH
P
NH
3
HC
H COH
O
N
1
-5H11032-Phosphoribulosyl-
formimino-5-amino-
imidazole-4-carboxamide
ribonucleotide
Histidine
4
Glutamate H9251-Ketoglutarate
1
OH
PCH
2
HC
HCOH
O
HC
C
H
N
N
CH
Imidazole glycerol
3-phosphate
COO
H11002
O
P
CH
2
C
O
HC
C
N
Imidazole acetol
3-phosphate
CH
2
HC
CH
2
Glutamine
Glutamate
6
9
2NADH H11001 2H
H11001
NC
H
P
C
O
C
N
NH
3
L-Histidinol
phosphate
CH
2
HC
H
2
PP
i
8
P
i
2NAD
H11001
H
C
O
C
N
NH
3
L-Histidinol
CH
2
HC
H
H11001H11001 H11001
H
N
CH
H
N
CH
CH
2
H
N
CH
CH
2
H
N
CH
CH
2
ATP phosphoribosyl transferase
pyrophosphohydrolase
phosphoribosyl-AMP cyclohydrolase
phosphoribosylformimino-5-aminoimidazole-
4-carboxamide ribonucleotide isomerase
glutamine amidotransferase
imidazole glycerol 3-phosphate dehydratase
L-histidinol phosphate aminotransferase
histidinol phosphate phosphatase
histidinol dehydrogenase
1
2
3
4
5
6
7
8
9
FIGURE 22–20 Biosynthesis of histidine
in bacteria and plants. Atoms derived
from PRPP and ATP are shaded red and
blue, respectively. Two of the histidine
nitrogens are derived from glutamine and
glutamate (green). Note that the derivative
of ATP remaining after step 5 (AICAR) is
an intermediate in purine biosynthesis
(see Fig. 22–33, step 9 ), so ATP is
rapidly regenerated.
8885d_c22_833-880 2/6/04 8:35 AM Page 852 mac76 mac76:385_reb:
Because the 20 common amino acids must be made
in the correct proportions for protein synthesis, cells
have developed ways not only of controlling the rate of
synthesis of individual amino acids but also of coordi-
nating their formation. Such coordination is especially
well developed in fast-growing bacterial cells. Figure
22–22 shows how E. coli cells coordinate the synthesis
of lysine, methionine, threonine, and isoleucine, all
made from aspartate. Several important types of inhibi-
tion patterns are evident. The step from aspartate to
aspartyl-H9252-phosphate is catalyzed by three isozymes,
each independently controlled by different modulators.
This enzyme multiplicity prevents one biosynthetic
end product from shutting down key steps in a pathway
when other products of the same pathway are required.
The steps from aspartate H9252-semialdehyde to homoser-
ine and from threonine to H9251-ketobutyrate (detailed in
Fig. 22–15) are also catalyzed by dual, independently
controlled isozymes. One isozyme for the conversion of
aspartate to aspartyl-H9252-phosphate is allosterically inhib-
ited by two different modulators, lysine and isoleucine,
whose action is more than additive—another example
of concerted inhibition. The sequence from aspartate
to isoleucine undergoes multiple, overlapping negative
feedback inhibition; for example, isoleucine inhibits
the conversion of threonine to H9251-ketobutyrate (as de-
scribed above), and threonine inhibits its own forma-
tion at three points: from homoserine, from aspartate
H9252-semialdehyde, and from aspartate (steps 4 , 3 , and
1 in Fig. 22–15). This overall regulatory mechanism
is called sequential feedback inhibition.
Similar patterns are evident in the pathways lead-
ing to the aromatic amino acids. The first step of the
early pathway to the common intermediate chorismate
is catalyzed by the enzyme 2-keto-3-deoxy-D-arabino-
heptulosonate 7-phosphate (DAHP) synthase (step 1
in Fig. 22–16). Most microorganisms and plants have
three DAHP synthase isozymes. One is allosterically in-
hibited (feedback inhibition) by phenylalanine, another
by tyrosine, and the third by tryptophan. This scheme
helps the overall pathway to respond to cellular
22.2 Biosynthesis of Amino Acids 853
COO
H11002
H11001
NH
3
CH
2
H9251-Ketobutyrate
OO
CH
A
A
O
O Isoleucine
OCH
OH
CH
3
CH
3
CH
COO
H11002
O
threonine dehydratase
COO
H11002
CH
2
OO
O
OCH
3
B
C
A
CH
3
OCH
H11001
NH
3
A
Threonine
5 steps
FIGURE 22–21 Allosteric regulation of isoleucine biosynthesis. The
first reaction in the pathway from threonine to isoleucine is inhibited
by the end product, isoleucine. This was one of the first examples of
allosteric feedback inhibition to be discovered. The steps from H9251-
ketobutyrate to isoleucine correspond to steps 18 through 21 in Fig-
ure 22–15 (five steps because 19 is a two-step reaction).
FIGURE 22–22 Interlocking regulatory mechanisms in the biosyn-
thesis of several amino acids derived from aspartate in E. coli. Three
enzymes (A, B, C) have either two or three isozyme forms, indicated
by numerical subscripts. In each case, one isozyme (A
2
, B
1
, and C
2
)
has no allosteric regulation; these isozymes are regulated by changes
in the amount synthesized (Chapter 28). Synthesis of isozymes A
2
and
B
1
is repressed when methionine levels are high, and synthesis of
isozyme C
2
is repressed when isoleucine levels are high. Enzyme A
is aspartokinase; B, homoserine dehydrogenase; C, threonine dehy-
dratase.
Aspartate b-semialdehyde
Lysine
Methionine
B
1
B
2
A
2
A
3
A
1
Aspartate
Threonine
C
1
C
2
a-Ketobutyrate
Isoleucine
Aspartyl-b-phosphate
Homoserine
5 steps
3 steps
6 steps
8885d_c22_833-880 2/6/04 8:35 AM Page 853 mac76 mac76:385_reb:
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules854
requirements for one or more of the aromatic amino
acids. Additional regulation takes place after the path-
way branches at chorismate. For example, the enzymes
catalyzing the first two steps of the tryptophan branch
are subject to allosteric inhibition by tryptophan.
SUMMARY 22.2 Biosynthesis of Amino Acids
■ Plants and bacteria synthesize all 20 common
amino acids. Mammals can synthesize about
half; the others are required in the diet
(essential amino acids).
■ Among the nonessential amino acids, glutamate
is formed by reductive amination of
H9251-ketoglutarate and serves as the precursor of
glutamine, proline, and arginine. Alanine and
aspartate (and thus asparagine) are formed
from pyruvate and oxaloacetate, respectively,
by transamination. The carbon chain of serine
is derived from 3-phosphoglycerate. Serine is a
precursor of glycine; the H9252-carbon atom of
serine is transferred to tetrahydrofolate. In
microorganisms, cysteine is produced from
serine and from sulfide produced by the
reduction of environmental sulfate. Mammals
produce cysteine from methionine and serine
by a series of reactions requiring
S-adenosylmethionine and cystathionine.
■ Among the essential amino acids, the aromatic
amino acids (phenylalanine, tyrosine, and
tryptophan) form by a pathway in which
chorismate occupies a key branch point.
Phosphoribosyl pyrophosphate is a precursor of
tryptophan and histidine. The pathway to
histidine is interconnected with the purine
synthetic pathway. Tyrosine can also be formed
by hydroxylation of phenylalanine (and thus is
considered conditionally essential). The
pathways for the other essential amino acids
are complex.
■ The amino acid biosynthetic pathways are
subject to allosteric end-product inhibition; the
regulatory enzyme is usually the first in the
sequence. Regulation of the various synthetic
pathways is coordinated.
22.3 Molecules Derived from Amino Acids
In addition to their role as the building blocks of proteins,
amino acids are precursors of many specialized biomol-
ecules, including hormones, coenzymes, nucleotides,
alkaloids, cell wall polymers, porphyrins, antibiotics,
pigments, and neurotransmitters. We describe here the
pathways to a number of these amino acid derivatives.
Glycine Is a Precursor of Porphyrins
The biosynthesis of porphyrins, for which glycine is a
major precursor, is our first example, because of the
central importance of the porphyrin nucleus in heme
proteins such as hemoglobin and the cytochromes. The
porphyrins are constructed from four molecules of the
monopyrrole derivative porphobilinogen, which itself is
derived from two molecules of H9254-aminolevulinate. There
are two major pathways to H9254-aminolevulinate. In higher
eukaryotes (Fig. 22–23a), glycine reacts with succinyl-
CoA in the first step to yield H9251-amino-H9252-ketoadipate,
which is then decarboxylated to H9254-aminolevulinate. In
plants, algae, and most bacteria, H9254-aminolevulinate is
formed from glutamate (Fig. 22–23b). The glutamate is
first esterified to glutamyl-tRNA
Glu
(see Chapter 27 on
the topic of transfer RNAs); reduction by NADPH con-
verts the glutamate to glutamate 1-semialdehyde, which
is cleaved from the tRNA. An aminotransferase converts
the glutamate 1-semialdehyde to H9254-aminolevulinate.
In all organisms, two molecules of H9254-aminolevulinate
condense to form porphobilinogen and, through a series
of complex enzymatic reactions, four molecules of por-
phobilinogen come together to form protoporphyrin
(Fig. 22–24). The iron atom is incorporated after the
protoporphyrin has been assembled, in a step catalyzed
by ferrochelatase. Porphyrin biosynthesis is regulated
in higher eukaryotes by the concentration of the heme
product, which serves as a feedback inhibitor of early
steps in the synthetic pathway. Genetic defects in the
biosynthesis of porphyrins can lead to the accumulation
of pathway intermediates, causing a variety of human
diseases known collectively as porphyrias (Box 22–1).
Heme Is the Source of Bile Pigments
The iron-porphyrin (heme) group of hemoglobin,
released from dying erythrocytes in the spleen,
is degraded to yield free Fe
3H11001
and, ultimately, biliru-
bin. This pathway is arresting for its capacity to inject
color into human biochemistry.
The first step in the two-step pathway, catalyzed by
heme oxygenase (HO), converts heme to biliverdin, a
linear (open) tetrapyrrole derivative (Fig. 22–25). The
other products of the reaction are free Fe
2H11001
and CO.
The Fe
2H11001
is quickly bound by ferritin. Carbon monox-
ide is a poison that binds to hemoglobin (see Box 5–1),
and the production of CO by heme oxygenase ensures
that, even in the absence of environmental exposure,
about 1% of an individual’s heme is complexed with CO.
Biliverdin is converted to bilirubin in the second
step, catalyzed by biliverdin reductase. You can monitor
this reaction colorimetrically in a familiar in situ exper-
iment. When you are bruised, the black and/or purple
color results from hemoglobin released from damaged
erythrocytes. Over time, the color changes to the green
of biliverdin, and then to the yellow of bilirubin. Biliru-
8885d_c22_833-880 2/6/04 8:35 AM Page 854 mac76 mac76:385_reb:
H11001
H11001
Glycine
Glutamate
CH
2
CO
2
COO
H11002
Succinyl-CoA
NH
3
H11001
CH
2
COO
H11002
CH
2
CH
H11001
C
NH
3
O
H9251-Amino-H9252-
ketoadipate
COO
H11002
CH
2
S-CoA
COO
H11002
CH
2
CH
2
COO
H11002
CH
2
C
O
CH
2
CH
2
C
COO
H11002
O
CH
2
H9254-Aminolevulinate
H11001
NH
3
CoA-SH
H11001
NH
3
HC
COO
H11002
CH
2
COO
H11002
CH
2
H11001
NH
3
HC
CH
2
COO
H11002
CH
2
H11001
NH
3
HC
C
ATP AMP PP
i
tRNA
Glu
tRNA
Glu
NADPH NADP
H11001
tRNA
Glu
O C
H
O
Glutamyl-tRNA
Glu
Glutamate 1-semialdehyde
H9254-aminolevulinate
synthase
H9254-aminolevulinate
synthase
glutamyl-tRNA
synthetase
glutamyl-tRNA
reductase
glutamate-1-semialdehyde
aminomutase
(a)
(b)
22.3 Molecules Derived from Amino Acids 855
FIGURE 22–23 Biosynthesis of H9254-aminolevulinate. (a) In mammals
and other higher eukaryotes, H9254-aminolevulinate is synthesized from
glycine and succinyl-CoA. The atoms furnished by glycine are shown
in red. (b) In bacteria and plants, the precursor of H9254-aminolevulinate
is glutamate.
CH
3
CH
3
COO
H11002
H11002
OOC
COO
H11002
Fe
2H11001
Fe
2H11001
CH
3
NH
3
H
3
N
8 H
2
O
Pr Pr
CH
3
CH
3
CH
3
CH
3
CH
3
NN
N N
O
N
H
Heme
CH
3
CH
3
CH
3
Pr Pr
CH
3
N
N
Protoporphyrin
Pr
P
NH
NH HN
HN
Pr
Pr
Ac
Ac
Ac
Ac
HO
Preuroporphyrinogen
PrPr
NH
NH HN
HN
Pr
Pr
AcAc
Ac
Ac
Uroporphyrinogen III
PrPr
NH
NH HN
HN
Pr
Pr
Coproporphyrinogen III
CH
3
CH
3
CH
3
CH
3
PrPr
NH
NH HN
HN
HN
HN
Protoporphyrinogen
1
6
2
84
H9254-Aminolevulinate
4 NH
4
H11001
3
H
2
O
5
4
4 CO
2
2 CO
2
7
Porphobilinogen
Vinyl group
porphobilinogen synthase
uroporphyrinogen synthase
uroporphyrinogen III cosynthase
uroporphyrinogen decarboxylase
coproporphyrinogen oxidase
protoporphyrinogen oxidase
ferrochelatase
1
2
3
4
5
6
7
FIGURE 22–24 Biosynthesis of heme from H9254-aminolevulinate. Ac rep-
resents acetyl (OCH
2
COO
H11002
); Pr, propionyl (OCH
2
CH
2
COO
H11002
).
8885d_c22_833-880 2/6/04 8:35 AM Page 855 mac76 mac76:385_reb:
bin is largely insoluble, and it travels in the bloodstream
as a complex with serum albumin. In the liver, bilirubin
is transformed to the bile pigment bilirubin diglu-
curonide. This product is sufficiently water-soluble to be
secreted with other components of bile into the small
intestine, where microbial enzymes convert it to several
products, predominantly urobilinogen. Some urobilino-
gen is reabsorbed into the blood and transported to the
kidney, where it is converted to urobilin, the compound
that gives urine its yellow color (Fig. 22–25, left branch).
Urobilinogen remaining in the intestine is converted (in
another microbe-dependent reaction) to stercobilin (Fig.
22–25, right branch), which imparts the red-brown color
to feces.
Impaired liver function or blocked bile secretion
causes bilirubin to leak from the liver into the blood,
resulting in a yellowing of the skin and eyeballs, a con-
dition called jaundice. In cases of jaundice, determina-
tion of the concentration of bilirubin in the blood may
be useful in the diagnosis of underlying liver disease.
Newborn infants sometimes develop jaundice because
they have not yet produced enough glucuronyl bilirubin
transferase to process their bilirubin. A traditional
treatment to reduce excess bilirubin, exposure to a fluo-
rescent lamp, causes a photochemical conversion of
bilirubin to compounds that are more soluble and easily
excreted.
These pathways of heme breakdown play significant
roles in protecting cells from oxidative damage and in
regulating certain cellular functions. The CO produced
by heme oxygenase is toxic at high concentrations, but
at the very low concentrations generated during heme
degradation it appears to have some regulatory and/or
signaling functions. It acts as a vasodilator, much the
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules856
FIGURE 22–25 Bilirubin and its breakdown
products. M represents methyl; V, vinyl; Pr,
propionyl; E, ethyl. For ease of comparison,
these structures are shown in linear form,
rather than in their correct stereochemical
conformations.
transport to intestine
heme oxygenase CO
Fe
H110012
Heme
O
M V
N
H
N
H
M Pr
N
MPr
O
M V
N
H
O
M V
N
H
N
H
M Pr
N
H
MPr
H H
O
M V
N
H
biliverdin
reductase
NADPH, H
NADP
H11001
Biliverdin
H11001
Bilirubin
(in blood)
Bilirubin diglucuronide Bilirubin
(in bile)
transport to kidney
Urobilinogen Urobilinogen
(liver)
transport in blood
as complex with
serum albumin
glucuronyl-
bilirubin
transferase
O
M E
N
H
N
H
M Pr
N
MPr
O
M E
N
H
H
O
M E
N
H
N
H
M Pr
N
MPr
O
M E
N
H
H
H H HH
Urobilin Stercobilin
8885d_c22_833-880 2/6/04 8:35 AM Page 856 mac76 mac76:385_reb:
same as (but less potent than) nitric oxide (discussed
below). Low levels of CO also have some regulatory ef-
fects on neurotransmission. Bilirubin is the most abun-
dant antioxidant in mammalian tissues and is responsi-
ble for most of the antioxidant activity in serum. Its
protective effects appear to be especially important in
the developing brain of newborn infants. The cell toxi-
city associated with jaundice may be due to bilirubin
levels in excess of the serum albumin needed to solu-
bilize it.
Given these varied roles of heme degradation prod-
ucts, the degradative pathway is subject to regulation,
mainly at the first step. Humans have at least three
isozymes of heme oxygenase. HO-1 is highly regulated;
the expression of its gene is induced by a wide range of
stress conditions (shear stress, angiogenesis (uncon-
trolled development of blood vessels), hypoxia, hyper-
oxia, heat shock, exposure to ultraviolet light, hydrogen
peroxide, and many other metabolic insults). HO-2 is
found mainly in brain and testes, where it is continu-
ously expressed. The third isozyme, HO-3, is not yet well
characterized. ■
Amino Acids Are Precursors of Creatine
and Glutathione
Phosphocreatine, derived from creatine, is an im-
portant energy buffer in skeletal muscle (see Fig. 13–5).
Creatine is synthesized from glycine and arginine (Fig.
22–26); methionine, in the form of S-adenosylmethionine,
acts as methyl group donor.
Glutathione (GSH), present in plants, animals,
and some bacteria, often at high levels, can be thought
of as a redox buffer. It is derived from glycine, gluta-
mate, and cysteine (Fig. 22–27). The H9253-carboxyl group
of glutamate is activated by ATP to form an acyl phos-
phate intermediate, which is then attacked by the H9251-
amino group of cysteine. A second condensation reac-
tion follows, with the H9251-carboxyl group of cysteine
activated to an acyl phosphate to permit reaction with
glycine. The oxidized form of glutathione (GSSG), pro-
duced in the course of its redox activities, contains two
glutathione molecules linked by a disulfide bond.
Glutathione probably helps maintain the sulfhydryl
groups of proteins in the reduced state and the iron of
22.3 Molecules Derived from Amino Acids 857
BOX 22–1 BIOCHEMISTRY IN MEDICINE
Biochemistry of Kings and Vampires
Porphyrias (listed at right) are a group of genetic dis-
eases in which, because of defects in enzymes of the
biosynthetic pathway from glycine to porphyrins, spe-
cific porphyrin precursors accumulate in erythrocytes,
body fluids, and the liver. The most common form is
acute intermittent porphyria. Most affected individu-
als are heterozygotes and are usually asymptomatic,
because the single copy of the normal gene provides
a sufficient level of enzyme function. However, certain
nutritional or environmental factors (as yet poorly
understood) can cause a buildup of H9254-aminolevulinate
and porphobilinogen, leading to attacks of acute
abdominal pain and neurological dysfunction. King
George III, British monarch during the American
Revolution, suffered several episodes of apparent
madness that tarnished the record of this otherwise
accomplished man. The symptoms of his condition
suggest that George III suffered from acute intermit-
tent porphyria.
One of the rarer porphyrias results in an accu-
mulation of uroporphyrinogen I, an abnormal isomer
of a protoporphyrin precursor. This compound stains
the urine red, causes the teeth to fluoresce strongly
in ultraviolet light, and makes the skin abnormally sen-
sitive to sunlight. Many individuals with this porphyria
are anemic, because insufficient heme is synthesized.
This genetic condition may have given rise to the vam-
pire myths of folk legend.
The symptoms of most porphyrias are now read-
ily controlled with dietary changes or the administra-
tion of heme or heme derivatives.
H9254-Aminolevulinate
Porphobilinogen
Preuroporphyrinogen
Urophorphyrinogen III
Coproporphyrinogen
Protoporphyrinogen
Protoporphyrin
Heme
Doss porphyria
Acute intermittent porphyria
Congenital erythropoietic porphyria
Porphyria cutanea tarda
Hereditary coproporphyria
Variegate porphyria
Erythropoietic protoporphyria
8885d_c22_833-880 2/6/04 8:35 AM Page 857 mac76 mac76:385_reb:
bound selenium (Se) atom in the form of selenocysteine
(see Fig. 3–8a), which is essential for its activity.
D-Amino Acids Are Found Primarily in Bacteria
Although D-amino acids do not generally occur
in proteins, they do serve some special functions
in the structure of bacterial cell walls and peptide an-
tibiotics. Bacterial peptidoglycans (see Fig. 20–23) con-
tain both D-alanine and D-glutamate. D-Amino acids arise
directly from the L isomers by the action of amino acid
racemases, which have pyridoxal phosphate as cofactor
(see Fig. 18–6). Amino acid racemization is uniquely
important to bacterial metabolism, and enzymes such as
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules858
ATP
Glutamate
Cysteine
H9253-Glu–Cys–Gly
glutathione
synthetase
H9253-Glu Cys Gly
O
CCH N
SH
H11001
CH
2
(reduced)
Glutathione (GSH)
Glutathione (GSSG)
(oxidized)
Glycine
CH
CH
2
CH
2
C
O
N
HH
CH
2
NH
3
ATP
ADP H11001 P
i
H5008
OOC COO
H5008
ADP H11001 P
i
H9253-Glu–Cys–Gly
H9253-Glu–Cys
H9253-glutamyl
cysteine synthetase
(b)
(a)
S
S
FIGURE 22–27 Glutathione metabolism. (a) Biosynthesis of glu-
tathione. (b) Reduced form of glutathione.
heme in the ferrous (Fe
2H11001
) state, and it serves as a re-
ducing agent for glutaredoxin in deoxyribonucleotide
synthesis (see Fig. 22–39). Its redox function is also used
to remove toxic peroxides formed in the normal course
of growth and metabolism under aerobic conditions:
2GSH H11001 ROOOOOH 88n GSSG H11001 H
2
O H11001 ROOH
This reaction is catalyzed by glutathione peroxidase,
a remarkable enzyme in that it contains a covalently
A
OP
CH
2
H11001
N
COO
H11002
C
A
O
H11002
A
A
NH
3
H11001
NH
3
A
CH
3
A
CH
2
A
A
CH
H11001
NH
2
Arginine
Glycine
Ornithine
Guanidinoacetate
A
P
NH
C
A
NH
2
A
(CH
2
)
3
H11001
NH
2
COO
H11002
adoMet
adoHcy
Methionine
Creatine
Phosphocreatine
methyltransferase
creatine
kinase
amidinotransferase
A
P
NH
C
A
COO
H11002
NH
2
A
CH
2
A
H11001
NH
2
O
NH
2
ADP
A
A
A
A
A
A
O
P
P
O
COO
H11002
OP
NH
C
O
H11002
N
CH
2
COO
H11002
CH
3
NH
2
H11001
ATP
FIGURE 22–26 Biosynthesis of creatine and phosphocreatine. Crea-
tine is made from three amino acids: glycine, arginine, and methion-
ine. This pathway shows the versatility of amino acids as precursors
of other nitrogenous biomolecules.
8885d_c22_858 2/6/04 1:08 PM Page 858 mac76 mac76:385_reb:
alanine racemase are prime targets for pharmaceutical
agents. One such agent, L-fluoroalanine, is being
tested as an antibacterial drug. Another, cycloserine,
is used to treat tuberculosis. Because these inhibitors
also affect some PLP-requiring human enzymes, how-
ever, they have potentially undesirable side effects. ■
Aromatic Amino Acids Are Precursors
of Many Plant Substances
Phenylalanine, tyrosine, and tryptophan are converted
to a variety of important compounds in plants. The rigid
polymer lignin, derived from phenylalanine and tyro-
sine, is second only to cellulose in abundance in plant
tissues. The structure of the lignin polymer is complex
and not well understood. Tryptophan is also the pre-
cursor of the plant growth hormone indole-3-acetate, or
auxin (Fig. 22–28a), which has been implicated in the
regulation of a wide range of biological processes in
plants.
Phenylalanine and tyrosine also give rise to many
commercially significant natural products, including the
tannins that inhibit oxidation in wines; alkaloids such as
morphine, which have potent physiological effects; and
the flavoring of cinnamon oil (Fig. 22–28b), nutmeg,
cloves, vanilla, cayenne pepper, and other products.
Biological Amines Are Products
of Amino Acid Decarboxylation
Many important neurotransmitters are primary
or secondary amines, derived from amino acids
in simple pathways. In addition, some polyamines that
form complexes with DNA are derived from the amino
acid ornithine, a component of the urea cycle. A com-
mon denominator of many of these pathways is amino
acid decarboxylation, another PLP-requiring reaction
(see Fig. 18–6).
The synthesis of some neurotransmitters is illus-
trated in Figure 22–29. Tyrosine gives rise to a family of
catecholamines that includes dopamine, norepineph-
rine, and epinephrine. Levels of catecholamines are
correlated with, among other things, changes in blood
pressure. The neurological disorder Parkinson’s disease
is associated with an underproduction of dopamine, and
it has traditionally been treated by administering L-dopa.
Overproduction of dopamine in the brain may be linked
to psychological disorders such as schizophrenia.
Glutamate decarboxylation gives rise to H9253-amino-
butyrate (GABA), an inhibitory neurotransmitter. Its
underproduction is associated with epileptic seizures.
H
3
NH
HC
O
CH
2
NH
3
H11001
COO
C
OF
C
H11001
L-Fluoroalanine Cycloserine
H5008
H
2
NCH
GABA analogs are used in the treatment of epilepsy and
hypertension. Levels of GABA can also be increased by
administering inhibitors of the GABA-degrading enzyme
GABA aminotransferase. Another important neuro-
transmitter, serotonin, is derived from tryptophan in a
two-step pathway.
Histidine undergoes decarboxylation to histamine,
a powerful vasodilator in animal tissues. Histamine is re-
leased in large amounts as part of the allergic response,
and it also stimulates acid secretion in the stomach. A
growing array of pharmaceutical agents are being de-
signed to interfere with either the synthesis or the ac-
tion of histamine. A prominent example is the histamine
receptor antagonist cimetidine (Tagamet), a structural
analog of histamine:
It promotes the healing of duodenal ulcers by inhibiting
secretion of gastric acid.
CH
3
C
NHCH
2
N
CH
2
CH
2
NH
SNH
CH
3
N
C
N
22.3 Molecules Derived from Amino Acids 859
COO
H5008
CH
C
O
CH
2
NH
3
H11001
COO
H5008
CHCH
COO
H5008
amino-
transferase
decarboxylase
CH
2
N
H
COO
H5008
CH
2
CH
NH
3
H11001
COO
H5008
CH
2
N
H
N
H
Tryptophan
Indole-
3-pyruvate
Cinnamate
Indole-3-acetate
(auxin)
Phenylalanine
CO
2
phenylalanine
ammonia
lyase
NH
3
(a) (b)
FIGURE 22–28 Biosynthesis of two plant substances from amino acids.
(a) Indole-3-acetate (auxin) and (b) cinnamate (cinnamon flavor).
8885d_c22_859 2/6/04 2:00 PM Page 859 mac76 mac76:385_reb:
Polyamines such as spermine and spermidine, in-
volved in DNA packaging, are derived from methionine
and ornithine by the pathway shown in Figure 22–30. The
first step is decarboxylation of ornithine, a precursor of
arginine (Fig. 22–10). Ornithine decarboxylase, a
PLP-requiring enzyme, is the target of several powerful
inhibitors used as pharmaceutical agents (Box 22–2). ■
Arginine Is the Precursor for Biological Synthesis
of Nitric Oxide
A surprise finding in the mid-1980s was the role of ni-
tric oxide (NO)—previously known mainly as a compo-
nent of smog—as an important biological messenger.
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules860
COO
H5008
CH
CH
2
H
2
N
H11001
HO
Tyrosine
H
2
O
CH
Ascorbate
Dehydroascorbate
dopamine
CO
2
O
2
CH
CH
2
NH
3
Dopa
HO
PLP
CH
2
phenylethanolamine
N-methyltransferase
CH
2
NH
3
H11001
HO
CH
3
Dopamine
Tetrahydrobiopterin
Dihydrobiopterin
tryptophan
H
2
O
HO
HO
OH
NH
3
H11001
Serotonin
aromatic
amino acid
decarboxylase
adoMet
adoHcy
NH
3
H11001
HO
CHHO CH
2
OH
CH
2
Epinephrine
CO
2
PLP
aromatic
amino acid
decarboxylase
CHCH
2
H11001
H5008
OOC
5-Hydroxy-
tryptophan
CH
2
NH
3
H11001
CH
2
H5008
OOC
CH
2
H9253-Aminobutyrate
(GABA)
HO COO
H5008
CHCH
2
NH
3
H11001
COO
H5008
Glutamate
NH
3
COO
H5008
CO
2
NH
3
H11001
PLP
glutamate
CH
2
CH
2
Norepinephrine
CHCH
2
H11001
Histidine
NH
3
COO
H5008
N NH
H
2
O
Tetrahydrobiopterin
Dihydrobiopterin
tyrosine
O
2
CH
2
H11001
NH
3
N
H
N
H
CO
2
H11001
PLP
histidine
HO
CHCH
2
H11001
NH
3
COO
H5008
Tryptophan
N
H
CH
2
decarboxylase
H9252-hydroxylase
decarboxylase
HO
hydroxylase
hydroxylase
CH
2
N NH
Histamine
O
2
HO
FIGURE 22–29 Biosynthesis of some neurotransmitters from amino
acids. The key step is the same in each case: a PLP-dependent de-
carboxylation (shaded in pink).
This simple gaseous substance diffuses readily through
membranes, although its high reactivity limits its range
of diffusion to about a 1 mm radius from the site of syn-
thesis. In humans NO plays a role in a range of physio-
logical processes, including neurotransmission, blood
clotting, and the control of blood pressure. Its mode of
action is described in Chapter 12 (p. 434).
Nitric oxide is synthesized from arginine in an
NADPH-dependent reaction catalyzed by nitric oxide
synthase (Fig. 22–31), a dimeric enzyme structurally re-
lated to NADPH cytochrome P-450 reductase (see Box
21–1). The reaction is a five-electron oxidation. Each
subunit of the enzyme contains one bound molecule of
each of four different cofactors: FMN, FAD, tetrahydro-
biopterin, and Fe
3H11001
heme. NO is an unstable molecule
and cannot be stored. Its synthesis is stimulated by in-
teraction of nitric oxide synthase with Ca
2H11001
-calmodulin
(see Fig. 12–21).
8885d_c22_833-880 2/6/04 8:35 AM Page 860 mac76 mac76:385_reb:
SUMMARY 22.3 Molecules Derived
from Amino Acids
■ Many important biomolecules are derived from
amino acids. Glycine is a precursor of por-
phyrins. Degradation of iron-porphyrin (heme)
generates bilirubin, which is converted to bile
pigments, with several physiological functions.
■ Glycine and arginine give rise to creatine and
phosphocreatine, an energy buffer. Glutathione,
formed from three amino acids, is an important
cellular reducing agent.
■ Bacteria synthesize D-amino acids from L-amino
acids in racemization reactions requiring
pyridoxal phosphate.
■ The aromatic amino acids give rise to many
plant substances. The PLP-dependent
decarboxylation of some amino acids yields
important biological amines, including
neurotransmitters.
■ Arginine is the precursor of nitric oxide, a
biological messenger.
Arginine
COO
H11002
CHH
3
N
CH
2
CH
2
CH
2
NH
NH
2
CNH
2
H11001
H11001
Hydroxyarginine
NADPH, O
2
NADP
H11001
, H
2
O
NADPH, O
2
NADP
H11001
, H
2
O
COO
H11002
CHH
3
N
CH
2
CH
2
CH
2
NH
NH
2
CNOH
H11001
Citrulline
COO
H11002
CHH
3
N
CH
2
CH
2
CH
2
NH
NH
2
CO
H11001
H11001 NO
?
Nitric
oxide
2
1
2
1
FIGURE 22–31 Biosynthesis of nitric oxide. Both steps are catalyzed by nitric oxide synthase.
The nitrogen of the NO is derived from the guanidino group of arginine.
H
3
N
CO
2
CH
H11001
S
Methylthioadenosine
Adenosine
CH
2
CH
2
H11001
ornithine
PLP
CH
3
CH
2
H
3
N
CH
2
CH
3
Adenosine
H11001
H
3
N
CH
2
(CH
2
)
4
H11001
S
COO
Ornithine
Putrescine
Decarboxylated
adoMet
CO
2
adoMet
PLP
COO
H11002
S
H11001
CH
2
CH
2
CHCH
2
NH
propylaminotransferase II
Spermidine
Adenosine
H
3
N
PP
i
H11001 P
i
S-Adenosylmethionine
H
3
N
H11001 H11001
NH
3
NH (CH
2
)
3
Spermine
AdenosineS
propylaminotransferase I
NH
3
H
3
N
H11001
(CH
2
)
4
H11001
NH
3
NH(CH
2
)
3
(CH
2
)
3
CH
3
H11001
NH
3
decarboxylase
decarboxylase
(CH
2
)
4
Methionine
ATP
CH
3
H11001H11001
FIGURE 22–30 Biosynthesis of spermidine and spermine. The PLP-
dependent decarboxylation steps are shaded in pink. In these reac-
tions, S-adenosylmethionine (in its decarboxylated form) acts as a
source of propylamino groups (shaded blue).
861
8885d_c22_833-880 2/6/04 8:35 AM Page 861 mac76 mac76:385_reb:
22.4 Biosynthesis and Degradation
of Nucleotides
As discussed in Chapter 8, nucleotides play a variety
of important roles in all cells. They are the precursors
of DNA and RNA. They are essential carriers of chem-
ical energy—a role primarily of ATP and to some
extent GTP. They are components of the cofactors
NAD, FAD, S-adenosylmethionine, and coenzyme A, as
well as of activated biosynthetic intermediates such
as UDP-glucose and CDP-diacylglycerol. Some, such as
cAMP and cGMP, are also cellular second messengers.
Two types of pathways lead to nucleotides: the de
novo pathways and the salvage pathways. De novo
synthesis of nucleotides begins with their metabolic pre-
cursors: amino acids, ribose 5-phosphate, CO
2
, and NH
3
.
Salvage pathways recycle the free bases and nucleosides
released from nucleic acid breakdown. Both types of
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules862
BOX 22–2 BIOCHEMISTRY IN MEDICINE
Curing African Sleeping Sickness with a
Biochemical Trojan Horse
African sleeping sickness, or African trypanosomiasis,
is caused by protists (single-celled eukaryotes) called
trypanosomes (Fig. 1). This disease (and related
trypanosome-caused diseases) is medically and eco-
nomically significant in many developing nations.
Until recently, the disease was virtually incurable. Vac-
cines are ineffective, because the parasite has a novel
mechanism to evade the host immune system.
The cell coat of trypanosomes is covered with a sin-
gle protein, which is the antigen to which the immune
system responds. Every so often, however, by a process
of genetic recombination (see Table 28–1), a few cells
in the population of infecting trypanosomes switch
to a new protein coat, not recognized by the immune
system. This process of “changing coats” can occur
hundreds of times. The result is a chronic cyclic infec-
tion: the human host develops a fever, which subsides
as the immune system beats back the first infection;
trypanosomes with changed coats then become the
seed for a second infection, and the fever recurs. This
cycle can repeat for weeks, and the weakened person
eventually dies.
Some modern approaches to treating African
sleeping sickness have been based on an under-
standing of enzymology and metabolism. In at least
one such approach, this involves pharmaceutical
agents designed as mechanism-based enzyme inacti-
vators (suicide inactivators; p. 211). A vulnerable
point in trypanosome metabolism is the pathway of
polyamine biosynthesis. The polyamines spermine
and spermidine, used in DNA packaging, are required
in large amounts in rapidly dividing cells. The first
step in their synthesis is catalyzed by ornithine
decarboxylase, a PLP-requiring enzyme (see Fig.
FIGURE 1 Trypanosoma brucei rhodesiense, one of several
trypanosomes known to cause African sleeping sickness.
O
H5008
CH
OH
H
3
N
H
H11001
CH
3
P O
H
2
O
CH
C
O
CH
2
N
H11001
NH
Ornithine
Putrescine
O
H5008
CH
OH
H
2
N
H
H11001
CH
3
P O
(CH
2
)
3
CH
C
O
O
CH
2
N
NH
2
Pyridoxal
phosphate
H11001
CH
OH
H
2
N
H
CH
3
P O
CH
CH
2
N
NH
Schiff base
(CH
2
)
3
(CH
2
)
3
CO
2
H
2
O
H11001
H
H
H11001
H11001
FIGURE 2 Mechanism of ornithine decarboxylase reaction.
8885d_c22_862 2/6/04 1:09 PM Page 862 mac76 mac76:385_reb:
pathways are important in cellular metabolism and both
are presented in this section.
The de novo pathways for purine and pyrimidine
biosynthesis appear to be nearly identical in all living
organisms. Notably, the free bases guanine, adenine,
thymine, cytidine, and uracil are not intermediates in
these pathways; that is, the bases are not synthesized
and then attached to ribose, as might be expected. The
purine ring structure is built up one or a few atoms at
a time, attached to ribose throughout the process. The
pyrimidine ring is synthesized as orotate, attached to
ribose phosphate, and then converted to the common
pyrimidine nucleotides required in nucleic acid synthe-
sis. Although the free bases are not intermediates in the
de novo pathways, they are intermediates in some of the
salvage pathways.
Several important precursors are shared by the de
novo pathways for synthesis of pyrimidines and purines.
22.4 Biosynthesis and Degradation of Nucleotides 863
22–30). In mammalian cells, ornithine decarboxylase
undergoes rapid turnover—that is, a constant round
of enzyme degradation and synthesis. In some try-
panosomes, however, the enzyme—for reasons not
well understood—is stable, not readily replaced by
newly synthesized enzyme. An inhibitor of ornithine
decarboxylase that binds permanently to the enzyme
would thus have little effect on human cells, which
could rapidly replace inactivated enzyme, but would
adversely affect the parasite.
The first few steps of the normal reaction cat-
alyzed by ornithine decarboxylase are shown in Fig-
H11001
H
2
O
DFMO
F
additional
rearrangements
H5008
OH
H
2
N
H
H11001
CH
3
O
CH
C
O
O
CH
2
N
NH
2
Pyridoxal
phosphate
H11001
OH
H
2
N
H
CH
3
P O
CH
CH
2
N
NH
Stuck!
P
C
CH
F
C
F
O
H5008
C
O
CH
F
C
H11001
OH
H
2
CH
3
P O
CH
CH
2
N
H11001
NH
F
C
EnzymeF
H11001
Cys-S
C
H11001
OH
H
2
N
H
CH
3
P O
CH
CH
2
N
NH
C
Enzyme
Schiff base
O
H5008
(CH
2
)
3
(CH
2
)
3
(CH
2
)
3
N
H
HH
(CH
2
)
3
Cys-S
CO
2
H11001 F
H5008
ure 2. Once CO
2
is released, the electron movement
is reversed and putrescine is produced (see Fig.
22–30). Based on this mechanism, several suicide in-
activators have been designed, one of which is difluor-
omethylornithine (DFMO). DFMO is relatively inert in
solution. When it binds to ornithine decarboxylase,
however, the enzyme is quickly inactivated (Fig. 3).
The inhibitor acts by providing an alternative electron
sink in the form of two strategically placed fluorine
atoms, which are excellent leaving groups. Instead of
electrons moving into the ring structure of PLP, the
reaction results in displacement of a fluorine atom.
The S of a Cys residue at the enzyme’s active site then
forms a covalent complex with the highly reactive
PLP-inhibitor adduct in an essentially irreversible re-
action. In this way, the inhibitor makes use of the en-
zyme’s own reaction mechanisms to kill it.
DFMO has proved highly effective against African
sleeping sickness in clinical trials and is now used to
treat African sleeping sickness caused by T. brucei
gambiense. Approaches such as this show great
promise for treating a wide range of diseases. The
design of drugs based on enzyme mechanism and
structure can complement the more traditional trial-
and-error methods of developing pharmaceuticals.
FIGURE 3 Inhibition of ornithine
decarboxylase by DFMO.
8885d_c22_863 2/6/04 1:09 PM Page 863 mac76 mac76:385_reb:
Phosphoribosyl pyrophosphate (PRPP) is important in
both, and in these pathways the structure of ribose is
retained in the product nucleotide, in contrast to its
fate in the tryptophan and histidine biosynthetic path-
ways discussed earlier. An amino acid is an important
precursor in each type of pathway: glycine for purines
and aspartate for pyrimidines. Glutamine again is the
most important source of amino groups—in five differ-
ent steps in the de novo pathways. Aspartate is also
used as the source of an amino group in the purine
pathways, in two steps.
Two other features deserve mention. First, there is
evidence, especially in the de novo purine pathway, that
the enzymes are present as large, multienzyme com-
plexes in the cell, a recurring theme in our discussion
of metabolism. Second, the cellular pools of nucleotides
(other than ATP) are quite small, perhaps 1% or less of
the amounts required to synthesize the cell’s DNA.
Therefore, cells must continue to synthesize nucleotides
during nucleic acid synthesis, and in some cases nu-
cleotide synthesis may limit the rates of DNA replica-
tion and transcription. Because of the importance of
these processes in dividing cells, agents that inhibit nu-
cleotide synthesis have become particularly important
to modern medicine.
We examine here the biosynthetic pathways of
purine and pyrimidine nucleotides and their regulation,
the formation of the deoxynucleotides, and the degra-
dation of purines and pyrimidines to uric acid and urea.
We end with a discussion of chemotherapeutic agents
that affect nucleotide synthesis.
De Novo Purine Nucleotide
Synthesis Begins with PRPP
The two parent purine nu-
cleotides of nucleic acids are
adenosine 5H11032-monophosphate
(AMP; adenylate) and guano-
sine 5H11032-monophosphate (GMP;
guanylate), containing the
purine bases adenine and gua-
nine. Figure 22–32 shows the
origin of the carbon and nitro-
gen atoms of the purine ring
system, as determined by John Buchanan using isotopic
tracer experiments in birds. The detailed pathway of
purine biosynthesis was worked out primarily by
Buchanan and G. Robert Greenberg in the 1950s.
In the first committed step of the pathway, an amino
group donated by glutamine is attached at C-1 of PRPP
(Fig. 22–33). The resulting 5-phosphoribosylamine is
highly unstable, with a half-life of 30 seconds at pH 7.5.
The purine ring is subsequently built up on this struc-
ture. The pathway described here is identical in all or-
ganisms, with the exception of one step that differs in
higher eukaryotes as noted below.
The second step is the addition of three atoms from
glycine (Fig. 22–33, step 2 ). An ATP is consumed to
activate the glycine carboxyl group (in the form of an
acyl phosphate) for this condensation reaction. The
added glycine amino group is then formylated by N
10
-
formyltetrahydrofolate (step 3 ), and a nitrogen is con-
tributed by glutamine (step 4 ), before dehydration and
ring closure yield the five-membered imidazole ring of
the purine nucleus, as 5-aminoimidazole ribonucleotide
(AIR; step 5 ).
At this point, three of the six atoms needed for the
second ring in the purine structure are in place. To com-
plete the process, a carboxyl group is first added (step
6 ). This carboxylation is unusual in that it does not re-
quire biotin, but instead uses the bicarbonate generally
present in aqueous solutions. A rearrangement transfers
the carboxylate from the exocyclic amino group to po-
sition 4 of the imidazole ring (step 7 ). Steps 6 and 7
are found only in bacteria and fungi. In higher eukary-
otes, including humans, the 5-aminoimidazole ribonu-
cleotide product of step 5 is carboxylated directly to
carboxyaminoimidazole ribonucleotide in one step in-
stead of two (step 6a). The enzyme catalyzing this re-
action is AIR carboxylase.
Aspartate now donates its amino group in two steps
( 8 and 9 ): formation of an amide bond, followed by
elimination of the carbon skeleton of aspartate (as fu-
marate). Recall that aspartate plays an analogous role
in two steps of the urea cycle (see Fig. 18–10). The fi-
nal carbon is contributed by N
10
-formyltetrahydrofolate
(step 10 ), and a second ring closure takes place to yield
the second fused ring of the purine nucleus (step 11).
864
Amide N
of glutamine
CO
2
C
C
C
N
Formate
Aspartate
Formate
Glycine
C
N
N
C
N
FIGURE 22–32 Origin of the ring atoms of purines. This information
was obtained from isotopic experiments with
14
C- or
15
N-labeled pre-
cursors. Formate is supplied in the form of N
10
-formyltetrahydrofolate.
FIGURE 22–33 (facing page) De novo synthesis of purine nucleotides:
construction of the purine ring of inosinate (IMP). Each addition to
the purine ring is shaded to match Figure 22–32. After step 2 , R sym-
bolizes the 5-phospho-D-ribosyl group on which the purine ring is
built. Formation of 5-phosphoribosylamine (step 1 ) is the first com-
mitted step in purine synthesis. Note that the product of step 9 ,
AICAR, is the remnant of ATP released during histidine biosynthesis
(see Fig. 22–20, step 5 ). Abbreviations are given for most interme-
diates to simplify the naming of the pathway enzymes. Step 6a is the
alternative path from AIR to CAIR occurring in higher eukaryotes.
John Buchanan
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules
8885d_c22_833-880 2/6/04 8:35 AM Page 864 mac76 mac76:385_reb:
H
2
H
OP
H
OH
N
O
Aspartate
CH
2
O
H
OH
OP
P
H
H
OH
H
P
O
CH
2
H
H
4
ATP
OH
NH
H
2
C
H11001
O
5
ADP H11001 P
i
H
R
Glutamate
OC
N
H
H
N
Glycinamide
ribonucleotide (GAR)
N-Formylaminoimidazole-
4-carboxamide ribonucleotide (FAICAR)
11R
Glutamine
ATP
H
OC
H
H
O
R
Formylglycinamide
ribonucleotide (FGAR)
PP
i
1
Glutamine
Glutamate
ADP H11001 P
i
ATP
H
2
C
H
2
O
H
CH
C
CH
R
N
Formylglycinamidine
ribonucleotide (FGAM)
8
ADP H11001 P
i
ATP
6
5-Phospho-H9252-
D-ribosylamine
HCO
3
H5008
H
C
C
C
COO
H5008
H5008
N
H
2
C
Glycine
ATP
2
ADP H11001 P
i
5-Aminoimidazole
ribonucleotide (AIR)
H
C
O
N
H
2
N
C
N
N
H
H
H
2
O
C
R
COO
H5008
5-Aminoimidazole-4-carboxamide
ribonucleotide (AICAR)
10
7
N
10
-Formyl H
4
folate
H
4
folate
H
N
C
C
C
R
N
H
2
N
O
H
H
2
N
O
5-Phosphoribosyl
1-pyrophosphate (PRPP)
3
N
10
-Formyl H
4
folate
H
4
folate
HC
C
O
N
N
C
R
O
C
H
2
N
HC
C
H
R
O
N
Carboxyamino-
imidazole ribonucleotide
(CAIR)
N
5
-Carboxyaminoimidazole
ribonucleotide
(N
5
-CAIR)
carboxamide
-Succinyl-5-aminoimidazole-4-
ribonucleotide (SAICAR)
9 Fumarate
H
2
N
C
N
C
HC
N
C
N
N
H
NHC
C
N
O
OH H
H
H
Inosinate (IMP)
O
C
O
H5008
PO
CH
2
CH
2
O
H
NH
3
C
C
N
N
N
C
H
2
N
N
CH
C
CH
R
N
ADP H11001 P
i
H5008
OO
H5008
O
C
N
AIR
H
ONC
CO
2
6a
glutamine-PRPP
amidotransferase
GAR synthetase
GAR transformylase
FGAR amidotransferase
FGAM cyclase
(AIR synthetase)
N
5
-CAIR synthetase
AIR carboxylase
N
5
-CAIR mutase
SAICAR synthetase
SAICAR lyase
AICAR transformylase
IMP synthase
1
2
3
4
5
6
6a
7
8
9
10
11
865
8885d_c22_833-880 2/6/04 8:35 AM Page 865 mac76 mac76:385_reb:
The first intermediate with a complete purine ring is
inosinate (IMP).
As in the tryptophan and histidine biosynthetic
pathways, the enzymes of IMP synthesis appear to be
organized as large, multienzyme complexes in the cell.
Once again, evidence comes from the existence of sin-
gle polypeptides with several functions, some catalyz-
ing nonsequential steps in the pathway. In eukaryotic
cells ranging from yeast to fruit flies to chickens, steps
1 , 3 , and 5 in Figure 22–33 are catalyzed by a mul-
tifunctional protein. An additional multifunctional pro-
tein catalyzes steps 10 and 11. In humans, a multi-
functional enzyme combines the activities of AIR
carboxylase and SAICAR synthetase (steps 6a and 8 ).
In bacteria, these activities are found on separate pro-
teins, but a large noncovalent complex may exist in
these cells. The channeling of reaction intermediates
from one enzyme to the next permitted by these com-
plexes is probably especially important for unstable in-
termediates such as 5-phosphoribosylamine.
Conversion of inosinate to adenylate requires the
insertion of an amino group derived from aspartate (Fig.
22–34); this takes place in two reactions similar to those
used to introduce N-1 of the purine ring (Fig. 22–33,
steps 8 and 9 ). A crucial difference is that GTP rather
than ATP is the source of the high-energy phosphate in
synthesizing adenylosuccinate. Guanylate is formed by
the NAD
H11001
-requiring oxidation of inosinate at C-2, fol-
lowed by addition of an amino group derived from glu-
tamine. ATP is cleaved to AMP and PP
i
in the final step
(Fig. 22–34).
Purine Nucleotide Biosynthesis Is Regulated
by Feedback Inhibition
Three major feedback mechanisms cooperate in regu-
lating the overall rate of de novo purine nucleotide syn-
thesis and the relative rates of formation of the two end
products, adenylate and guanylate (Fig. 22–35). The
first mechanism is exerted on the first reaction that is
unique to purine synthesis—transfer of an amino group
to PRPP to form 5-phosphoribosylamine. This reaction
is catalyzed by the allosteric enzyme glutamine-PRPP
amidotransferase, which is inhibited by the end prod-
ucts IMP, AMP, and GMP. AMP and GMP act synergisti-
cally in this concerted inhibition. Thus, whenever either
AMP or GMP accumulates to excess, the first step in its
biosynthesis from PRPP is partially inhibited.
In the second control mechanism, exerted at a later
stage, an excess of GMP in the cell inhibits formation of
xanthylate from inosinate by IMP dehydrogenase, with-
out affecting the formation of AMP (Fig. 22–35). Con-
versely, an accumulation of adenylate inhibits formation
of adenylosuccinate by adenylosuccinate synthetase,
without affecting the biosynthesis of GMP. In the third
mechanism, GTP is required in the conversion of IMP
to AMP (Fig. 22–34, step 1 ), whereas ATP is required
for conversion of IMP to GMP (step 4 ), a reciprocal
arrangement that tends to balance the synthesis of the
two ribonucleotides.
The final control mechanism is the inhibition of
PRPP synthesis by the allosteric regulation of ribose
phosphate pyrophosphokinase. This enzyme is inhibited
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules866
P
Inosinate
(IMP)
NH
2
Rib
H
2
O
H
C
H5008
OOC COO
H5008
P
Guanylate
(GMP)
N
Rib
P
Adenylosuccinate
NH
Rib
P
NAD
H11001
O
Rib
P
Xanthylate
(XMP)
O
Rib
O
GTP
Aspartate
GDP H11001 P
i
CH
2
NADH H11001 H
H11001
Adenylate
(AMP)
Fumarate
ATPGlu AMP H11001 PP
i
H
2
N
Gln
H
2
O
N
HN
N
N
N
HN
N
N
N
N
N
N
N
N
N
N
H
O
N
HN
N
adenylosuccinate
synthetase
adenylosuccinate
lyase
IMP
dehydrogenase
XMP-glutamine
amidotransferase
FIGURE 22–34 Biosynthesis of AMP
and GMP from IMP.
8885d_c22_833-880 2/6/04 8:35 AM Page 866 mac76 mac76:385_reb:
cytidylate
synthetase
ATP
ADP H11001 P
i
Uridine 5H11032-triphosphate (UTP)
kinases
2 ATP
2ADP
Uridylate (UMP)
orotidylate
decarboxylase
CO
2
Orotidylate
orotate
phosphoribosyl-
transferase
PRPP
PP
i
Orotate
dihydroorotate
dehydrogenase
NAD
H11001
NADH H11001 H
H11001
L-Dihydroorotate
dihydroorotase
H
2
O
N-Carbamoylaspartate
aspartate
trans-
carbamoylase
Carbamoyl
phosphate
P
i
Aspartate
Gln
Glu
Cytidine 5H11032-triphosphate (CTP)
OCH
2
O
H
OH
H
HH
OH
O
O
PPP
C
N
CH
N
C
CH
NH
2
OCH
2
O
H
OH
H
HH
OH
O
O
P
P
C
N
O
C
HN
C
CH
O
OCH
2
O
H
OH
H
HH
OH
O
O
C
N
CH
HN
C
CH
O
C
N
H
C
HN
C
CH
COO
H5008
COO
H5008
O
O
C
N
H
CH
HN
C
CH
2
COO
H5008
O
O
O
C
N
H
CH
H
2
N
C
CH
2
COO
H5008
H5008
O
22.4 Biosynthesis and Degradation of Nucleotides 867
Ribose 5-phosphate
ribose phosphate
pyrophosphokinase
(PRPP synthetase)
glutamine-PRPP
amidotransferase
IMP
GMP
AMP
XMP-glutamine
amidotransferase
IMP
dehydrogenase
adenylosuccinate
synthetase
adenylosuccinate
lyase
AMP
GMP
IMP
ADP
AMP GMP
PRPP
5-Phosphoribosylamine
Adenylosuccinate
XMP
9 steps
ATPADP
FIGURE 22–35 Regulatory mechanisms in the biosynthesis of ade-
nine and guanine nucleotides in E. coli. Regulation of these pathways
differs in other organisms.
FIGURE 22–36 De novo synthesis of pyrimidine nucleotides: biosyn-
thesis of UTP and CTP via orotidylate. The pyrimidine is constructed
from carbamoyl phosphate and aspartate. The ribose 5-phosphate is
then added to the completed pyrimidine ring by orotate phosphori-
bosyltransferase. The first step in this pathway (not shown here; see
Fig. 18–11a) is the synthesis of carbamoyl phosphate from CO
2
and
NH
4
H11001
, catalyzed in eukaryotes by carbamoyl phosphate synthetase II.
by ADP and GDP, in addition to metabolites from other
pathways of which PRPP is a starting point.
Pyrimidine Nucleotides Are Made from Aspartate,
PRPP, and Carbamoyl Phosphate
The common pyrimidine ribonucleotides are cytidine 5H11032-
monophosphate (CMP; cytidylate) and uridine 5H11032-
monophosphate (UMP; uridylate), which contain the
pyrimidines cytosine and uracil. De novo pyrimidine nu-
cleotide biosynthesis (Fig. 22–36) proceeds in a some-
what different manner from purine nucleotide synthe-
sis; the six-membered pyrimidine ring is made first and
then attached to ribose 5-phosphate. Required in this
process is carbamoyl phosphate, also an intermediate in
the urea cycle (see Fig. 18–10). However, as we noted
8885d_c22_833-880 2/6/04 8:35 AM Page 867 mac76 mac76:385_reb:
in Chapter 18, in animals the carbamoyl phosphate re-
quired in urea synthesis is made in mitochondria by car-
bamoyl phosphate synthetase I, whereas the carbamoyl
phosphate required in pyrimidine biosynthesis is made
in the cytosol by a different form of the enzyme, car-
bamoyl phosphate synthetase II. In bacteria, a sin-
gle enzyme supplies carbamoyl phosphate for the syn-
thesis of arginine and pyrimidines. The bacterial enzyme
has three separate active sites, spaced along a channel
nearly 100 ? long (Fig. 22–37). Bacterial carbamoyl
phosphate synthetase provides a vivid illustration of the
channeling of unstable reaction intermediates between
active sites.
Carbamoyl phosphate reacts with aspartate to yield
N-carbamoylaspartate in the first committed step of
pyrimidine biosynthesis (Fig. 22–36). This reaction is
catalyzed by aspartate transcarbamoylase. In bacte-
ria, this step is highly regulated, and bacterial aspartate
transcarbamoylase is one of the most thoroughly stud-
ied allosteric enzymes (see below). By removal of wa-
ter from N-carbamoylaspartate, a reaction catalyzed by
dihydroorotase, the pyrimidine ring is closed to form
L-dihydroorotate. This compound is oxidized to the
pyrimidine derivative orotate, a reaction in which NAD
H11001
is the ultimate electron acceptor. In eukaryotes, the first
three enzymes in this pathway—carbamoyl phosphate
synthetase II, aspartate transcarbamoylase, and dihy-
droorotase—are part of a single trifunctional protein.
The protein, known by the acronym CAD, contains three
identical polypeptide chains (each of M
r
230,000), each
with active sites for all three reactions. This suggests
that large, multienzyme complexes may be the rule in
this pathway.
Once orotate is formed, the ribose 5-phosphate side
chain, provided once again by PRPP, is attached to yield
orotidylate (Fig. 22–36). Orotidylate is then decarboxy-
lated to uridylate, which is phosphorylated to UTP. CTP
is formed from UTP by the action of cytidylate syn-
thetase, by way of an acyl phosphate intermediate
(consuming one ATP). The nitrogen donor is normally
glutamine, although the cytidylate synthetases in many
species can use NH
4
H11001
directly.
Pyrimidine Nucleotide Biosynthesis Is Regulated
by Feedback Inhibition
Regulation of the rate of pyrimidine nucleotide synthe-
sis in bacteria occurs in large part through aspartate
transcarbamoylase (ATCase), which catalyzes the first
reaction in the sequence and is inhibited by CTP, the
end product of the sequence (Fig. 22–36). The bacter-
ial ATCase molecule consists of six catalytic subunits
and six regulatory subunits (see Fig. 6–27). The cat-
alytic subunits bind the substrate molecules, and the al-
losteric subunits bind the allosteric inhibitor, CTP. The
entire ATCase molecule, as well as its subunits, exists
in two conformations, active and inactive. When CTP is
not bound to the regulatory subunits, the enzyme is
maximally active. As CTP accumulates and binds to the
regulatory subunits, they undergo a change in confor-
mation. This change is transmitted to the catalytic sub-
units, which then also shift to an inactive conformation.
ATP prevents the changes induced by CTP. Figure 22–38
shows the effects of the allosteric regulators on the ac-
tivity of ATCase.
Nucleoside Monophosphates Are Converted
to Nucleoside Triphosphates
Nucleotides to be used in biosynthesis are generally con-
verted to nucleoside triphosphates. The conversion
pathways are common to all cells. Phosphorylation of
AMP to ADP is promoted by adenylate kinase, in the
reaction
868
FIGURE 22–37 Channeling of intermediates in bacterial carbamoyl
phosphate synthetase. (Derived from PDB ID 1M6V.) The reaction cat-
alyzed by this enzyme is illustrated in Figure 18–11a. The large and
small subunits are shown in gray and blue, respectively; the channel
between active sites (almost 100 ? long) is shown as a yellow mesh.
A glutamine molecule (green) binds to the small subunit, donating its
amido nitrogen as NH
4
H11001
in a glutamine amidotransferase–type reac-
tion. The NH
4
H11001
enters the channel, which takes it to a second active
site, where it combines with bicarbonate in a reaction requiring ATP
(bound ADP in blue). The carbamate then reenters the channel to reach
the third active site, where it is phosphorylated to carbamoyl phos-
phate (bound ADP in red).
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules
8885d_c22_833-880 2/6/04 8:35 AM Page 868 mac76 mac76:385_reb:
ATP H11001 AMP 2ADP
The ADP so formed is phosphorylated to ATP by the
glycolytic enzymes or through oxidative phosphorylation.
ATP also brings about the formation of other nu-
cleoside diphosphates by the action of a class of en-
zymes called nucleoside monophosphate kinases.
These enzymes, which are generally specific for a par-
ticular base but nonspecific for the sugar (ribose or de-
oxyribose), catalyze the reaction
ATP H11001 NMP ADP H11001 NDP
The efficient cellular systems for rephosphorylating
ADP to ATP tend to pull this reaction in the direction
of products.
Nucleoside diphosphates are converted to triphos-
phates by the action of a ubiquitous enzyme, nucleo-
side diphosphate kinase, which catalyzes the reac-
tion
NTP
D
H11001 NDP
A
NDP
D
H11001 NTP
A
This enzyme is notable in that it is not specific for the
base (purines or pyrimidines) or the sugar (ribose or
deoxyribose). This nonspecificity applies to both phos-
phate acceptor (A) and donor (D), although the donor
(NTP
D
) is almost invariably ATP, because it is present
in higher concentration than other nucleoside triphos-
phates under aerobic conditions.
Ribonucleotides Are the Precursors
of Deoxyribonucleotides
Deoxyribonucleotides, the building blocks of DNA, are
derived from the corresponding ribonucleotides by di-
rect reduction at the 2H11032-carbon atom of the D-ribose to
form the 2H11032-deoxy derivative. For example, adenosine
diphosphate (ADP) is reduced to 2H11032-deoxyadenosine
z
y
z
y
z
y
diphosphate (dADP), and GDP is reduced to dGDP. This
reaction is somewhat unusual in that the reduction oc-
curs at a nonactivated carbon; no closely analogous
chemical reactions are known. The reaction is catalyzed
by ribonucleotide reductase, best characterized in E.
coli, in which its substrates are ribonucleoside diphos-
phates.
The reduction of the D-ribose portion of a ribonu-
cleoside diphosphate to 2H11032-deoxy-D-ribose requires a
pair of hydrogen atoms, which are ultimately donated
by NADPH via an intermediate hydrogen-carrying pro-
tein, thioredoxin. This ubiquitous protein serves a sim-
ilar redox function in photosynthesis (see Fig. 20–19)
and other processes. Thioredoxin has pairs of OSH
groups that carry hydrogen atoms from NADPH to the
ribonucleoside diphosphate. Its oxidized (disulfide)
form is reduced by NADPH in a reaction catalyzed by
thioredoxin reductase (Fig. 22–39), and reduced
thioredoxin is then used by ribonucleotide reductase to
reduce the nucleoside diphosphates (NDPs) to de-
oxyribonucleoside diphosphates (dNDPs). A second
source of reducing equivalents for ribonucleotide re-
ductase is glutathione (GSH). Glutathione serves as the
reductant for a protein closely related to thioredoxin,
22.4 Biosynthesis and Degradation of Nucleotides 869
V
max
1
2
V
max
10 20 30
K
0.5
H11005 12 mM K
0.5
H11005 23 mM
[Aspartate] (mM)
Normal
activity
(no CTP)
CTP
H11001 ATP
CTP
M
/min)
V
0
(
H9262
FIGURE 22–38 Allosteric regulation of aspartate transcarbamoylase
by CTP and ATP. Addition of 0.8 mM CTP, the allosteric inhibitor of
ATCase, increases the K
0.5
for aspartate (lower curve) and the rate of
conversion of aspartate to N-carbamoylaspartate. ATP at 0.6 mM fully
reverses this effect (middle curve).
glutathione
reductase
glutaredoxin
reductase
thioredoxin
reductase
(a) (b)
dNDP NDP
S
S
HS
HS
SH
S
S
SH
S
S
SH SH
H
2
O
Ribonucleotide
reductase
Glutaredoxin Glutaredoxin Thioredoxin Thioredoxin
FADH
2
2GSH FADGSSG
NADPH H11001 H
H11001
NADP
H11001
NADP
H11001
NADPH H11001 H
H11001
Ribonucleotide
reductase
FIGURE 22–39 Reduction of ribonucleotides to deoxyribonu-
cleotides by ribonucleotide reductase. Electrons are transmitted (blue
arrows) to the enzyme from NADPH by (a) glutaredoxin or (b) thiore-
doxin. The sulfide groups in glutaredoxin reductase are contributed by
two molecules of bound glutathione (GSH; GSSG indicates oxidized
glutathione). Note that thioredoxin reductase is a flavoenzyme, with
FAD as prosthetic group.
8885d_c22_833-880 2/6/04 8:35 AM Page 869 mac76 mac76:385_reb:
glutaredoxin, which then transfers the reducing power
to ribonucleotide reductase (Fig. 22–39).
Ribonucleotide reductase is notable in that its
reaction mechanism provides the best-characterized
example of the involvement of free radicals in bio-
chemical transformations, once thought to be rare in
biological systems. The enzyme in E. coli and most eu-
karyotes is a dimer, with subunits designated R1 and
R2 (Fig. 22–40). The R1 subunit contains two kinds of
regulatory sites, as described below. The two active
sites of the enzyme are formed at the interface between
the R1 and R2 subunits. At each active site, R1 con-
tributes two sulfhydryl groups required for activity and
R2 contributes a stable tyrosyl radical. The R2 subunit
also has a binuclear iron (Fe
3H11001
) cofactor that helps gen-
erate and stabilize the tyrosyl radicals (Fig. 22–40). The
tyrosyl radical is too far from the active site to interact
directly with the site, but it generates another radical
at the active site that functions in catalysis.
A likely mechanism for the ribonucleotide reductase
reaction is illustrated in Figure 22–41. The 3H11032-ribonu-
cleotide radical formed in step 1 helps stabilize the
cation formed at the 2H11032 carbon after the loss of H
2
O
(steps 2 and 3 ). Two one-electron transfers accom-
panied by oxidation of the dithiol reduce the radical
cation (step 4 ). Step 5 is the reverse of step 1 , regenerat-
ing the active site radical (ultimately, the tyrosyl radi-
cal) and forming the deoxy product. The oxidized dithiol
is reduced to complete the cycle (step 6 ). In E. coli,
likely sources of the required reducing equivalents for
this reaction are thioredoxin and glutaredoxin, as noted
above.
Four classes of ribonucleotide reductase have been
reported. Their mechanisms (where known) generally
conform to the scheme in Figure 22–41, but they differ in
the identity of the group supplying the active-site radical
and in the cofactors used to generate it. The E. coli en-
zyme (class I) requires oxygen to regenerate the tyrosyl
radical if it is quenched, so this enzyme functions only in
an aerobic environment. Class II enzymes, found in other
microorganisms, have 5H11032-deoxyadenosylcobalamin (see
Box 17–2) rather than a binuclear iron center. Class III
enzymes have evolved to function in an anaerobic envi-
ronment. E. coli contains a separate class III ribonu-
cleotide reductase when grown anaerobically; this
enzyme contains an iron-sulfur cluster (structurally dis-
tinct from the binuclear iron center of the class I en-
zyme) and requires NADPH and S-adenosylmethionine
for activity. It uses nucleoside triphosphates rather than
nucleoside diphosphates as substrates. A class IV ribo-
nucleotide reductase, containing a binuclear manganese
center, has been reported in some microorganisms. The
evolution of different classes of ribonucleotide reductase
for production of DNA precursors in different environ-
ments reflects the importance of this reaction in nu-
cleotide metabolism.
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules870
(b)
SH
(a)
Fe
3H11001
OO
OFe
3H11001
Fe
3H11001
HS
XH HX
SH
HS
R1
subunit
R1
subunit
R2 subunit
ATP,
dGTP, dTTP
dATP,
ATP,
dATP
Primary
regulation
Active
site
site
Substrate
specificity
site
Regulatory
sites
Substrates
Allosteric
effectors
F
O
ADP,
CDP,
UDP,
GDP
e
3H11001
OH H11001O XH11001 XH
(c)
FIGURE 22–40 Ribonucleotide reductase. (a) Subunit structure.
The functions of the two regulatory sites are explained in Figure
22–42. Each active site contains two thiols and a group (OXH) that
can be converted to an active-site radical; this group is probably
the OSH of Cys
439
, which functions as a thiyl radical. (b) The R2
subunits of E. coli ribonucleotide reductase (PDB ID 1PFR). The Tyr
residue that acts as the tyrosyl radical is shown in red; the binu-
clear iron center is orange. (c) The tyrosyl radical functions to gen-
erate the active-site radical (OX
H11080
), which is used in the mechanism
shown in Figure 22–41.
8885d_c22_870 2/6/04 1:10 PM Page 870 mac76 mac76:385_reb:
22.4 Biosynthesis and Degradation of Nucleotides 871
HH
O
NDP
Thioredoxin (SH)
2
(glutaredoxin)
Thioredoxin S S
(glutaredoxin)
NDP
dNDP
dNDP
1
2
5
6
3
4
SH
X
H
CH
2
H
Base
O
H
OH
H
P OP
S
H11002
H11001
O
H
H
SH
X
H
CH
2
H
Base
O
H
OH
H
P OP
HS
O
H
3H11032 2H11032
H
SH
X
CH
2
H
Base
O
H
OH
H
P OP
HS
O
H
R2 subunit
Ribonucleotide
reductase
R1 subunit
H
H
S
X
CH
2
Base
O
H
OH
H
P OP
S
H
H
H
S
X
CH
2
Base
O
H
OH
H
P OP
S
H
S
H
CH
2
Base
O
H
OH
H
P OP
H11001
H
HX
S
H11002
MECHANISM FIGURE 22–41 Proposed mechanism for ribonu-
cleotide reductase. In the enzyme of E. coli and most eukaryotes, the
active thiol groups are on the R1 subunit; the active-site radical (OX
H11080
)
is on the R2 subunit and in E. coli is probably a thiyl radical of Cys
439
(see Fig. 22–40). Steps 1 through 6 are described in the text.
8885d_c22_871 2/6/04 1:10 PM Page 871 mac76 mac76:385_reb:
Regulation of E. coli ribonucleotide reductase is
unusual in that not only its activity but its substrate
specificity is regulated by the binding of effector mol-
ecules. Each R1 subunit has two types of regulatory site
(Fig. 22–40). One type affects overall enzyme activity
and binds either ATP, which activates the enzyme, or
dATP, which inactivates it. The second type alters sub-
strate specificity in response to the effector molecule—
ATP, dATP, dTTP, or dGTP—that is bound there (Fig.
22–42). When ATP or dATP is bound, reduction of UDP
and CDP is favored. When dTTP or dGTP is bound,
reduction of GDP or ADP, respectively, is stimulated.
The scheme is designed to provide a balanced pool of
precursors for DNA synthesis. ATP is also a general
activator for biosynthesis and ribonucleotide reduction.
The presence of dATP in small amounts increases the
reduction of pyrimidine nucleotides. An oversupply of
the pyrimidine dNTPs is signaled by high levels of
dTTP, which shifts the specificity to favor reduction of
GDP. High levels of dGTP, in turn, shift the specificity
to ADP reduction, and high levels of dATP shut the en-
zyme down. These effectors are thought to induce sev-
eral distinct enzyme conformations with altered speci-
ficities.
Thymidylate Is Derived from dCDP and dUMP
DNA contains thymine rather than uracil, and the de
novo pathway to thymine involves only deoxyribonu-
cleotides. The immediate precursor of thymidylate
(dTMP) is dUMP. In bacteria, the pathway to dUMP be-
gins with formation of dUTP, either by deamination of
dCTP or by phosphorylation of dUDP (Fig. 22–43). The
dUTP is converted to dUMP by a dUTPase. The latter
reaction must be efficient to keep dUTP pools low and
prevent incorporation of uridylate into DNA.
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules872
dCDP
dGDP
dADP
dTTP
dCTP
dGTP
dATP
Regulation at substrate-
specificity sites
Products
dUDP
dCDP
dGDP
dADP
dTTP
dCTP
dGTP
dATP
CDP
ADP
Regulation at primary
regulatory sites
Products Substrates
ATP (d)ATP
GDP
dUDPUDP
FIGURE 22–42 Regulation of ribonucleotide reductase by deoxynu-
cleoside triphosphates. The overall activity of the enzyme is affected
by binding at the primary regulatory site (left). The substrate specificity
of the enzyme is affected by the nature of the effector molecule bound
at the second type of regulatory site (right). The diagram indicates in-
hibition or stimulation of enzyme activity with the four different sub-
strates. The pathway from dUDP to dTTP is described later (see Figs
22–43, 22–44).
dCDP dCTP
UDP dUDP dUTP
dUMP
dTMP
ribonucleotide
reductase
nucleoside
diphosphate
kinase
deaminase
dUTPase
thymidylate
synthase
CDP
FIGURE 22–43 Biosynthesis of thymidylate (dTMP). The pathways
are shown beginning with the reaction catalyzed by ribonucleotide
reductase. Figure 22–44 gives details of the thymidylate synthase
reaction.
8885d_c22_833-880 2/6/04 8:35 AM Page 872 mac76 mac76:385_reb:
Conversion of dUMP to dTMP is catalyzed by thy-
midylate synthase. A one-carbon unit at the hydroxy-
methyl (OCH
2
OH) oxidation level (see Fig. 18–17) is
transferred from N
5
,N
10
-methylenetetrahydrofolate to
dUMP, then reduced to a methyl group (Fig. 22–44).
The reduction occurs at the expense of oxidation of
tetrahydrofolate to dihydrofolate, which is unusual in
tetrahydrofolate-requiring reactions. (The mechanism
of this reaction is shown in Fig. 22–50.) The dihydrofo-
late is reduced to tetrahydrofolate by dihydrofolate
reductase—a regeneration that is essential for the
many processes that require tetrahydrofolate. In plants
and at least one protist, thymidylate synthase and dihy-
drofolate reductase reside on a single bifunctional protein.
Degradation of Purines and Pyrimidines Produces
Uric Acid and Urea, Respectively
Purine nucleotides are degraded by a pathway in which
they lose their phosphate through the action of 5H11541-
nucleotidase (Fig. 22–45). Adenylate yields adenosine,
which is deaminated to inosine by adenosine deami-
nase, and inosine is hydrolyzed to hypoxanthine (its
purine base) and D-ribose. Hypoxanthine is oxidized
successively to xanthine and then uric acid by xanthine
oxidase, a flavoenzyme with an atom of molybdenum
and four iron-sulfur centers in its prosthetic group. Mol-
ecular oxygen is the electron acceptor in this complex
reaction.
22.4 Biosynthesis and Degradation of Nucleotides 873
FIGURE 22–44 Conversion of dUMP to dTMP by thymidylate syn-
thase and dihydrofolate reductase. Serine hydroxymethyltransferase is
required for regeneration of the N
5
,N
10
-methylene form of tetrahy-
drofolate. In the synthesis of dTMP, all three hydrogens of the added
methyl group are derived from N
5
,N
10
-methylenetetrahydrofolate (pink
and gray).
O
CH
2
OH
N
N
H
H
H
C
H
CH
2
dTMP
O
P
O
HN
O
OH
O
N
H
H
H
H
H
CH
2
dUMP
O
P
O
HN
H
H
O
O
N
H
2
N
H
N
H
HN
CH
2
N
H
H
O
NH
2
N
C
HN
H
H
H
N
H
HN
R
N
R
thymidylate
synthase
N
5
,N
10
-Methylene-
tetrahydrofolate
7,8-Dihydrofolate
Glycine
Serine
NADPH H11001 H
H11001
H11001
NADP
serine
hydroxymethyl-
transferase
dihydrofolate
reductase
Tetrahydrofolate
CH
2
N
H
O
NH
2
N
HN
H
N
HN R
PLP
8885d_c22_873 2/6/04 1:10 PM Page 873 mac76 mac76:385_reb:
N
H
C
GMP
Guanosine
nucleosidase
AMP
guanine
deaminase
Adenosine
Inosine
Hypoxanthine
(keto form)
xanthine
oxidase
Xanthine
(keto form)
Uric acid
H
2
O
Guanine
nucleosidase
H
2
O
2
O
5H11032-nucleotidase
H
2
O
P
i
5H11032-nucleotidase
H
2
O
P
i
NH
3
H
2
O
Ribose
O
HN
N
N
N
N
H
Ribose
H
2
O
H
2
O H11001 O
2
H
O
HO
xanthine
oxidase
OH
HO
HN
4NH
4
H11001
O
adenosine
deaminase
NH
2
H
2
O
C
C
C
C
C
N
N
H
N
C
O
H
2
O
2
H
2
O H11001 O
2
allantoicase
N
H
OH Uric acid
Allantoin
C
O
H
N
C
H
C
NH
3
O
H
2
O
allantoinase
NH
2
NH
2
C
ON
H
C
H
COO
H11002
Allantoate
urease
COO
H11002
CHO
Glyoxylate
NH
2
C
O
H
2
O
CO
2
urate oxidase
2H
2
N
2H
2
O
2CO
2
Urea
Excreted by:
Primates, birds,
reptiles, insects
Most mammals
Bony fishes
Amphibians,
cartilaginous
fishes
Marine
invertebrates
O
2
H11001 H
2
O
H
2
N
O
HN
N
N
N
N
N
H
N
N
N
H
HO
HN
O
N
H
HN
2
1
FIGURE 22–45 Catabolism of purine nucleotides. Note that primates
excrete much more nitrogen as urea via the urea cycle (Chapter 18) than
as uric acid from purine degradation. Similarly, fish excrete much more
nitrogen as NH
4
H11001
than as urea produced by the pathway shown here.
GMP catabolism also yields uric acid as end prod-
uct. GMP is first hydrolyzed to guanosine, which is then
cleaved to free guanine. Guanine undergoes hydrolytic
removal of its amino group to yield xanthine, which is
converted to uric acid by xanthine oxidase (Fig. 22–45).
Uric acid is the excreted end product of purine ca-
tabolism in primates, birds, and some other animals. A
healthy adult human excretes uric acid at a rate of about
0.6 g/24 h; the excreted product arises in part from in-
gested purines and in part from turnover of the purine
nucleotides of nucleic acids. In most mammals and many
other vertebrates, uric acid is further degraded to al-
lantoin by the action of urate oxidase. In other or-
ganisms the pathway is further extended, as shown in
Figure 22–45.
The pathways for degradation of pyrimidines gen-
erally lead to NH
4
H11001
production and thus to urea syn-
thesis. Thymine, for example, is degraded to methyl-
malonylsemialdehyde (Fig. 22–46), an intermediate
of valine catabolism. It is further degraded through
propionyl-CoA and methylmalonyl-CoA to succinyl-
CoA (see Fig. 18–27).
Genetic aberrations in human purine metabolism
have been found, some with serious conse-
quences. For example, adenosine deaminase (ADA)
deficiency leads to severe immunodeficiency disease
in which T lymphocytes and B lymphocytes do not de-
velop properly. Lack of ADA leads to a 100-fold increase
in the cellular concentration of dATP, a strong inhibitor
of ribonucleotide reductase (Fig. 22–42). High levels
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules874
8885d_c22_833-880 2/6/04 8:35 AM Page 874 mac76 mac76:385_reb:
NADPH H11001 H
H11001
O
HN
C
N
H
C
C
O
NADP
H11001
O
HN C
CH
N
H
CH
2
C
C
O
H
H
2
N
CH
3
Thymine
dihydrouracil
dehydrogenase
Dihydrothymine
dihydropyrimidinase
H9252-Ureidoisobutyrate
CH
3
C
O
NH
CH
2
CH
CH
3
C
O
O
H11002
H
2
O
H
3
N
H11002
CH
2
CH
C
O
O
H11002
CH
3
aminotransferase
H9252-Aminoisobutyrate
H
2
O
H9252-ureidopropionase
H9251-Ketoglutarate
Glutamate
CH
H
3
C
C
O
H11002
O
O
H
Methylmalonyl-
semialdehyde
H11001
NH
4
H11001 HCO
3
H11001
FIGURE 22–46 Catabolism of a pyrimidine. Shown here is the path-
way for thymine. The methylmalonylsemialdehyde is further degraded
to succinyl-CoA.
of dATP produce a general deficiency of other dNTPs
in T lymphocytes. The basis for B-lymphocyte toxicity
is less clear. Individuals with ADA deficiency lack an ef-
fective immune system and do not survive unless iso-
lated in a sterile “bubble” environment. ADA deficiency
is one of the first targets of human gene therapy trials
(see Box 9–2). ■
Purine and Pyrimidine Bases Are Recycled
by Salvage Pathways
Free purine and pyrimidine bases are constantly re-
leased in cells during the metabolic degradation of nu-
cleotides. Free purines are in large part salvaged and
reused to make nucleotides, in a pathway much simpler
than the de novo synthesis of purine nucleotides de-
scribed earlier. One of the primary salvage pathways
consists of a single reaction catalyzed by adenosine
phosphoribosyltransferase, in which free adenine
reacts with PRPP to yield the corresponding adenine
nucleotide:
Adenine H11001 PRPP On AMP H11001 PP
i
Free guanine and hypoxanthine (the deamination prod-
uct of adenine; Fig. 22–45) are salvaged in the same
way by hypoxanthine-guanine phosphoribosyltrans-
ferase. A similar salvage pathway exists for pyrimidine
bases in microorganisms, and possibly in mammals.
A genetic lack of hypoxanthine-guanine phos-
phoribosyltransferase activity, seen almost ex-
clusively in male children, results in a bizarre set of
symptoms called Lesch-Nyhan syndrome. Children
with this genetic disorder, which becomes manifest by
the age of 2 years, are sometimes poorly coordinated
and mentally retarded. In addition, they are extremely
hostile and show compulsive self-destructive tenden-
cies: they mutilate themselves by biting off their fingers,
toes, and lips.
The devastating effects of Lesch-Nyhan syndrome
illustrate the importance of the salvage pathways. Hy-
poxanthine and guanine arise constantly from the break-
down of nucleic acids. In the absence of hypoxanthine-
guanine phosphoribosyltransferase, PRPP levels rise
and purines are overproduced by the de novo pathway,
resulting in high levels of uric acid production and gout-
like damage to tissue (see below). The brain is espe-
cially dependent on the salvage pathways, and this may
account for the central nervous system damage in chil-
dren with Lesch-Nyhan syndrome. This syndrome is
another target of early trials in gene therapy (see Box
9–2). ■
Excess Uric Acid Causes Gout
Long thought, erroneously, to be due to “high liv-
ing,” gout is a disease of the joints caused by an
elevated concentration of uric acid in the blood and tis-
sues. The joints become inflamed, painful, and arthritic,
owing to the abnormal deposition of sodium urate crys-
tals. The kidneys are also affected, as excess uric acid
is deposited in the kidney tubules. Gout occurs pre-
dominantly in males. Its precise cause is not known, but
it often involves an underexcretion of urate. A genetic
deficiency of one or another enzyme of purine metabo-
lism may also be a factor in some cases.
22.4 Biosynthesis and Degradation of Nucleotides 875
8885d_c22_833-880 2/6/04 8:35 AM Page 875 mac76 mac76:385_reb:
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules876
CH
2
NH
3
O
H
N
H11002
H11001
NH
2
C
CH
NH
3
H
COO
H11002
H11001
C
Acivicin
O
C
O
CH
2
COO
N
H11001
Glutamine
C
N
H11001
CH
2
C
H
Cl
COO
H11002
CH
C
O
CH
2
NH
3
Azaserine
FIGURE 22–48 Azaserine and acivicin, inhibitors of glutamine ami-
dotransferases. These analogs of glutamine interfere in a number of
amino acid and nucleotide biosynthetic pathways.
Gout is effectively treated by a combination of nu-
tritional and drug therapies. Foods especially rich in nu-
cleotides and nucleic acids, such as liver or glandular
products, are withheld from the diet. Major alleviation of
the symptoms is provided by the drug allopurinol (Fig.
22–47), which inhibits xanthine oxidase, the enzyme that
catalyzes the conversion of purines to uric acid. Allop-
urinol is a substrate of xanthine oxidase, which converts
allopurinol to oxypurinol (alloxanthine). Oxypurinol in-
activates the reduced form of the enzyme by remaining
tightly bound in its active site. When xanthine oxidase
is inhibited, the excreted products of purine metabolism
are xanthine and hypoxanthine, which are more water-
soluble than uric acid and less likely to form crystalline
deposits. Allopurinol was developed by Gertrude Elion
and George Hitchings, who also developed acyclovir,
used in treating people with AIDS, and other purine
analogs used in cancer chemotherapy. ■
Many Chemotherapeutic Agents Target Enzymes
in the Nucleotide Biosynthetic Pathways
The growth of cancer cells is not controlled in the
same way as cell growth in most normal tissues.
Cancer cells have greater requirements for nucleotides
as precursors of DNA and RNA, and consequently are
generally more sensitive than normal cells to inhibitors
of nucleotide biosynthesis. A growing array of important
chemotherapeutic agents—for cancer and other dis-
eases—act by inhibiting one or more enzymes in these
pathways. We describe here several well-studied exam-
ples that illustrate productive approaches to treatment
and help us understand how these enzymes work.
The first set of agents includes compounds that in-
hibit glutamine amidotransferases. Recall that glutamine
is a nitrogen donor in at least half a dozen separate re-
actions in nucleotide biosynthesis. The binding sites for
glutamine and the mechanism by which NH
4
H11001
is ex-
tracted are quite similar in many of these enzymes. Most
are strongly inhibited by glutamine analogs such as aza-
serine and acivicin (Fig. 22–48). Azaserine, charac-
terized by John Buchanan in the 1950s, was one of the
first examples of a mechanism-based enzyme inactiva-
tor (suicide inactivator; p. 211 and Box 22–2). Acivicin
shows promise as a cancer chemotherapeutic agent.
Other useful targets for pharmaceutical agents are
thymidylate synthase and dihydrofolate reductase, en-
zymes that provide the only cellular pathway for
thymine synthesis (Fig. 22–49). One inhibitor that acts
on thymidylate synthase, fluorouracil, is an important
chemotherapeutic agent. Fluorouracil itself is not the
enzyme inhibitor. In the cell, salvage pathways convert
it to the deoxynucleoside monophosphate FdUMP,
which binds to and inactivates the enzyme. Inhibition
by FdUMP (Fig. 22–50) is a classic example of mecha-
nism-based enzyme inactivation. Another prominent
chemotherapeutic agent, methotrexate, is an inhibitor
of dihydrofolate reductase. This folate analog acts as a
competitive inhibitor; the enzyme binds methotrexate
with about 100 times higher affinity than dihydrofolate.
Aminopterin is a related compound that acts similarly.
OH
N
HC
N
N
H
C
C
N
Hypoxanthine
(enol form)
H
C
N
C
HC
N
C
C
N
Allopurinol
OH
N
H
CH
C
H
C
N
C
N
C
C
N
Oxypurinol
OH
N
H
C
HO
xanthine
oxidase
FIGURE 22–47 Allopurinol, an inhibitor of xanthine
oxidase. Hypoxanthine is the normal substrate of
xanthine oxidase. Only a slight alteration in the
structure of hypoxanthine (shaded pink) yields the
medically effective enzyme inhibitor allopurinol. At
the active site, allopurinol is converted to oxypuri-
nol, a strong competitive inhibitor that remains
tightly bound to the reduced form of the enzyme.
Gertrude Elion (1918–1999) and
George Hitchings (1905–1998)
8885d_c22_833-880 2/6/04 8:35 AM Page 876 mac76 mac76:385_reb:
22.4 Biosynthesis and Degradation of Nucleotides 877
HN
N
H
F
NH
2
CH
CH
3
COO
H11002
O
Methotrexate
CH
2
O
NH
O
H
N
C
H
2
CH
2
N
N
N
N
N
CH
2
COO
H11002
Fluorouracil
(b)(a)
10
5
dUMP dTMP
FdUMP
Methotrexate
Aminopterin
Trimethoprim
Serine
PLP
Glycine
NADP
H11001
thymidylate
synthase
dihydrofolate
reductase
serine
hydroxymethyl-
transferase
O
Trimethoprim
CH
3
H
3
CO
H
3
CO NH
2
NH
2
N
N
N
5
, N
10
-Methylene
H
4
folate
7,8-Dihydrofolate
H
4
folate
NADPH H11001 H
H11001
FIGURE 22–49 Thymidylate synthesis and folate metabolism
as targets of chemotherapy. (a) During thymidylate synthesis,
N
5
,N
10
-methylenetetrahydrofolate is converted to 7,8-dihydrofolate;
the N
5
,N
10
-methylenetetrahydrofolate is regenerated in two steps (see
Fig. 22–44). This cycle is a major target of several chemotherapeutic
agents. (b) Fluorouracil and methotrexate are important chemothera-
peutic agents. In cells, fluorouracil is converted to FdUMP, which
inhibits thymidylate synthase. Methotrexate, a structural analog of
tetrahydrofolate, inhibits dihydrofolate reductase; the shaded amino
and methyl groups replace a carbonyl oxygen and a proton, respec-
tively, in folate (see Fig. 22–44). Another important folate analog,
aminopterin, is identical to methotrexate except that it lacks the shaded
methyl group. Trimethoprim, a tight-binding inhibitor of bacterial di-
hydrofolate reductase, was developed as an antibiotic.
MECHANISM FIGURE 22–50 Conversion of dUMP to dTMP and its
inhibition by FdUMP. The top row is the normal reaction mechanism
of thymidylate synthase. The nucleophilic sulfhydryl group contributed
by the enzyme in step 1 and the ring atoms of dUMP taking part in
the reaction are shown in red; :B denotes an amino acid side chain
that acts as a base to abstract a proton in step 3 . The hydrogens de-
rived from the methylene group of N
5
,N
10
-methylenetetrahydrofolate
are shaded in gray. A novel feature of this reaction mechanism is a
1,3 hydride shift (step 3 ), which moves a hydride ion (shaded pink)
from C-6 of H
4
folate to the methyl group of thymidine, resulting in
the oxidation of tetrahydrofolate to dihydrofolate. It is this hydride shift
that apparently does not occur when FdUMP is the substrate (bottom).
Steps 1 and 2 proceed normally, but result in a stable complex—
with FdUMP linked covalently to the enzyme and to tetrahydrofolate—
that inactivates the enzyme. Thymidylate Synthase Mechanism
O
HN
N
R
O
H
CH
3
FdUMP
N
5
,N
10
-Methylene
H
4
folate
O
FHN C
N
RH11032
R
C H
O
N
H
H
C
CH
2
O
HN
N
RH11032
R
C H
O
N
H
H
C
CH
2
O
FHN C
N
R
C H
O
H
H
C
1,3
hydride
shift
Enzyme H11001
dihydrofolate
O
HHN
C
N
R
C H
O
dTMP
dUMP
O
HHN C
N
RH11032
R
C H
O
N
H
H
C
CH
2
B
B B B
B B B
10
O
HN
N
RH11032
R
C H
O
N
H
H
C
CH
2
S
O
HHN C
N
CH
2
R
C H
O
H
C
H
H
N
C
S
S
CH
2
C
H
N
CH
2
C
H
N
CH
2
C
H
N
5
CH
2
C
H
N
CH
2
C
H
N
CH
2
C
H
N
C
H
H
FC
H5008
S
S
Dead-end
covalent
complex
RH11032
NH
CH
2
SH
SH
C
H5008
RH11032
NH
CH
2
RH11032
NH
CH
2
1 2
1 2
3
8885d_c22_877 2/6/04 1:10 PM Page 877 mac76 mac76:385_reb:
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules878
Key Terms
nitrogen cycle 834
nitrogen fixation 834
symbionts 834
nitrogenase complex 835
leghemoglobin 836
glutamine synthetase 838
glutamate synthase 838
glutamine amidotransferases 840
5-phosphoribosyl-1-
pyrophosphate (PRPP) 842
tryptophan synthase 849
porphyrin 854
porphyria 854
bilirubin 854
phosphocreatine 857
creatine 857
glutathione (GSH) 857
auxin 859
dopamine 859
norepinephrine 859
epinephrine 859
H9253-aminobutyrate (GABA) 859
serotonin 859
histamine 859
cimetidine 859
spermine 860
spermidine 860
ornithine decarboxylase 860
de novo pathway 862
salvage pathway 862
inosinate (IMP) 866
carbamoyl phosphate synthetase II
868
aspartate transcarbamoylase 868
nucleoside monophosphate kinase
869
nucleoside diphosphate kinase
869
ribonucleotide reductase 869
thioredoxin 869
thymidylate synthase 873
dihydrofolate reductase 873
adenosine deaminase
deficiency 874
Lesch-Nyhan syndrome 875
allopurinol 876
azaserine 876
acivicin 876
fluorouracil 876
methotrexate 876
aminopterin 876
Terms in bold are defined in the glossary.
Further Reading
Nitrogen Fixation
Burris, R.H. (1995) Breaking the N–N bond. Annu. Rev. Plant
Physiol. Plant Mol. Biol. 46, 1–19.
Igarishi, R.Y. & Seefeldt, L.C. (2003) Nitrogen fixation: the
mechanism of the Mo-dependent nitrogenase. Crit. Rev. Biochem.
Mol. Biol., 38, 351–384.
The medical potential of inhibitors of nucleotide
biosynthesis is not limited to cancer treatment. All fast-
growing cells (including bacteria and protists) are po-
tential targets. Trimethoprim, an antibiotic developed
by Hitchings and Elion, binds to bacterial dihydrofolate
reductase nearly 100,000 times better than to the mam-
malian enzyme. It is used to treat certain urinary and
middle ear bacterial infections. Parasitic protists, such
as the trypanosomes that cause African sleeping sickness
(African trypanosomiasis), lack pathways for de novo
nucleotide biosynthesis and are particularly sensitive to
agents that interfere with their scavenging of nucleo-
tides from the surrounding environment using salvage
pathways. Allopurinol (Fig. 22–47) and a number of re-
lated purine analogs have shown promise for the treat-
ment of African trypanosomiasis and related afflictions.
See Box 22–2 for another approach to combating African
trypanosomiasis, made possible by advances in our un-
derstanding of metabolism and enzyme mechanism. ■
SUMMARY 22.4 Biosynthesis and Degradation
of Nucleotides
■ The purine ring system is built up step-by-step
beginning with 5-phosphoribosylamine. The
amino acids glutamine, glycine, and aspartate
furnish all the nitrogen atoms of purines. Two
ring-closure steps form the purine nucleus.
■ Pyrimidines are synthesized from carbamoyl
phosphate and aspartate, and ribose
5-phosphate is then attached to yield the
pyrimidine ribonucleotides.
■ Nucleoside monophosphates are converted to
their triphosphates by enzymatic phosphory-
lation reactions. Ribonucleotides are converted
to deoxyribonucleotides by ribonucleotide re-
ductase, an enzyme with novel mechanistic and
regulatory characteristics. The thymine nu-
cleotides are derived from dCDP and dUMP.
■ Uric acid and urea are the end products of
purine and pyrimidine degradation.
■ Free purines can be salvaged and rebuilt into
nucleotides. Genetic deficiencies in certain
salvage enzymes cause serious disorders such
as Lesch-Nyhan syndrome and ADA deficiency.
■ Accumulation of uric acid crystals in the joints,
possibly caused by another genetic deficiency,
results in gout.
■ Enzymes of the nucleotide biosynthetic pathways
are targets for an array of chemotherapeutic
agents used to treat cancer and other diseases.
8885d_c22_878 2/6/04 2:00 PM Page 878 mac76 mac76:385_reb:
Chapter 22 Problems 879
Patriarca, E.J., Tate, R., & Iaccarino, M. (2002) Key role of
bacterial NH
4
H11001
metabolism in rhizobium-plant symbiosis. Microbiol.
Mol. Biol. Rev. 66, 203–222.
A good overview of ammonia assimilation in bacterial systems
and its regulation.
Sinha, S.C. & Smith, J.L. (2001) The PRT protein family. Curr.
Opin. Struct. Biol. 11, 733–739.
Description of a protein family that includes many amidotrans-
ferases, with channels for the movement of NH
3
from one
active site to another.
Ye, R.W. & Thomas, S.M. (2001) Microbial nitrogen cycles:
physiology, genomics and applications. Curr. Opin. Microbiol. 4,
307–312.
Amino Acid Biosynthesis
Abeles, R.H., Frey, P.A., & Jencks, W.P. (1992) Biochemistry,
Jones and Bartlett Publishers, Boston.
This book includes excellent accounts of reaction mechanisms,
including one-carbon metabolism and pyridoxal phosphate
enzymes.
Bender, D.A. (1985) Amino Acid Metabolism, 2nd edn, Wiley-
Interscience, New York.
Neidhardt, F.C. (ed.) (1996) Escherichia coli and Salmonella:
Cellular and Molecular Biology, 2nd edn, ASM Press, Washing-
ton, DC.
Volume 1 of this two-volume set has 13 chapters devoted to
detailed descriptions of amino acid and nucleotide biosynthesis
in bacteria. The web-based version, EcoSal, is updated regularly.
A valuable resource.
Pan P., Woehl, E., & Dunn, M.F. (1997) Protein architecture,
dynamics and allostery in tryptophan synthase channeling. Trends
Biochem. Sci. 22, 22–27.
Compounds Derived from Amino Acids
Bredt, D.S. & Snyder, S.H. (1994) Nitric oxide: a physiologic
messenger molecule. Annu. Rev. Biochem. 63, 175–195.
Meister, A. & Anderson, M.E. (1983) Glutathione. Annu. Rev.
Biochem. 52, 711–760.
Morse, D. & Choi, A.M.K. (2002) Heme oxygenase-1—the
“emerging molecule” has arrived. Am. J. Resp. Cell Mol. Biol. 27,
8–16.
Rondon, M.R., Trzebiatowski, J.R., & Escalante-Semerena,
J.C. (1997) Biochemistry and molecular genetics of cobalamin
biosynthesis. Prog. Nucleic Acid Res. Mol. Biol. 56, 347–384.
Stadtman, T.C. (1996) Selenocysteine. Annu. Rev. Biochem. 65,
83–100.
Nucleotide Biosynthesis
Blakley, R.L. & Benkovic, S.J. (1985) Folates and Pterins, Vol.
2: Chemistry and Biochemistry of Pterins, Wiley-Interscience,
New York.
Carreras, C.W. & Santi, D.V. (1995) The catalytic mechanism
and structure of thymidylate synthase. Annu. Rev. Biochem. 64,
721–762.
Eliasson, R., Pontis, E., Sun, X., & Reichard, P. (1994)
Allosteric control of the substrate specificity of the anaerobic
ribonucleotide reductase from Escherichia coli. J. Biol. Chem.
269, 26,052–26,057.
Holmgren, A. (1989) Thioredoxin and glutaredoxin systems. J.
Biol. Chem. 264, 13,963–13,966.
Jordan, A. & Reichard P. (1998) Ribonucleotide reductases.
Annu. Rev. Biochem. 67, 71–98.
Kappock, T.J., Ealick, S.E., & Stubbe, J. (2000) Modular
evolution of the purine biosynthetic pathway. Curr. Opin. Chem.
Biol. 4, 567–572.
Kornberg, A. & Baker, T.A. (1991) DNA Replication, 2nd edn,
W. H. Freeman and Company, New York.
This text includes a good summary of nucleotide biosynthesis.
Lee, L., Kelly, R.E., Pastra-Landis, S.C., & Evans, D.R.
(1985) Oligomeric structure of the multifunctional protein CAD
that initiates pyrimidine biosynthesis in mammalian cells. Proc.
Natl. Acad. Sci. USA 82, 6802–6806.
Licht, S., Gerfen, G.J., & Stubbe, J. (1996) Thiyl radicals in
ribonucleotide reductases. Science 271, 477–481.
Schachman, H.K. (2000) Still looking for the ivory tower. Annu.
Rev. Biochem. 69, 1–29.
A lively description of research on aspartate transcarbamoylase,
accompanied by delightful tales of science and politics.
Stubbe, J. & Riggs-Gelasco, P. (1998) Harnessing free radicals:
formation and function of the tyrosyl radical in ribonucleotide
reductase. Trends Biochem. Sci. 23, 438–443.
Villafranca, J.E., Howell, E.E., Voet, D.H., Strobel, M.S.,
Ogden, R.C., Abelson, J.N., & Kraut, J. (1983) Directed
mutagenesis of dihydrofolate reductase. Science 222, 782–788.
A report of structural studies on this important enzyme.
Genetic Diseases
Scriver, C.R., Beaudet, A.L., Valle, D., Sly, W.S., Childs, B.,
Kinzler, L.W., & Vogelstein, B. (eds) (2001) The Metabolic
and Molecular Bases of Inherited Disease, 8th edn, McGraw-Hill
Professional, New York.
This four-volume set has good chapters on disorders of amino
acid, porphyrin, and heme metabolism. See also the chapters
on inborn errors of purine and pyrimidine metabolism.
1. ATP Consumption by Root Nodules in Legumes
Bacteria residing in the root nodules of the pea plant con-
sume more than 20% of the ATP produced by the plant. Sug-
gest why these bacteria consume so much ATP.
2. Glutamate Dehydrogenase and Protein Synthesis
The bacterium Methylophilus methylotrophus can synthe-
size protein from methanol and ammonia. Recombinant DNA
techniques have improved the yield of protein by introduc-
ing into M. methylotrophus the glutamate dehydrogenase
gene from E. coli. Why does this genetic manipulation in-
crease the protein yield?
Problems
8885d_c22_833-880 2/6/04 8:35 AM Page 879 mac76 mac76:385_reb:
Chapter 22 Biosynthesis of Amino Acids, Nucleotides, and Related Molecules880
3. Transformation of Aspartate to Asparagine There
are two routes for transforming aspartate to asparagine at the
expense of ATP. Many bacteria have an asparagine synthetase
that uses ammonium ion as the nitrogen donor. Mammals have
an asparagine synthetase that uses glutamine as the nitrogen
donor. Given that the latter requires an extra ATP (for the
synthesis of glutamine), why do mammals use this route?
4. Equation for the Synthesis of Aspartate from Glu-
cose Write the net equation for the synthesis of aspartate
(a nonessential amino acid) from glucose, carbon dioxide, and
ammonia.
5. Phenylalanine Hydroxylase Deficiency and Diet
Tyrosine is normally a nonessential amino acid, but individu-
als with a genetic defect in phenylalanine hydroxylase require
tyrosine in their diet for normal growth. Explain.
6. Cofactors for One-Carbon Transfer Reactions
Most one-carbon transfers are promoted by one of three co-
factors: biotin, tetrahydrofolate, or S-adenosylmethionine
(Chapter 18). S-Adenosylmethionine is generally used as a
methyl group donor; the transfer potential of the methyl
group in N
5
-methyltetrahydrofolate is insufficient for most
biosynthetic reactions. However, one example of the use of
N
5
-methyltetrahydrofolate in methyl group transfer is in me-
thionine formation by the methionine synthase reaction (step
9 of Fig. 22–15); methionine is the immediate precursor of
S-adenosylmethionine (see Fig. 18–18). Explain how the
methyl group of S-adenosylmethionine can be derived from
N
5
-methyltetrahydrofolate, even though the transfer poten-
tial of the methyl group in N
5
-methyltetrahydrofolate is one-
thousandth of that in S-adenosylmethionine.
7. Concerted Regulation in Amino Acid Biosynthesis
The glutamine synthetase of E. coli is independently modu-
lated by various products of glutamine metabolism (see Fig.
22–6). In this concerted inhibition, the extent of enzyme in-
hibition is greater than the sum of the separate inhibitions
caused by each product. For E. coli grown in a medium rich
in histidine, what would be the advantage of concerted inhi-
bition?
8. Relationship between Folic Acid Deficiency
and Anemia Folic acid deficiency, believed to be
the most common vitamin deficiency, causes a type of ane-
mia in which hemoglobin synthesis is impaired and erythro-
cytes do not mature properly. What is the metabolic rela-
tionship between hemoglobin synthesis and folic acid
deficiency?
9. Nucleotide Biosynthesis in Amino Acid Auxotrophic
Bacteria Normal E. coli cells can synthesize all 20 com-
mon amino acids, but some mutants, called amino acid aux-
otrophs, are unable to synthesize a specific amino acid and
require its addition to the culture medium for optimal growth.
Besides their role in protein synthesis, some amino acids are
also precursors for other nitrogenous cell products. Consider
the three amino acid auxotrophs that are unable to synthe-
size glycine, glutamine, and aspartate, respectively. For each
mutant, what nitrogenous products other than proteins would
the cell fail to synthesize?
10. Inhibitors of Nucleotide Biosynthesis Suggest
mechanisms for the inhibition of (a) alanine racemase by L-
fluoroalanine and (b) glutamine amidotransferases by aza-
serine.
11. Mode of Action of Sulfa Drugs Some bacte-
ria require p-aminobenzoate in the culture medium
for normal growth, and their growth is severely inhibited by
the addition of sulfanilamide, one of the earliest sulfa drugs.
Moreover, in the presence of this drug, 5-aminoimidazole-4-
carboxamide ribonucleotide (AICAR; see Fig. 22–33) accu-
mulates in the culture medium. These effects are reversed by
addition of excess p-aminobenzoate.
(a) What is the role of p-aminobenzoate in these bacte-
ria? (Hint: See Fig. 18–16).
(b) Why does AICAR accumulate in the presence of sul-
fanilamide?
(c) Why are the inhibition and accumulation reversed by
addition of excess p-aminobenzoate?
12. Pathway of Carbon in Pyrimidine Biosynthesis
Predict the locations of
14
C in orotate isolated from cells
grown on a small amount of uniformly labeled [
14
C]succinate.
Justify your prediction.
13. Nucleotides As Poor Sources of Energy Under
starvation conditions, organisms can use proteins and amino
acids as sources of energy. Deamination of amino acids pro-
duces carbon skeletons that can enter the glycolytic pathway
and the citric acid cycle to produce energy in the form of
ATP. Nucleotides, on the other hand, are not similarly de-
graded for use as energy-yielding fuels. What observations
about cellular physiology support this statement? What as-
pect of the structure of nucleotides makes them a relatively
poor source of energy?
14. Treatment of Gout Allopurinol (see Fig.
22–47), an inhibitor of xanthine oxidase, is used to
treat chronic gout. Explain the biochemical basis for this
treatment. Patients treated with allopurinol sometimes de-
velop xanthine stones in the kidneys, although the incidence
of kidney damage is much lower than in untreated gout. Ex-
plain this observation in the light of the following solubilities
in urine: uric acid, 0.15 g/L; xanthine, 0.05 g/L; and hypox-
anthine, 1.4 g/L.
15. Inhibition of Nucleotide Synthesis by Azaserine
The diazo compound O-(2-diazoacetyl)-L-serine, known also
as azaserine (see Fig. 22–48), is a powerful inhibitor of glut-
amine amidotransferases. If growing cells are treated with
azaserine, what intermediates of nucleotide biosynthesis
would accumulate? Explain.
O
H5008
O
SCH
2
NH
2
O
O
NNH
2
p-Aminobenzoate Sulfanilamide
8885d_c22_833-880 2/6/04 8:35 AM Page 880 mac76 mac76:385_reb:
chapter
I
n Chapters 13 through 22 we have discussed metabo-
lism at the level of the individual cell, emphasizing cen-
tral pathways common to almost all cells, prokaryotic
and eukaryotic. We have seen how metabolic processes
within cells are regulated at the level of individual en-
zyme reactions, by substrate availability, by allosteric
mechanisms, and by phosphorylation or other covalent
modifications of enzymes.
To appreciate fully the significance of individual
metabolic pathways and their regulation, we must view
these pathways in the context of the whole organism.
An essential characteristic of multicellular organisms is
cell differentiation and division of labor. The specialized
functions of the tissues and organs of complex organ-
isms such as humans impose characteristic fuel re-
quirements and patterns of metabolism. Hormonal sig-
nals integrate and coordinate the metabolic activities of
different tissues and optimize the allocation of fuels and
precursors to each organ.
In this chapter we focus on mammals, looking at the
specialized metabolism of several major organs and tis-
sues and the integration of metabolism in the whole or-
ganism. We begin by examining the broad range of hor-
mones and hormonal mechanisms, then turn to the
tissue-specific functions regulated by these mecha-
nisms. We discuss the distribution of nutrients to vari-
ous organs—emphasizing the central role played by the
liver—and the metabolic cooperation among these or-
gans. To illustrate the integrative role of hormones, we
describe the interplay of insulin, glucagon, and epi-
nephrine in coordinating fuel metabolism in muscle,
liver, and adipose tissue. The metabolic disturbances in
diabetes further illustrate the importance of hormonal
regulation of metabolism. Finally we discuss the long-
term hormonal regulation of body mass.
23.1 Hormones: Diverse Structures
for Diverse Functions
Virtually every process in a complex organism is regu-
lated by one or more hormones: maintenance of blood
pressure, blood volume, and electrolyte balance; em-
bryogenesis; sexual differentiation, development, and
reproduction; hunger, eating behavior, digestion, and
HORMONAL REGULATION
AND INTEGRATION OF
MAMMALIAN METABOLISM
23.1 Hormones: Diverse Structures for Diverse
Functions 881
23.2 Tissue-Specific Metabolism: The Division
of Labor 892
23.3 Hormonal Regulation of Fuel Metabolism 902
23.4 Obesity and the Regulation of Body Mass 910
We recognize that each tissue and, more generally, each
cell of the organism secretes . . . special products or
ferments into the blood which thereby influence all the
other cells thus integrated with each other by a
mechanism other than the nervous system.
—Charles édouard Brown-Séquard and J. d’Arsonval, article in
Comptes Rendus de la Société de Biologie, 1891
23
881
8885d_c23_881-919 3/1/04 1:26 PM Page 881 mac76 mac76:385_reb:
fuel allocation—to name but a few. We examine here the
methods for detecting and measuring hormones and
their interaction with receptors, and consider a repre-
sentative selection of hormone types.
The coordination of metabolism in mammals is
achieved by the neuroendocrine system. Individual
cells in one tissue sense a change in the organism’s cir-
cumstances and respond by secreting a chemical mes-
senger that passes to another cell in the same or dif-
ferent tissue, where it binds to a receptor molecule and
triggers a change in this second cell. In neuronal sig-
naling (Fig. 23–1a), the chemical messenger (neuro-
transmitter; acetylcholine, for example) may travel only
a fraction of a micrometer, across the synaptic cleft to
the next neuron in a network. In hormonal signaling, the
messengers—hormones—are carried in the bloodstream
to neighboring cells or to distant organs and tissues;
they may travel a meter or more before encountering
their target cell (Fig. 23–1b). Except for this anatomic
difference, these two chemical signaling mechanisms are
remarkably similar. Epinephrine and norepinephrine,
for example, serve as neurotransmitters in certain
synapses of the brain and smooth muscle and as hor-
mones that regulate fuel metabolism in liver and muscle.
The following discussion of cellular signaling emphasizes
hormone action, drawing on discussions of fuel metab-
olism in earlier chapters, but most of the fundamental
mechanisms described here also occur in neurotrans-
mitter action.
The Discovery and Purification of Hormones
Requires a Bioassay
How is a hormone discovered and isolated? First, re-
searchers find that a physiological process in one tissue
depends on a signal that originates in another tissue. In-
sulin, for example, was first recognized as a substance that
is produced in the pancreas and affects the volume and
composition of urine (Box 23–1). Once a physiological ef-
fect of the putative hormone is discovered, a quantitative
bioassay for the hormone can be developed. In the case
of insulin, the assay consisted of injecting extracts of pan-
creas (a crude source of insulin) into experimental ani-
mals deficient in insulin, then quantifying the resulting
changes in glucose concentration in blood and urine. To
isolate a hormone, the biochemist fractionates extracts
containing the putative hormone, with the same tech-
niques used to purify other biomolecules (solvent frac-
tionation, chromatography, and electrophoresis), and then
assays each fraction for hormone activity. Once the chem-
ical has been purified, its composition and structure can
be determined.
This protocol for hormone characterization is de-
ceptively simple. Hormones are extremely potent and
are produced in very small amounts. Obtaining sufficient
hormone to allow its chemical characterization often in-
volves biochemical isolations on a heroic scale. When
Roger Guillemin and Andrew Schally independently
purified and characterized thyrotropin-releasing hor-
mone (TRH) from the hypothalamus, Schally’s group
processed about 20 tons of hypothalamus from nearly
two million sheep, and Guillemin’s group extracted the
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism882
(a) Neuronal
signaling
(b) Endocrine signaling
Target
cells
Nerve
impulse
Contraction
Bloodstream
Metabolic
change
SecretionNerve
impulse
FIGURE 23–1 Signaling by the neuroendocrine system. (a) In neu-
ronal signaling, electrical signals (nerve impulses) originate in the cell
body of a neuron and travel very rapidly over long distances to the
axon tip, where neurotransmitters are released and diffuse to the tar-
get cell. The target cell (another neuron, a myocyte, or a secretory cell)
is only a fraction of a micrometer or a few micrometers away from
the site of neurotransmitter release. (b) In the endocrine system, hor-
mones are secreted into the bloodstream, which carries them through-
out the body to target tissues that may be a meter or more away from
the secreting cell. Both neurotransmitters and hormones interact with
specific receptors on or in their target cells, triggering responses.
8885d_c23_881-919 3/1/04 1:26 PM Page 882 mac76 mac76:385_reb:
23.1 Hormones: Diverse Structures for Diverse Functions 883
BOX 23–1 BIOCHEMISTRY IN MEDICINE
How Is a Hormone Discovered?
The Arduous Path to Purified Insulin
Millions of people with type I (insulin-dependent) di-
abetes mellitus inject themselves daily with pure in-
sulin to compensate for the lack of production of this
critical hormone by their own pancreatic H9252 cells. Insulin
injection is not a cure for diabetes, but it allows peo-
ple who otherwise would have died young to lead long
and productive lives. The discovery of insulin, which
began with an accidental observation, illustrates the
combination of serendipity and careful experimenta-
tion that led to the discovery of many of the hormones.
In 1889, Oskar Minkowski, a young assistant at the
Medical College of Strasbourg, and Josef von Mering,
at the Hoppe-Seyler Institute in Strasbourg, had a
friendly disagreement about whether the pancreas,
known to contain lipases, was important in fat diges-
tion in dogs. To resolve the issue, they began an ex-
periment on fat digestion. They surgically removed the
pancreas from a dog, but before their experiment got
any farther, Minkowski noticed that the dog was now
producing far more urine than normal (a common
symptom of untreated diabetes). Also, the dog’s urine
had glucose levels far above normal (another symp-
tom of diabetes). These findings suggested that lack
of some pancreatic product caused diabetes.
Minkowski tried unsuccessfully to prepare an ex-
tract of dog pancreas that would reverse the effect of
removing the pancreas—that is, would lower the uri-
nary or blood glucose levels. We now know that insulin
is a protein, and that the pancreas is very rich in pro-
teases (trypsin and chymotrypsin), normally released
directly into the small intestine to aid in digestion.
These proteases doubtless degraded the insulin in the
pancreatic extracts in Minkowski’s experiments.
Despite considerable effort, no significant prog-
ress was made in the isolation or characterization of
the “antidiabetic factor” until the summer of 1921,
when Frederick G. Banting, a young scientist working
in the laboratory of J. J. R. MacLeod at the University
of Toronto, and a student assistant, Charles Best, took
up the problem. By that time, several lines of evidence
pointed to a group of specialized cells in the pancreas
(the islets of Langerhans; see Fig. 23–24) as the source
of the antidiabetic factor, which came to be called in-
sulin (from Latin insula, “island”).
Taking precautions to prevent proteolysis, Bant-
ing and Best (later aided by biochemist J. B. Collip)
succeeded in December 1921 in preparing a purified
pancreatic extract that cured the symptoms of ex-
perimental diabetes in dogs. On January 25, 1922 (just
one month later!), their insulin preparation was in-
jected into Leonard Thompson, a 14-year-old boy se-
verely ill with diabetes mellitus. Within days, the lev-
els of ketone bodies and glucose in Thompson’s urine
dropped dramatically; the extract saved his life. In
1923, Banting and MacLeod won the Nobel Prize for
their isolation of insulin. Banting immediately an-
nounced that he would share his prize with Best;
MacLeod shared his with Collip.
By 1923, pharmaceutical companies were supply-
ing thousands of patients throughout the world with in-
sulin extracted from porcine pancreas. With the devel-
opment of genetic engineering techniques in the 1980s
(Chapter 9), it became possible to produce unlimited
quantities of human insulin by inserting the cloned hu-
man gene for insulin in a microorganism, which was
then cultured on an industrial scale. Some patients with
diabetes are now fitted with implanted insulin pumps,
which release adjustable amounts of insulin on demand
to meet changing needs at meal times and during ex-
ercise. There is a reasonable prospect that, in the fu-
ture, transplantation of pancreatic tissue will provide
diabetic patients with a source of insulin that responds
as well as normal pancreas, releasing insulin into the
bloodstream only when blood glucose rises.
Frederick G. Banting,
1891–1941
J. J. R. MacLeod,
1876–1935
Charles Best,
1899–1978
J. B. Collip,
1892–1965
8885d_c23_881-919 3/1/04 1:26 PM Page 883 mac76 mac76:385_reb:
hypothalamus from about a million pigs! TRH proved
to be a simple derivative of the tripeptide Glu–His–Pro
(Fig. 23–2). Once the structure of the hormone was
known, it could be chemically synthesized in large quan-
tities for use in physiological and biochemical studies.
For their work on hypothalamic hormones, Schally
and Guillemin shared the Nobel Prize in Physiology or
Medicine in 1977, along with Rosalyn Yalow, who (with
Solomon A. Berson) developed the extraordinarily sen-
sitive radioimmunoassay (RIA) for peptide hormones
and used it to study hormone action. RIA revolutionized
hormone research by making possible the rapid, quan-
titative, and specific measurement of hormones in
minute amounts.
Hormone-specific antibodies are the key to the ra-
dioimmunoassay. Purified hormone, injected into rab-
bits, elicits antibodies that bind to that hormone with
very high affinity and specificity. When a constant
amount of isolated antibody is incubated with a fixed
amount of the radioactively labeled hormone, a certain
fraction of the radioactive hormone binds to the anti-
body (Fig. 23–3). If, in addition to the radiolabeled hor-
mone, unlabeled hormone is also present, the unlabeled
hormone competes with and displaces some of the la-
beled hormone from its binding site on the antibody. This
binding competition can be quantified by reference to a
standard curve obtained with known amounts of unla-
beled hormone. The degree to which labeled hormone
is displaced from antibody is a measure of the amount
of unlabeled hormone in a sample of blood or tissue
extract. By using very highly radioactive hormone, re-
searchers can make the assay sensitive to picograms of
hormone. A newer variation of this technique, enzyme-
linked immunosorbent assay (ELISA), is illustrated in
Figure 5–28b.
Hormones Act through Specific High-Affinity
Cellular Receptors
As we saw in Chapter 12, all hormones act through
highly specific receptors in hormone-sensitive target
cells, to which the hormones bind with high affinity (see
Fig. 12–2). Each cell type has its own combination of
hormone receptors, which define the range of its hor-
mone responsiveness. Moreover, two cell types with the
same type of receptor may have different intracellular
targets of hormone action and thus may respond dif-
ferently to the same hormone. The specificity of hor-
mone action results from structural complementarity
between the hormone and its receptor; this interaction
is extremely selective, so structurally similar hormones
can have different effects. The high affinity of the in-
teraction allows cells to respond to very low concentra-
tions of hormone. In the design of drugs intended to
intervene in hormonal regulation, we need to know the
relative specificity and affinity of the drug and the nat-
ural hormone. Recall that hormone-receptor interac-
tions can be quantified by Scatchard analysis (see Box
12–1), which, under favorable conditions, yields a quan-
titative measure of affinity (the dissociation constant for
the complex) and the number of hormone-binding sites
in a preparation of receptor.
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism884
FIGURE 23–2 The structure of thyrotropin-releasing hormone (TRH).
Purified (by heroic efforts) from extracts of hypothalamus, TRH proved
to be a derivative of the tripeptide Glu–His–Pro. The side-chain car-
boxyl group of the amino-terminal Glu forms an amide (red bond)
with the residue’s H9251-amino group, creating pyroglutamate, and the car-
boxyl group of the carboxyl-terminal Pro is converted to an amide
(red ONH
2
). Such modifications are common among the small pep-
tide hormones. In a typical protein of M
r
~50, the charges on the
amino- and carboxyl-terminal groups contribute relatively little to the
overall charge on the protein, but in a tripeptide these two charges
dominate the properties of the peptide. Formation of the amide de-
rivatives removes these charges.
CO
NH
CH
2
CH C
O
NH CH
C
CH
2
Histidine ProlylamidePyroglutamate
CH C
CH
2
CH
2
HC
pyroGlu–His–Pro–NH
2
N
NH
CH
CH
2
CH
2
NH
2
C
O
O
N
Roger Guillemin Rosalyn S. Yalow Andrew V. Schally
8885d_c23_881-919 3/1/04 1:26 PM Page 884 mac76 mac76:385_reb:
The locus of the encounter between hormone and
receptor may be extracellular, cytosolic, or nuclear, de-
pending on the hormone type. The intracellular conse-
quences of hormone-receptor interaction are of at least
six general types: (1) a change in membrane potential
results from the opening or closing of a hormone-gated
23.1 Hormones: Diverse Structures for Diverse Functions 885
ion channel; (2) a receptor enzyme is activated by the
extracellular hormone; (3) a second messenger (such
as cAMP or inositol trisphosphate) generated inside the
cell acts as an allosteric regulator of one or more en-
zymes; (4) a receptor with no intrinsic enzyme activity
recruits a soluble protein kinase in the cytosol, which
passes on the signal; (5) an adhesion receptor on the
cell surface interacts with molecules in the extracellu-
lar matrix and conveys information to the cytoskeleton;
or (6) a steroid or steroidlike molecule causes a change
in the level of expression (transcription of DNA into
mRNA) of one or more genes, mediated by a nuclear
hormone receptor protein (see Fig. 12–2).
Water-soluble peptide and amine hormones (insulin
and epinephrine, for example) act extracellularly by
binding to cell surface receptors that span the plasma
membrane (Fig. 23–4). When the hormone binds to its
extracellular domain, the receptor undergoes a confor-
mational change analogous to that produced in an al-
losteric enzyme by binding of an effector molecule. The
conformational change triggers the downstream effects
of the hormone.
A single hormone molecule, in forming a hormone-
receptor complex, activates a catalyst that produces
many molecules of second messenger, so the receptor
Antibody
Radiolabeled
hormone
1
Radiolabeled and
unlabeled hormone
2
Radiolabeled ACTH
[unbound]
Unlabeled ACTH added (pg)
()
0
1000100
Unknown sample
101
0.2
0.4
0.6
0.8
1.0
1.2
ratio
[bound]
Standard curve
FIGURE 23–3 Radioimmunoassay (RIA). (a) A low concentration of
radiolabeled hormone (red) is incubated with 1 a fixed amount of
antibody specific for that hormone or 2 a fixed amount of antibody
and various concentrations of unlabeled hormone (blue). In the latter
case, unlabeled hormone competes with labeled hormone for bind-
ing to the antibody; the amount of labeled hormone bound varies in-
versely with the concentration of unlabeled hormone present. (b) A
radioimmunoassay for adrenocorticotropic hormone (ACTH). A stan-
dard curve of the ratio [bound] to [unbound radiolabeled ACTH] vs.
[unlabeled ACTH added] is constructed and used to determine the
amount of (unlabeled) ACTH in an unknown sample. If an aliquot con-
taining an unknown quantity of unlabeled hormone gives, say, a value
of 0.4 for the ratio [bound]/[unbound] (see arrow), the aliquot must
contain about 20 pg of ACTH.
Rec
Rec
Peptide or amine
hormone binds to
receptor on the
outside of the cell;
acts through receptor
without entering
the cell.
Steroid or thyroid hormone
enters the cell; hormone-
receptor complex acts in
the nucleus.
Second messenger
(e.g., cAMP)
Plasma
membrane
Nucleus
Altered transcription
of specific genes
Altered activity of
preexisting
enzyme
Altered amount of
newly synthesized
proteins
FIGURE 23–4 Two general mechanisms of hormone action. The pep-
tide and amine hormones are faster acting than steroid and thyroid
hormones.
(a)
(b)
8885d_c23_885 3/2/04 7:59 AM Page 885 mac76 mac76:385_reb:
serves not only as a signal transducer but also as a sig-
nal amplifier. The signal may be further amplified by a
signaling cascade, a series of steps in which a catalyst
activates a catalyst, resulting in very large amplifications
of the original signal. A cascade of this type occurs in
the regulation of glycogen synthesis and breakdown by
epinephrine (see Fig. 12–16). Epinephrine activates
(through its receptor) adenylyl cyclase, which produces
many molecules of cAMP for each molecule of receptor-
bound hormone. Cyclic AMP in turn activates cAMP-
dependent protein kinase, which activates phosphory-
lase kinase, which activates glycogen phosphorylase.
The result is signal amplification: one epinephrine mol-
ecule causes the production of many thousands of
molecules of glucose 1-phosphate from glycogen.
Water-insoluble hormones (steroid, retinoid, and thy-
roid hormones) readily pass through the plasma mem-
brane of their target cells to reach their receptor proteins
in the nucleus (Fig. 23–4). With this class of hormones,
the hormone-receptor complex itself carries the message;
it interacts with DNA to alter the expression of specific
genes, changing the enzyme complement of the cell and
thereby changing cellular metabolism (see Fig. 12–40).
Hormones that act through plasma membrane re-
ceptors generally trigger very rapid physiological or bio-
chemical responses. Just seconds after the adrenal
medulla secretes epinephrine into the bloodstream,
skeletal muscle responds by accelerating the breakdown
of glycogen. By contrast, the thyroid hormones and the
sex (steroid) hormones promote maximal responses in
their target tissues only after hours or even days. These
differences in response time correspond to different
modes of action. In general, the fast-acting hormones
lead to a change in the activity of one or more preex-
isting enzymes in the cell, by allosteric mechanisms or
covalent modification. The slower-acting hormones gen-
erally alter gene expression, resulting in the synthesis
of more or less of the regulated protein(s).
Hormones Are Chemically Diverse
Mammals have several classes of hormones, distin-
guishable by their chemical structures and their modes
of action (Table 23–1). Peptide, amine, and eicosanoid
hormones act from outside the target cell via surface re-
ceptors. Steroid, vitamin D, retinoid, and thyroid hor-
mones enter the cell and act through nuclear receptors.
Nitric oxide also enters the cell, but activates a cytoso-
lic enzyme, guanylyl cyclase (see Fig. 12–10).
Hormones can also be classified by the way they get
from the point of their release to their target tissue. En-
docrine (from the Greek endon, “within,” and krinein,
“to release”) hormones are released into the blood and
carried to target cells throughout the body (insulin is
an example). Paracrine hormones are released into the
extracellular space and diffuse to neighboring target
cells (the eicosanoid hormones are of this type). Au-
tocrine hormones are released by and affect the same
cell, binding to receptors on the cell surface.
Mammals are hardly unique in possessing hormonal
signaling systems. Insects and nematode worms have
highly developed systems for hormonal regulation, with
fundamental mechanisms similar to those in mammals.
Plants, too, use hormonal signals to coordinate the ac-
tivities of their various tissues (Chapter 12). The study
of hormone action is not as advanced in plants as in ani-
mals, but we do know that some mechanisms are shared.
To illustrate the structural diversity and range of action
of mammalian hormones, we consider representative
examples of each major class listed in Table 23–1.
Peptide Hormones Peptide hormones may have from 3
to 200 or more amino acid residues. They include the
pancreatic hormones insulin, glucagon, and somato-
statin, the parathyroid hormone, calcitonin, and all the
hormones of the hypothalamus and pituitary (described
below). These hormones are synthesized on ribosomes
in the form of longer precursor proteins (prohormones),
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism886
TABLE 23–1 Classes of Hormones
Type Example Synthetic path Mode of action
Peptide Insulin, glucagon Proteolytic processing of
prohormone
Catecholamine Epinephrine From tyrosine Plasma membrane receptors; second messengers
Eicosanoid PGE
1
From arachidonate
(20:4 fatty acid)
Steroid Testosterone From cholesterol
Vitamin D 1,25-Dihydroxycholecalciferol From cholesterol
Retinoid Retinoic acid From vitamin A
Thyroid Triiodothyronine (T
3
) From Tyr in thyroglobulin
Nitric oxide Nitric oxide From arginine H11001 O
2
Cytosolic receptor (guanylate cyclase) and
second messenger (cGMP)
Nuclear receptors; transcriptional regulation
?
?
?
?
?
?
?
?
?
?
?
?
8885d_c23_881-919 3/1/04 1:26 PM Page 886 mac76 mac76:385_reb:
then packaged into secretory vesicles and proteolyti-
cally cleaved to form the active peptides. Insulin is a
small protein (M
r
5,800) with two polypeptide chains,
A and B, joined by two disulfide bonds. It is synthesized
in the pancreas as an inactive single-chain precursor,
preproinsulin (Fig. 23–5), with an amino-terminal “sig-
nal sequence” that directs its passage into secretory
vesicles. (Signal sequences are discussed in Chapter 27;
see Fig. 27–33.) Proteolytic removal of the signal se-
quence and formation of three disulfide bonds produces
proinsulin, which is stored in secretory granules in pan-
creatic H9252 cells. When elevated blood glucose triggers in-
sulin secretion, proinsulin is converted to active insulin
by specific proteases, which cleave two peptide bonds
to form the mature insulin molecule.
In some cases, prohormone proteins yield a single
peptide hormone, but often several active hormones
are carved out of the same prohormone. Pro-opiomelano-
cortin (POMC) is a spectacular example of multiple
hormones encoded by a single gene. The POMC gene
encodes a large polypeptide that is progressively carved
up into at least nine biologically active peptides (Fig.
23–6). The terminal residues of peptide hormones are
often modified, as in TRH (Fig. 23–2).
23.1 Hormones: Diverse Structures for Diverse Functions 887
C
COO
H11002
C peptideSignal sequence
Signal
sequence
A B B chainA chain
Preproinsulin Proinsulin Mature
insulin
COO
H11002
S
S
COO
H11002
NH
3
H11001
NH
3
H11001
NH
3
H11001
H
3
N
H11001
SS
SS
S
S SS
SS
H11002
OOC
–
FIGURE 23–5 Insulin. Mature insulin is formed
from its larger precursor preproinsulin by proteolytic
processing. Removal of a 23 amino acid segment
(the signal sequence) at the amino terminus of
preproinsulin and formation of three disulfide bonds
produces proinsulin. Further proteolytic cuts remove
the C peptide from proinsulin to produce mature
insulin, composed of A and B chains. The amino
acid sequence of bovine insulin is shown in Figure
3–24.
Pro-opiomelanocortin (POMC) gene
Signal
peptide
5H11032 3H11032 mRNA
DNA
-Lipotropin
-MSH Met-enkephalin
CLIP-MSH -Lipotropin
COO
H11002
H
3
N
ACTH-MSH
H11001
H9253
H9251 H9253
H9252
H9252
-EndorphinH9252
FIGURE 23–6 Proteolytic processing of the pro-
opiomelanocortin (POMC) precursor. The initial gene
product of the POMC gene is a long polypeptide that
undergoes cleavage by a series of specific proteases to
produce ACTH, H9252- and H9253-lipotropin, H9251-, H9252-, and H9253-MSH
(melanocyte-stimulating hormone), CLIP (corticotropin-
like intermediary peptide), H9252-endorphin, and
Met-enkephalin. The points of cleavage are paired
basic residues, Arg–Lys, Lys–Arg, or Lys–Lys.
8885d_c23_881-919 3/1/04 1:26 PM Page 887 mac76 mac76:385_reb:
The concentration of peptide hormones within se-
cretory granules is so high that the vesicle contents are
virtually crystalline; when the contents are released by
exocytosis, a large amount of hormone is released sud-
denly. The capillaries that serve peptide-producing en-
docrine glands are fenestrated (and thus permeable to
peptides), so the hormone molecules readily enter the
bloodstream for transport to target cells elsewhere. As
noted earlier, all peptide hormones act by binding to re-
ceptors in the plasma membrane. They cause the gen-
eration of a second messenger in the cytosol, which
changes the activity of an intracellular enzyme, thereby
altering the cell’s metabolism.
Catecholamine Hormones The water-soluble compounds
epinephrine (adrenaline) and norepinephrine (nor-
adrenaline) are catecholamines, named for the struc-
turally related compound catechol. They are synthesized
from tyrosine.
Catecholamines produced in the brain and in other
neural tissues function as neurotransmitters, but
epinephrine and norepinephrine are also hormones, syn-
thesized and secreted by the adrenal glands. Like the
peptide hormones, catecholamines are highly concen-
trated within secretory vesicles and released by exocy-
tosis, and they act through surface receptors to gener-
ate intracellular second messengers. They mediate a
wide variety of physiological responses to acute stress
(see Table 23–6).
Eicosanoids The eicosanoid hormones (prostaglandins,
thromboxanes, and leukotrienes) are derived from the
20-carbon polyunsaturated fatty acid arachidonate.
Unlike the hormones described above, they are not
synthesized in advance and stored; they are produced,
when needed, from arachidonate enzymatically released
from membrane phospholipids by phospholipase A
2
(see
Fig. 10–18). The enzymes of the pathway leading to
prostaglandins and thromboxanes (see Fig. 21–15) are
very widely distributed in mammalian tissues; most cells
Phospholipids
Arachidonate
(20:4)
LeukotrienesProstaglandins Thromboxanes
Tyrosine L-DOPA Dopamine
Norepinephrine Epinephrine
can produce these signals, and cells of many tissues can
respond to them through specific plasma membrane re-
ceptors. The eicosanoid hormones are paracrine hor-
mones, secreted into the interstitial fluid (not primarily
into the blood) and acting on nearby cells.
Prostaglandins promote the contraction of
smooth muscle, including that of the intestine
and uterus (and can therefore be used medically to in-
duce labor). They also mediate pain and inflammation
in all tissues. Many antiinflammatory drugs act by in-
hibiting steps in the prostaglandin synthetic pathway
(see Box 21–2). Thromboxanes regulate platelet func-
tion and therefore blood clotting. Leukotrienes LTC
4
and LTD
4
act through plasma membrane receptors to
stimulate contraction of smooth muscle in the intestine,
pulmonary airways, and trachea. They are mediators of
the severe immune response called anaphylaxis. ■
Steroid Hormones The steroid hormones (adrenocorti-
cal hormones and sex hormones) are synthesized from
cholesterol in several endocrine tissues.
They travel to their target cells through the blood-
stream, bound to carrier proteins. More than 50 corti-
costeroid hormones are produced in the adrenal cortex
by reactions that remove the side chain from the D ring
of cholesterol and introduce oxygen to form keto and
hydroxyl groups. Many of these reactions involve cy-
tochrome P-450 enzymes (see Box 21–1). The steroid
hormones are of two general types. Glucocorticoids
(such as cortisol) primarily affect the metabolism of car-
bohydrates; mineralocorticoids (such as aldosterone)
regulate the concentrations of electrolytes in the blood.
Androgens (testosterone) and estrogens (such as estra-
diol; see Fig. 10–19) are synthesized in the testes and
ovaries. Their synthesis also involves cytochrome P-450
enzymes that cleave the side chain of cholesterol and
introduce oxygen atoms. These hormones affect sexual
development, sexual behavior, and a variety of other re-
productive and nonreproductive functions.
All steroid hormones act through nuclear recep-
tors to change the level of expression of specific genes
(p. 465). Recent evidence indicates that they also have
more rapid effects, mediated by receptors localized in
the plasma membrane.
Cholesterol
Progesterone
Cortisol
(glucocorticoid)
Aldosterone
(mineralocorticoid)
Testosterone
Estradiol
(sex hormones)
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism888
8885d_c23_881-919 3/1/04 1:26 PM Page 888 mac76 mac76:385_reb:
Vitamin D Hormone Calcitriol (1,25-dihydroxycholecal-
ciferol) is produced from vitamin D by enzyme-
catalyzed hydroxylation in the liver and kidneys (see
Fig. 10–20a).
Vitamin D is obtained in the diet or by photolysis of 7-
dehydrocholesterol in skin exposed to sunlight. Cal-
citriol works in concert with parathyroid hormone in
Ca
2H11001
homeostasis, regulating [Ca
2H11001
] in the blood and
the balance between Ca
2H11001
deposition and Ca
2H11001
mobi-
lization from bone. Acting through nuclear receptors,
calcitriol activates the synthesis of an intestinal Ca
2H11001
-
binding protein essential for uptake of dietary Ca
2H11001
. In-
adequate dietary vitamin D or defects in the biosynthe-
sis of calcitriol result in serious diseases such as rickets,
in which bones are weak and malformed (see Fig.
10–20b).
Retinoid Hormones Retinoids are potent hormones that
regulate the growth, survival, and differentiation of cells
via nuclear retinoid receptors. The prohormone retinol
is synthesized from vitamin A, primarily in liver (see Fig.
10–21), and many tissues convert retinol to the hormone
retinoic acid (RA).
All tissues are retinoid targets, as all cell types have at
least one form of nuclear retinoid receptor. In adults, the
most significant targets include cornea, skin, epithelia of
the lungs and trachea, and the immune system. RA reg-
ulates the synthesis of proteins essential for growth or
differentiation. Excessive vitamin A can cause birth de-
fects, and pregnant women are advised not to use the
retinoid creams that have been developed for treatment
of severe acne.
Thyroid Hormones The thyroid hormones T
4
(thyroxine)
and T
3
(triiodothyronine) are synthesized from the pre-
cursor protein thyroglobulin (M
r
660,000). Up to 20 Tyr
residues in thyroglobulin are enzymatically iodinated
b-Carotene
Vitamin A
1
(retinol)
Retinoic acid
7-Dehydrocholesterol
UV light
Vitamin D
3
(cholecalciferol)
25-Hydroxycholecalciferol
1,25-Dihydroxycholecalciferol
in the thyroid gland, then two iodotyrosine residues con-
dense to form the precursor to thyroxine. When needed,
thyroxine is released by proteolysis. Condensation of
monoiodotyrosine with diiodotyrosine produces T
3
,
which is also an active hormone released by proteolysis.
The thyroid hormones act through nuclear receptors to
stimulate energy-yielding metabolism, especially in liver
and muscle, by increasing the expression of genes en-
coding key catabolic enzymes.
Nitric Oxide (NO) Nitric oxide is a relatively stable free
radical synthesized from molecular oxygen and the
guanidino nitrogen of arginine (see Fig. 22–31) in a re-
action catalyzed by NO synthase.
Arginine H11001 1
1
?
2
NADPH H11001 2O
2
?→
NO H11001 citrulline H11001 2H
2
O H11001 1
1
?
2
NADP
H11001
This enzyme is found in many tissues and cell types:
neurons, macrophages, hepatocytes, myocytes of
smooth muscle, endothelial cells of the blood vessels,
and epithelial cells of the kidney. NO acts near its point
of release, entering the target cell and activating the cy-
tosolic enzyme guanylyl cyclase, which catalyzes the for-
mation of the second messenger cGMP (see Fig. 12–10).
Hormone Release Is Regulated by a Hierarchy
of Neuronal and Hormonal Signals
The changing levels of specific hormones regulate spe-
cific cellular processes, but what regulates the level of
each hormone? The brief answer is that the central nerv-
ous system receives input from many internal and ex-
ternal sensors—signals about danger, hunger, dietary in-
take, blood composition and pressure, for example—and
orchestrates the production of appropriate hormonal
signals by the endocrine tissues. For a more complete
answer, we must look at the hormone-producing sys-
tems of the human body and some of their functional
interrelationships.
Figure 23–7 shows the anatomic location of the ma-
jor endocrine glands in humans, and Figure 23–8 rep-
resents the “chain of command” in the hormonal sig-
naling hierarchy. The hypothalamus, a small region of
the brain (Fig. 23–9), is the coordination center of the
Thyroglobulin–Tyr
proteolysis
Thyroglobulin–Tyr–I
(iodinated Tyr residues)
Thyroxine (T
4
),
triiodothyronine (T
3
)
23.1 Hormones: Diverse Structures for Diverse Functions 889
8885d_c23_889 3/2/04 7:59 AM Page 889 mac76 mac76:385_reb:
endocrine system; it receives and integrates messages
from the central nervous system. In response to these
messages, the hypothalamus produces regulatory hor-
mones (releasing factors) that pass directly to the
nearby pituitary gland, through special blood vessels
and neurons that connect the two glands (Fig. 23–9b).
The pituitary gland has two functionally distinct parts.
The posterior pituitary contains the axonal endings
of many neurons that originate in the hypothalamus.
These neurons produce the short peptide hormones
oxytocin and vasopressin (Fig. 23–10), which then move
down the axon to the nerve endings in the pituitary,
where they are stored in secretory granules to await the
signal for their release.
The anterior pituitary responds to hypothalamic
hormones carried in the blood, producing tropic hor-
mones, or tropins (from the Greek tropos, “turn”).
These relatively long polypeptides activate the next
rank of endocrine glands (Fig. 23–8), which includes the
adrenal cortex, thyroid gland, ovaries, and testes. These
glands in turn secrete their specific hormones, which
are carried in the bloodstream to the receptors of cells in
the target tissues. For example, corticotropin-releasing
hormone from the hypothalamus stimulates the anterior
pituitary to release ACTH, which travels to the zona fas-
ciculata of the adrenal cortex and triggers the release of
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism890
Hypothalamus
Pituitary
Thyroid
Adrenals
Pancreas
Kidneys
Ovaries
(female)
Testes
(male)
Parathyroids
(behind the
thyroid)
Adipose tissue
FIGURE 23–7 The major endocrine glands. The glands are shaded
dark pink.
Sensory input from environment
Central nervous system
Hypothalamus
Anterior pituitary
Neuroendocrine
origins of
signals
First targets
Ultimate targets
Many
tissues
Reproductive organs
Thyroxine
(T
4
), triiodo-
thyronine (T
3
)
Progesterone,
estradiol Testosterone
ThyroidSecond targets
Adrenal
cortex
Ovaries/testes
Thyrotropin
M
r
28,000
Corticotropin
(ACTH)
M
r
4,500
Follicle-
stimulating
hormone
M
r
24,000
Luteinizing
hormone
M
r
20,500
Muscles,
liver
Cortisol,
corticosterone,
aldosterone
Liver,
bone
Mammary
glands
Smooth
muscle,
mammary
glands
Arterioles,
kidney
Liver,
muscles
Liver,
muscles,
heart
Somatotropin
(growth hormone)
M
r
21,500
Prolactin
M
r
22,000
Oxytocin
M
r
1,007
Blood
glucose
level
Posterior pituitary
Insulin,
glucagon,
somatostatin
Epinephrine
Adrenal
medulla
Islet cells of
pancreas
Vasopressin
(antidiuretic
hormone)
M
r
1,040
Hypothalamic hormones
(releasing factors)
FIGURE 23–8 The major endocrine systems and their target
tissues. Signals originating in the central nervous system (top)
pass via a series of relays to the ultimate target tissues (bottom).
In addition to the systems shown, the thymus, pineal gland,
and groups of cells in the gastrointestinal tract also secrete
hormones. Dashed lines represent neuronal connections.
8885d_c23_881-919 3/1/04 1:26 PM Page 890 mac76 mac76:385_reb:
cortisol. Cortisol, the ultimate hormone in this cascade,
acts through its receptor in many types of target cells
to alter their metabolism. In hepatocytes, one effect of
cortisol is to increase the rate of gluconeogenesis.
Hormonal cascades such as those responsible for the
release of cortisol and epinephrine result in large ampli-
fications of the initial signal and allow exquisite fine-
tuning of the output of the ultimate hormone (Fig.
23–11). At each level in the cascade, a small signal
elicits a larger response. The initial electrical signal
to the hypothalamus results in the release of a few
nanograms of corticotropin-releasing hormone, which
elicits the release of a few micrograms of corticotropin.
Corticotropin acts on the adrenal cortex to cause the re-
lease of milligrams of cortisol, for an overall amplifica-
tion of at least a millionfold.
At each level of a hormonal cascade, feedback in-
hibition of earlier steps in the cascade is possible; an
unnecessarily elevated level of the ultimate hormone
or of one of the intermediate hormones inhibits the re-
lease of earlier hormones in the cascade. These feed-
back mechanisms accomplish the same end as those that
limit the output of a biosynthetic pathway (compare
Fig. 23–11 with Fig. 6–28): a product is synthesized
(or released) only until the necessary concentration is
reached.
23.1 Hormones: Diverse Structures for Diverse Functions 891
FIGURE 23–10 Two hormones of the posterior pituitary gland. The
carboxyl-terminal residues are glycinamide, ONHOCH
2
OCONH
2
(as noted in Fig. 23–2, amidation of the carboxyl terminus is common
in short peptide hormones). These two hormones, identical in all but
two residues (shaded), have very different biological effects. Oxytocin
acts on the smooth muscles of the uterus and mammary gland, caus-
ing uterine contractions during labor and promoting milk release dur-
ing lactation. Vasopressin (also called antidiuretic hormone) increases
water reabsorption in the kidney and promotes the constriction of
blood vessels, thereby increasing blood pressure.
(a)
(b)
Posterior
pituitary
Hypothalamus
Arteries
Release of posterior
pituitary hormones
(vasopressin, oxytocin)
Release of
hypothalamic
factors into
arterial blood
Capillary
network
Release of
anterior
pituitary
hormones
(tropins)
Nerve
axons
Hypothalamus
Anterior
pituitary
Afferent nerve signals to
hypothalamus
Veins carry hormones
to systemic blood
Anterior
pituitary
Posterior
pituitary
Cys
Tyr
Ile
Gln
Asn
Pro
Leu
Gly
CO
Cys
S
S
NH
2
Cys
Tyr
Phe
Gln
Asn
Pro
Arg
Gly
CO
NH
3
Cys
S
S
H11001
Human vasopressin
(antidiuretic hormone)
Human oxytocin
NH
2
NH
3
H11001
FIGURE 23–9 Neuroendocrine origins of hormone signals. (a) Location of
the hypothalamus and pituitary gland. (b) Details of the hypothalamus-
pituitary system. Signals from connecting neurons stimulate the hypothalamus
to secrete releasing factors into a blood vessel that carries the hormones
directly to a capillary network in the anterior pituitary. In response to each
hypothalamic releasing factor, the anterior pituitary releases the appropriate
hormone into the general circulation. Posterior pituitary hormones are synthe-
sized in neurons arising in the hypothalamus, transported along axons to
nerve endings in the posterior pituitary, and stored there until released into
the blood in response to a neuronal signal.
(a)
8885d_c23_891 3/2/04 8:00 AM Page 891 mac76 mac76:385_reb:
■ Hormonal cascades, in which catalysts activate
catalysts, amplify the initial stimulus by several
orders of magnitude, often in a very short time
(seconds).
■ Nerve impulses stimulate the hypothalamus to
send specific hormones to the pituitary gland,
thus stimulating (or inhibiting) the release of
tropic hormones. The anterior pituitary
hormones in turn stimulate other endocrine
glands (thyroid, adrenals, pancreas) to secrete
their characteristic hormones, which in turn
stimulate specific target tissues.
■ Peptide, amine, and eicosanoid hormones act
outside the target cell on specific receptors in
the plasma membrane, altering the level of an
intracellular second messenger.
■ Steroid, vitamin D, retinoid, and thyroid
hormones enter target cells and alter gene
expression by interacting with specific nuclear
receptors.
23.2 Tissue-Specific Metabolism:
The Division of Labor
Each tissue of the human body has a specialized func-
tion, reflected in its anatomy and metabolic activity (Fig.
23–12). Skeletal muscle allows directed motion; adipose
tissue stores and releases energy in the form of fats,
which serve as fuel throughout the body; the brain
pumps ions across plasma membranes to produce elec-
trical signals. The liver plays a central processing and
distributing role in metabolism and furnishes all other
organs and tissues with an appropriate mix of nutrients
via the bloodstream. The functional centrality of the
liver is indicated by the common reference to all other
tissues and organs as “extrahepatic” or “peripheral.” We
therefore begin our discussion of the division of meta-
bolic labor by considering the transformations of carbo-
hydrates, amino acids, and fats in the mammalian liver.
This is followed by brief descriptions of the primary
metabolic functions of adipose tissue, muscle, brain, and
the medium that interconnects all others: the blood.
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism892
Cortisol
(mg)
Adrenocorticotropic
hormone (ACTH)
(H9262g)
Central
nervous
system
Hypothalamus
Anterior pituitary
Adrenal gland
Corticotropin-releasing
hormone (CRH)
(ng)
Infection
HemorrhageFear
Pain Hypoglycemia
Liver AdiposeMuscle
FIGURE 23–11 Cascade of hormone release following central nerv-
ous system input to the hypothalamus. In each endocrine tissue along
the pathway, a stimulus from the level above is received, amplified,
and transduced into the release of the next hormone in the cascade.
The cascade is sensitive to regulation at several levels through feed-
back inhibition by the ultimate hormone. The product therefore regu-
lates its own production, as in feedback inhibition of biosynthetic path-
ways within a single cell.
SUMMARY 23.1 Hormones: Diverse Structures
for Diverse Functions
■ Hormones are chemical messengers secreted
by certain tissues into the blood or interstitial
fluid, serving to regulate the activity of other
cells or tissues.
■ Radioimmunoassay (RIA) and ELISA are two
very sensitive techniques for detecting and
quantifying hormones.
8885d_c23_881-919 3/1/04 1:26 PM Page 892 mac76 mac76:385_reb:
The Liver Processes and Distributes Nutrients
During digestion in mammals, the three main classes of
nutrients (carbohydrates, proteins, and fats) undergo
enzymatic hydrolysis into their simple constituents. This
breakdown is necessary because the epithelial cells lin-
ing the intestinal lumen absorb only relatively small mol-
ecules. Many of the fatty acids and monoacylglycerols
released by digestion of fats in the intestine are re-
assembled within these epithelial cells into triacylglyc-
erols (TAGs).
After being absorbed, most sugars and amino acids
and some TAGs travel in the bloodstream to the liver;
the remaining TAGs enter adipose tissue via the lym-
phatic system. The portal vein is a direct route from the
digestive organs to the liver, and liver therefore has first
access to ingested nutrients. The liver has two main cell
types. Kupffer cells are phagocytes, important in im-
mune function. Hepatocytes, of primary interest here,
transform dietary nutrients into the fuels and precur-
sors required by other tissues and export them via the
blood. The kinds and amounts of nutrients supplied to
the liver vary with several factors, including the diet and
the time between meals. The demand of extrahepatic
tissues for fuels and precursors varies among organs and
with the level of activity and overall nutritional state of
the individual.
To meet these changing circumstances, the liver has
remarkable metabolic flexibility. For example, when the
diet is rich in protein, hepatocytes supply themselves
with high levels of enzymes for amino acid catabolism
and gluconeogenesis. Within hours after a shift to a high-
carbohydrate diet, the levels of these enzymes begin to
drop and the hepatocytes increase their synthesis of en-
zymes essential to carbohydrate metabolism and fat syn-
thesis. Liver enzymes turn over (are synthesized and
degraded) at five to ten times the rate of enzyme
turnover in other tissues, such as muscle. Extrahepatic
23.2 Tissue-Specific Metabolism: The Division of Labor 893
Secretes insulin
and glucagon in
response to changes
in blood glucose
concentration.
Pancreas
Brain
Transports ions to
maintain membrane
potential; integrates
inputs from body
and surroundings;
sends signals to
other organs.
Adipose
tissue
Synthesizes,
stores, and
mobilizes
triacylglycerols.
Uses ATP to do
mechanical work.
Carries lipids
from intestine to liver.
Skeletal muscle
Lymphatic
system
Small intestine
Absorbs nutrients
from the diet, moves
them into blood or
lymphatic system.
Liver
Processes fats,
carbohydrates,
proteins from diet;
synthesizes and
distributes lipids,
ketone bodies, and
glucose for other
tissues; converts
excess nitrogen
to urea.
Carries nutrients
from intestine to liver.
Portal vein
FIGURE 23–12 Specialized metabolic functions of mammalian tissues.
8885d_c23_881-919 3/1/04 1:26 PM Page 893 mac76 mac76:385_reb:
tissues also can adjust their metabolism to prevailing
conditions, but none is as adaptable as the liver, and
none is so central to the organism’s overall metabolism.
What follows is a survey of the possible fates of sugars,
amino acids, and lipids that enter the liver from the
bloodstream. To help you recall the metabolic transfor-
mations discussed here, Table 23–2 shows the major
pathways and processes to which we refer and indicates
by figure number where each pathway is presented in
detail. Here, we present summaries of the pathways, re-
ferring to the step numbers in Figures 23–13 to 23–15.
Sugars The glucose transporter in hepatocytes (GLUT2)
is so effective that the concentration of glucose within
a hepatocyte is essentially the same as that in the blood.
Glucose entering hepatocytes is phosphorylated by hex-
okinase IV (glucokinase) to yield glucose 6-phosphate.
Glucokinase has a much higher K
m
for glucose (10 mM)
than do the hexokinase isozymes in other cells (p. 578)
and, unlike these other isozymes, it is not inhibited by
its product, glucose 6-phosphate. The presence of glu-
cokinase allows hepatocytes to continue phosphory-
lating glucose when the glucose concentration rises
well above levels that would overwhelm other hexoki-
nases. The high K
m
of glucokinase also ensures that
the phosphorylation of glucose in hepatocytes is min-
imal when the glucose concentration is low, prevent-
ing the liver from consuming glucose as fuel via gly-
colysis. This spares glucose for other tissues. Fructose,
galactose, and mannose, all absorbed from the small
intestine, are also converted to glucose 6-phosphate by
enzymatic pathways examined in Chapter 14. Glucose
6-phosphate is at the crossroads of carbohydrate me-
tabolism in the liver. It may take any of several major
metabolic routes (Fig. 23–13), depending on the cur-
rent metabolic needs of the organism. By the action of
various allosterically regulated enzymes, and through
hormonal regulation of enzyme synthesis and activity,
the liver directs the flow of glucose into one or more
of these pathways.
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism894
TABLE 23–2 Pathways of Carbohydrate, Amino Acid, and Fat Metabolism Illustrated in Earlier Chapters
Pathway Figure reference
Citric acid cycle: acetyl-CoA ?→ 2CO
2
16–7
Oxidative phosphorylation: ATP synthesis 19–17
Carbohydrate catabolism
Glycogenolysis: glycogen ?→ glucose 1-phosphate ?→ blood glucose 15–3; 15–4
Hexose entry into glycolysis: fructose, mannose, galactose ?→ glucose 6-phosphate 14–9
Glycolysis: glucose ?→ pyruvate 14–2
Pyruvate dehydrogenase reaction: pyruvate ?→ acetyl-CoA 16–2
Lactic acid fermentation: glucose ?→ lactate H11001 2ATP 14–3
Pentose phosphate pathway: glucose 6-phosphate ?→ pentose phosphates + NADPH 14–21
Carbohydrate anabolism
Gluconeogenesis: citric acid cycle intermediates ?→ glucose 14–16
Glucose-alanine cycle: glucose ?→ pyruvate ?→ alanine ?→ glucose 18–9
Glycogen synthesis: glucose 6-phosphate ?→ glucose 1-phosphate ?→ glycogen 15–8
Amino acid and nucleotide metabolism
Amino acid degradation: amino acids ?→ acetyl-CoA, citric acid cycle intermediates 18–15
Amino acid synthesis 22–9
Urea cycle: NH
3
?→ urea 18–10
Glucose-alanine cycle: alanine ?→ glucose 18–9
Nucleotide synthesis: amino acids ?→ purines, pyrimidines 22–33; 22–36
Hormone and neurotransmitter synthesis 22–29
Fat catabolism
H9252 Oxidation of fatty acids: fatty acid ?→ acetyl-CoA 17–8
Oxidation of ketone bodies: H9252-hydroxybutyrate ?→ acetyl-CoA ?→ CO
2
citric acid cycle 17–19
Fat anabolism
Fatty acid synthesis: acetyl-CoA ?→ fatty acids 21–5
Triacylglycerol synthesis: acetyl-CoA ?→ fatty acids ?→ triacylglycerol 21–18; 21–19
Ketone body formation: acetyl-CoA ?→ acetoacetate, H9252-hydroxybutyrate 17–18
Cholesterol and cholesteryl ester synthesis: acetyl-CoA ?→ cholesterol ?→ cholesteryl esters 21–33 to 21–37
Phospholipid synthesis: fatty acids ?→ phospholipids 21–17; 21–23 to 21–28
8885d_c23_881-919 3/1/04 1:26 PM Page 894 mac76 mac76:385_reb:
1 Glucose 6-phosphate is dephosphorylated by
glucose 6-phosphatase to yield free glucose (p. 547),
which is exported to replenish blood glucose. Export is
the predominant pathway when glucose 6-phosphate is
in limited supply, because the blood glucose concentra-
tion must be kept sufficiently high (4 mM) to provide
adequate energy for the brain and other tissues. 2 Glu-
cose 6-phosphate not immediately needed to form blood
glucose is converted to liver glycogen, or it has one of
several other fates. Following glucose 6-phosphate
breakdown by glycolysis and decarboxylation of the
pyruvate (by the pyruvate dehydrogenase reaction), 3
the acetyl-CoA so formed can be oxidized for energy
production by the citric acid cycle, with ensuing elec-
tron transfer and oxidative phosphorylation yielding
ATP. (Normally, however, fatty acids are the preferred
fuel for energy production in hepatocytes.) 4 Acetyl-
CoA can also serve as the precursor of fatty acids, which
are incorporated into TAGs and phospholipids, and cho-
lesterol. Much of the lipid synthesized in the liver is
23.2 Tissue-Specific Metabolism: The Division of Labor 895
transported to other tissues by blood lipoproteins. 5
Finally, glucose 6-phosphate can enter the pentose
phosphate pathway, yielding both reducing power
(NADPH), needed for the biosynthesis of fatty acids and
cholesterol, and D-ribose 5-phosphate, a precursor for
nucleotide biosynthesis. NADPH is also an essential co-
factor in the detoxification and elimination of many
drugs and other xenobiotics metabolized in the liver.
Amino Acids Amino acids that enter the liver follow sev-
eral important metabolic routes (Fig. 23–14). 1 They
are precursors for protein synthesis, a process discussed
in Chapter 27. The liver constantly renews its own pro-
teins, which have a relatively high turnover rate (aver-
age half-life of only a few days), and is also the site of
biosynthesis of most plasma proteins. 2 Alternatively,
amino acids pass in the bloodstream to other organs, to
be used in the synthesis of tissue proteins. 3 Other
amino acids are precursors in the biosynthesis of nu-
cleotides, hormones, and other nitrogenous compounds
in the liver and other tissues.
4a Amino acids not needed as biosynthetic precur-
sors are transaminated or deaminated and degraded to
yield pyruvate and citric acid cycle intermediates, with
various fates; 4b the ammonia released is converted to
the excretory product urea. 5 Pyruvate can be con-
verted to glucose and glycogen via gluconeogenesis or
6 it can be converted to acetyl-CoA, which has several
possible fates. 7 It can be oxidized via the citric acid
cycle and 8 oxidative phosphorylation to produce ATP,
or 9 converted to lipids for storage. 10 Citric acid cy-
cle intermediates can be siphoned off into glucose syn-
thesis by gluconeogenesis.
The liver also metabolizes amino acids that arrive in-
termittently from other tissues. The blood is adequately
supplied with glucose just after the digestion and ab-
sorption of dietary carbohydrate or, between meals, by
the conversion of liver glycogen to blood glucose. During
the interval between meals, especially if prolonged, some
muscle protein is degraded to amino acids. These amino
acids donate their amino groups (by transamination) to
pyruvate, the product of glycolysis, to yield alanine, which
11 is transported to the liver and deaminated. Hepato-
cytes convert the resulting pyruvate to blood glucose (via
gluconeogenesis 5 ), and the ammonia to urea for ex-
cretion. One benefit of this glucose-alanine cycle (see Fig.
18–9) is the smoothing out of fluctuations in blood glu-
cose between meals. The amino acid deficit incurred in
the muscles is made up after the next meal by incoming
dietary amino acids.
Lipids The fatty acid components of the lipids entering
hepatocytes also have several different fates (Fig.
23–15). 1 Some are converted to liver lipids. 2 Under
most circumstances, fatty acids are the primary oxida-
tive fuel in the liver. Free fatty acids may be activated
FIGURE 23–13 Metabolic pathways for glucose 6-phosphate in the
liver. Here and in Figures 23–14 and 23–15, anabolic pathways are
shown leading upward, catabolic pathways leading downward, and
distribution to other organs horizontally. The numbered processes in
each figure are described in the text.
Cholesterol
Fatty
acids
citric
acid
cycle
Acetyl-CoA
e
H11002
O
2
H
2
O
ADP H11001 P
i
oxidative
phosphorylation
3
ATP
3
4
glycolysis
Pyruvate
Blood
glucose
Ribose 5-
phosphate
NADPH
Glucose 6-
phosphate
1
Liver glycogen
Hepatocyte
2
Nucleotides
pentose
phosphate
pathway
5
Triacylglycerols,
phospholipids
CO
2
8885d_c23_881-919 3/1/04 1:26 PM Page 895 mac76 mac76:385_reb:
and oxidized to yield acetyl-CoA and NADH. 3 and 4
The acetyl-CoA is further oxidized via the citric acid cy-
cle, and the oxidations in the cycle drive the synthesis
of ATP by oxidative phosphorylation. 5 Excess acetyl-
CoA released by oxidation of fatty acids and not required
by the liver is converted to the ketone bodies, acetoac-
etate and H9252-hydroxybutyrate; these circulate in the blood
to other tissues, to be used as fuel for the citric acid cy-
cle. Ketone bodies may be regarded as a transport form
of acetyl groups. They can supply a significant fraction
of the energy in some extrahepatic tissues—up to one-
third in the heart, and as much as 60% to 70% in the
brain during prolonged fasting. 6 Some of the acetyl-
CoA derived from fatty acids (and from glucose) is used
for the biosynthesis of cholesterol, which is required for
membrane synthesis. Cholesterol is also the precursor of
all steroid hormones and of the bile salts, which are es-
sential for the digestion and absorption of lipids.
The final two metabolic fates of lipids involve spe-
cialized mechanisms for the transport of insoluble lipids
in the blood. 7 Fatty acids are converted to the phos-
pholipids and TAGs of plasma lipoproteins, which carry
lipids to adipose tissue for storage as TAGs. 8 Some
free fatty acids become bound to serum albumin and are
carried to the heart and skeletal muscles, which absorb
and oxidize free fatty acids as a major fuel. Serum al-
bumin is the most abundant plasma protein; one mole-
cule of serum albumin can carry up to 10 molecules of
free fatty acid to the tissues where the fatty acids are
released and consumed.
The liver thus serves as the body’s distribution cen-
ter, exporting nutrients in the correct proportions to
other organs, smoothing out fluctuations in metabolism
caused by intermittent food intake, and processing ex-
cess amino groups into urea and other products to be
disposed of by the kidneys. Certain nutrients are stored
in the liver, including Fe ions and vitamin A. The liver
also detoxifies foreign organic compounds, such as
drugs, food additives, preservatives, and other possibly
harmful agents with no food value. Detoxification often
involves the cytochrome P-450–dependent hydroxyla-
tion of relatively insoluble organic compounds, making
them sufficiently soluble for further breakdown and ex-
cretion (see Box 21–1).
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism896
citric
acid
cycle
Fatty
acids
Lipids
Acetyl-CoA
O
2
H
2
O
ADP H11001 P
i
oxidative
phosphorylation
ATP
7
6
9
Pyruvate
Amino
acids
in blood
Amino acids
2
Liver
proteins
1
11
CO
2
Plasma
proteins
Tissue
proteins
Glucose
Glycogen
10
4a
Glucose
Glycogen
in muscle
5
gluconeogenesis
NH
3
Urea
Alanine
4b
urea cycle
Nucleotides,
hormones,
porphyrins
8
3
e
H11002
Hepatocyte
Amino
acids in
muscle
FIGURE 23–14 Metabolism of amino acids in the liver.
8885d_c23_881-919 3/1/04 1:26 PM Page 896 mac76 mac76:385_reb:
Adipose Tissue Stores and Supplies Fatty Acids
Adipose tissue, which consists of adipocytes (fat cells)
(Fig. 23–16), is amorphous and widely distributed in the
body: under the skin, around the deep blood vessels,
and in the abdominal cavity. It typically makes up about
15% of the mass of a young adult human, with approx-
imately 65% of this mass in the form of triacylglycerols.
Adipocytes are metabolically very active, responding
quickly to hormonal stimuli in a metabolic interplay with
the liver, skeletal muscles, and the heart.
Like other cell types, adipocytes have an active gly-
colytic metabolism, use the citric acid cycle to oxidize
pyruvate and fatty acids, and carry out oxidative phos-
phorylation. During periods of high carbohydrate intake,
adipose tissue can convert glucose (via pyruvate and
acetyl-CoA) to fatty acids, convert the fatty acids to
TAGs, and store them as large fat globules—although,
in humans, much of the fatty acid synthesis occurs in
hepatocytes. Adipocytes store TAGs arriving from the
liver (carried in the blood as VLDLs; see Fig. 21–40a)
and from the intestinal tract (carried in chylomicrons),
particularly after meals rich in fat.
When fuel demand rises, lipases in adipocytes hy-
drolyze stored TAGs to release free fatty acids, which
can travel in the bloodstream to skeletal muscles and
the heart. The release of fatty acids from adipocytes is
greatly accelerated by epinephrine, which stimulates
the cAMP-dependent phosphorylation of perilipin; this
gives triacylglycerol lipase access to TAGs in the lipid
droplet. The hormone-sensitive lipase is also stimulated
by phosphorylation, but this is not the main cause of in-
creased lipolysis (see Fig. 17–3). Insulin counterbal-
ances this effect of epinephrine, decreasing the activity
of triacylglycerol lipase.
23.2 Tissue-Specific Metabolism: The Division of Labor 897
Liver lipids
Fatty
acids
Plasma
lipoproteins
Free fatty
acids in blood
Ketone
bodies
in blood
Bile salts
Steroid
hormones
Cholesterol
b oxidation
citric
acid
cycle
Acetyl-CoA
O
2
H
2
O
ADP H11001 P
i
oxidative
phosphorylation
ATP
3
6
8
5
7
1
2
CO
2
4
e
H11002
Hepatocyte
NADH
FIGURE 23–15 Metabolism of fatty acids in the liver.
10 H9262m
FIGURE 23–16 Scanning electron micrograph of human adipocytes.
In fat tissues, capillaries and collagen fibers form a supporting net-
work around spherical adipocytes. Almost the entire volume of these
metabolically active cells is taken up by fat droplets.
8885d_c23_881-919 3/1/04 1:26 PM Page 897 mac76 mac76:385_reb:
The breakdown and synthesis of TAGs in adipose
tissues constitute a substrate cycle; up to 70% of the
fatty acids released by triacylglycerol lipase are reester-
ified in adipocytes, re-forming TAGs. Recall from Chap-
ter 15 that such substrate cycles allow fine regulation
of the rate and direction of flow of intermediates
through a bidirectional pathway. In adipose tissue, glyc-
erol liberated by triacylglycerol lipase cannot be reused
in the synthesis of TAGs, because adipocytes lack the
enzyme glycerol kinase. Instead, the glycerol phosphate
required for TAG synthesis is made from pyruvate by
glyceroneogenesis, involving the cytosolic enzyme PEP
carboxykinase (see Fig. 21–22). This enzyme is one tar-
get of the drugs (thiazolidinediones) used in the treat-
ment of type II diabetes, raising the possibility that de-
fective regulation of cytosolic PEP carboxykinase in fat
tissue may be a causative factor in type II diabetes.
Human infants, and many hibernating animals, have
adipose tissue called brown fat, which is specialized to
generate heat rather than ATP during oxidation of fatty
acids. Adult humans have very little brown fat tissue.
Muscles Use ATP for Mechanical Work
Metabolism in the cells of skeletal muscle—the my-
ocytes—is specialized to generate ATP as the immedi-
ate source of energy for contraction. Moreover, skeletal
muscle is adapted to do its mechanical work in an
intermittent fashion, on demand. Sometimes skeletal
muscles must work at their maximum capacity for a
short time, as in a 100 m sprint; at other times more
prolonged work is required, as in running a marathon
or extended physical labor.
There are two general classes of muscle tissue,
which differ in physiological role and fuel utilization.
Slow-twitch muscle, also called red muscle, provides
relatively low tension but is highly resistant to fatigue.
It produces ATP by the relatively slow but steady
process of oxidative phosphorylation. Red muscle is very
rich in mitochondria and is served by very dense net-
works of blood vessels, which bring the oxygen essen-
tial to ATP production. It is the cytochromes in mito-
chondria and the hemoglobin in blood that give the
tissue its characteristic red color. Fast-twitch muscle,
or white muscle, has fewer mitochondria than red mus-
cle and is less well supplied with blood vessels, but it
can develop greater tension, and do so faster. White
muscle is quicker to fatigue, because when active, it uses
ATP faster than it can replace it. There is a genetic com-
ponent to the proportion of red and white muscle in any
individual; with training, the endurance of fast-twitch
muscle can be improved.
Skeletal muscle can use free fatty acids, ketone bod-
ies, or glucose as fuel, depending on the degree of mus-
cular activity (Fig. 23–17). In resting muscle, the pri-
mary fuels are free fatty acids from adipose tissue and
ketone bodies from the liver. These are oxidized and de-
graded to yield acetyl-CoA, which enters the citric acid
cycle for oxidation to CO
2
. The ensuing transfer of elec-
trons to O
2
provides the energy for ATP synthesis by
oxidative phosphorylation. Moderately active muscle
uses blood glucose in addition to fatty acids and ketone
bodies. The glucose is phosphorylated, then degraded
by glycolysis to pyruvate, which is converted to acetyl-
CoA and oxidized via the citric acid cycle and oxidative
phosphorylation.
In maximally active fast-twitch muscles, the de-
mand for ATP is so great that the blood flow cannot pro-
vide O
2
and fuels fast enough to supply sufficient ATP
by aerobic respiration alone. Under these conditions,
stored muscle glycogen is broken down to lactate by fer-
mentation (p. 523). Each glucose unit degraded yields
three ATP, because phosphorolysis of glycogen produces
glucose 6-phosphate, sparing the ATP normally con-
sumed in the hexokinase reaction. Lactic acid fermen-
tation thus responds to an increased need for ATP more
quickly than does oxidative phosphorylation, supple-
menting basal ATP production that results from aerobic
oxidation of other fuels via the citric acid cycle and res-
piratory chain. The use of blood glucose and muscle
glycogen as fuels for muscular activity is greatly en-
hanced by the secretion of epinephrine, which stimu-
lates both the release of glucose from liver glycogen and
the breakdown of glycogen in muscle tissue.
The relatively small amount of glycogen in skeletal
muscle (about 1% of its total weight) limits the amount
of glycolytic energy available during all-out exertion.
Moreover, the accumulation of lactate and consequent
decrease in pH in maximally active muscles reduces their
efficiency. Skeletal muscle, however, contains another
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism898
ATP
Creatine
ADP H11001 P
i
CO
2
Lactate
Muscle
contraction
Bursts of
heavy activity
Muscle glycogen
Bursts of
heavy activity
Fatty acids,
ketone bodies,
blood glucose
Light activity
or rest
Phosphocreatine
FIGURE 23–17 Energy sources for muscle contraction. Different fu-
els are used for ATP synthesis during bursts of heavy activity and dur-
ing light activity or rest. Phosphocreatine can rapidly supply ATP.
8885d_c23_881-919 3/1/04 1:26 PM Page 898 mac76 mac76:385_reb:
source of ATP, in the form of phosphocreatine (10 to
30 mM), which can rapidly regenerate ATP from ADP by
the creatine kinase reaction:
During periods of active contraction and glycolysis,
this reaction proceeds predominantly in the direction of
ATP synthesis; during recovery from exertion, the same
enzyme resynthesizes phosphocreatine from creatine at
the expense of ATP.
After a period of intense muscular activity, the in-
dividual continues breathing heavily for some time, us-
ing much of the extra O
2
for oxidative phosphorylation
in the liver. The ATP produced is used for gluconeoge-
nesis from lactate that has been carried in the blood
from the muscles. The glucose thus formed returns to
the muscles to replenish their glycogen, completing the
Cori cycle (Fig. 23–18; see also Box 15–1).
Heart muscle differs from skeletal muscle in that
it is continuously active in a regular rhythm of
contraction and relaxation, and it has a completely aer-
obic metabolism at all times. Mitochondria are much
more abundant in heart muscle than in skeletal muscle,
making up almost half the volume of the cells (Fig.
23–19). The heart uses as its fuel mainly free fatty acids,
but also some glucose and ketone bodies taken up from
the blood; these fuels are oxidized via the citric acid cy-
cle and oxidative phosphorylation to generate ATP. Like
skeletal muscle, heart muscle does not store lipids or
glycogen in large amounts. It does have small amounts
of reserve energy in the form of phosphocreatine,
enough for a few seconds of contraction. Because the
heart is normally aerobic and obtains its energy from
oxidative phosphorylation, the failure of O
2
to reach a
portion of the heart muscle when the blood vessels are
blocked by lipid deposits (atherosclerosis) or blood clots
(coronary thrombosis) can cause that region of the
heart muscle to die. This is what happens in myocardial
infarction, more commonly known as a heart attack. ■
23.2 Tissue-Specific Metabolism: The Division of Labor 899
NH
2
N
Creatine
COO
H11002
C
H11001
NH
2
CH
3
CH
2
ATP H11001
Phosphocreatine
CH
2
C
H11002
O
H
P O
N
H11001
NH
2
CH
3
N
O
H11002
COO
H11002
H11001 ADP
during
activity
during
recovery
Muscle: ATP produced by
glycolysis for rapid contraction.
Lactate Glycogen
Lactate Glucose
ATP
ATP
Blood
lactate
Blood
glucose
Liver: ATP used in synthesis
of glucose (gluconeogenesis)
during recovery.
FIGURE 23–18 Metabolic cooperation between skeletal muscle and
the liver. Extremely active muscles use glycogen as energy source,
generating lactate via glycolysis. During recovery, some of this lactate
is transported to the liver and converted to glucose via gluconeogen-
esis. This glucose is released to the blood and returned to the muscles
to replenish their glycogen stores. The overall pathway (glucose →
lactate → glucose) constitutes the Cori cycle.
FIGURE 23–19 Electron micrograph of heart muscle. In the profuse
mitochondria of heart tissue, pyruvate, fatty acids, and ketone bodies
are oxidized to drive ATP synthesis. This steady aerobic metabolism
allows the human heart to pump blood at a rate of nearly 6 L/min, or
about 350 L/hr—or 200 H11003 10
6
L over 70 years.
1 mH9262
8885d_c23_881-919 3/1/04 1:26 PM Page 899 mac76 mac76:385_reb:
The Brain Uses Energy for Transmission
of Electrical Impulses
The metabolism of the brain is remarkable in several re-
spects. The neurons of the adult mammalian brain nor-
mally use only glucose as fuel (Fig. 23–20). (Astrocytes,
the other major cell type in the brain, can oxidize fatty
acids.) The brain has a very active respiratory metabo-
lism (Fig. 23–21); it uses O
2
at a fairly constant rate, ac-
counting for almost 20% of the total O
2
consumed by
the body at rest. Because the brain contains very little
glycogen, it is constantly dependent on incoming glu-
cose from the blood. Should blood glucose fall signifi-
cantly below a critical level for even a short time, severe
and sometimes irreversible changes in brain function
may result.
Although the neurons of the brain cannot directly
use free fatty acids or lipids from the blood as fuels, they
can, when necessary, use H9252-hydroxybutyrate (a ketone
body), which is formed from fatty acids in the liver. The
capacity of the brain to oxidize H9252-hydroxybutyrate via
acetyl-CoA becomes important during prolonged fasting
or starvation, after liver glycogen has been depleted, be-
cause it allows the brain to use body fat as an energy
source. This spares muscle proteins—until they become
the brain’s ultimate source of glucose (via gluconeogen-
esis in the liver) during severe starvation.
Neurons oxidize glucose by glycolysis and the cit-
ric acid cycle, and the flow of electrons from these ox-
idations through the respiratory chain provides almost
all the ATP used by these cells. Energy is required to
create and maintain an electrical potential across the
neuronal plasma membrane. The membrane contains an
electrogenic ATP-driven antiporter, the Na
H11001
K
H11001
ATPase,
which simultaneously pumps 2 K
H11001
ions into and 3 Na
H11001
ions out of the neuron (see Fig. 11–37). The resulting transmembrane potential changes transiently as an elec-
trical signal (action potential) sweeps from one end of
a neuron to the other (see Fig. 12–5). Action potentials
are the chief mechanism of information transfer in the
nervous system, so a depletion of ATP in neurons has
disastrous effects on all activities coordinated by neu-
ronal signaling.
Blood Carries Oxygen, Metabolites, and Hormones
Blood mediates the metabolic interactions among all tis-
sues. It transports nutrients from the small intestine to
the liver, and from the liver and adipose tissue to other
organs; it also transports waste products from the tis-
sues to the kidneys for excretion. Oxygen moves in the
bloodstream from the lungs to the tissues, and CO
2
gen-
erated by tissue respiration returns via the bloodstream
to the lungs for exhalation. Blood also carries hormonal
signals from one tissue to another. In its role as signal
carrier, the circulatory system resembles the nervous
system; both regulate and integrate the activities of dif-
ferent organs.
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism900
CO
2
ADP + P
i
Ketone bodies
Electrogenic transport
by Na
+
K
+
ATPase
Starvation
ATP
Normal diet
Glucose
FIGURE 23–20 Energy sources in the brain vary with nutritional
state. The ketone body used by the brain is H9252-hydroxybutyrate.
(a)
(b)
12.00 2.00
mg/100 g /min
FIGURE 23–21 Glucose metabolism in the brain. The technique of
positron emission tomography (PET) scanning shows metabolic activ-
ity in specific regions of the brain. PET scans allow visualization of
isotopically labeled glucose in precisely localized regions of the brain
of a living person, in real time. A positron-emitting glucose analog (2-
[
18
F]-fluoro-2-deoxy-D-glucose) is injected into the bloodstream; a few
seconds later, a PET scan shows how much of the glucose has been
taken up by each region of the brain—a measure of metabolic activ-
ity. Shown here are PET scans of front-to-back cross sections of the
brain at three levels, from the top (at the left) downward (to the right).
The scans compare glucose metabolism (in mg/100 g/min) when the
experimental subject (a) is rested and (b) has been deprived of sleep
for 48 hours.
8885d_c23_881-919 3/1/04 1:26 PM Page 900 mac76 mac76:385_reb:
The average adult human has 5 to 6 L of blood. Al-
most half of this volume is occupied by three types of
blood cells (Fig. 23–22): erythrocytes (red cells), filled
with hemoglobin and specialized for carrying O
2
and
CO
2
; much smaller numbers of leukocytes (white cells)
of several types (including lymphocytes, also found in
lymphatic tissue), which are central to the immune sys-
tem that defends against infections; and platelets,
which help to mediate blood clotting. The liquid portion
is the blood plasma, which is 90% water and 10%
solutes. Dissolved or suspended in the plasma is a large
variety of proteins, lipoproteins, nutrients, metabolites,
waste products, inorganic ions, and hormones. More
than 70% of the plasma solids are plasma proteins
(Fig. 23–22), primarily immunoglobulins (circulating
antibodies), serum albumin, apolipoproteins involved in
the transport of lipids, transferrin (for iron transport),
and blood-clotting proteins such as fibrinogen and
prothrombin.
The ions and low molecular weight solutes in blood
plasma are not fixed components but are in constant
flux between blood and various tissues. Dietary uptake
of the inorganic ions that are the predominant elec-
trolytes of blood and cytosol (Na
H11001
, K
H11001
, and Ca
2H11001
) is, in
general, counterbalanced by their excretion in the urine.
For many blood components, something near a dynamic
steady state is achieved; the concentration of the com-
ponent changes little, although a continuous flux occurs
between the digestive tract, blood, and urine. The
plasma levels of Na
H11001
, K
H11001
, and Ca
2H11001
remain close to 140,
5, and 2.5 mM, respectively, with little change in re-
sponse to dietary intake. Any significant departure from
these values can result in serious illness or death. The
kidneys play an especially important role in maintaining
ion balance by selectively filtering waste products and
excess ions out of the blood while preventing the loss
of essential nutrients and ions.
The concentration of glucose in the plasma is also
subject to tight regulation. We have noted the
constant requirement of the brain for glucose and the
role of the liver in maintaining blood glucose in the nor-
mal range of 60 to 90 mg/100 mL. When blood glucose
in a human drops to 40 mg/100 mL (the hypoglycemic
condition), the person experiences discomfort and men-
tal confusion (Fig. 23–23); further reductions lead to
coma, convulsions, and in extreme hypoglycemia, death.
23.2 Tissue-Specific Metabolism: The Division of Labor 901
0
10
20
30
40
50
60
70
80
90
100
Blood
glucose
(mg/100 mL)
Normal
range
Subtle neurological signs; hunger
Release of glucagon, epinephrine, cortisol
Sweating, trembling
Lethargy
Convulsions, coma
Permanent brain damage (if prolonged)
Death
Blood plasmaCells
Inorganic components (10%)
NaCl, bicarbonate, phosphate,
CaCl
2
, MgCl
2
, KCl, Na
2
SO
4
Organic metabolites and waste
products (20%)
glucose, amino acids, lactate,
pyruvate, ketone bodies,
citrate, urea, uric acid
Plasma proteins (70%)
Major plasma proteins: serum albumin, very-low-density
lipoproteins (VLDL), low-density lipoproteins (LDL),
high-density lipoproteins (HDL), immunoglobulins (hundreds
of kinds), fibrinogen, prothrombin, many specialized transport
proteins such as transferrin
H
2
OErythrocytes
Leukocytes
Platelets
Plasma
solutes
FIGURE 23–22 The composition of blood. Whole blood can be sep-
arated into blood plasma and cells by centrifugation. About 10% of
blood plasma is solutes, of which about 10% consists of inorganic
salts, 20% small organic molecules, and 70% plasma proteins. The
major dissolved components are listed. Blood contains many other
substances, often in trace amounts. These include other metabolites,
enzymes, hormones, vitamins, trace elements, and bile pigments. Mea-
surements of the concentrations of components in blood plasma are
important in the diagnosis and treatment of many diseases.
FIGURE 23–23 Physiological effects of low blood glucose in humans.
Blood glucose levels of 40 mg/100 mL and below constitute severe
hypoglycemia.
8885d_c23_881-919 3/1/04 1:26 PM Page 901 mac76 mac76:385_reb:
Maintaining the normal concentration of glucose in the
blood is therefore a very high priority of the organism,
and a variety of regulatory mechanisms have evolved
to achieve that end. Among the most important regu-
lators of blood glucose are the hormones insulin,
glucagon, and epinephrine, as discussed in Section
23.3. ■
SUMMARY 23.2 Tissue-Specific Metabolism:
The Division of Labor
■ In mammals there is a division of metabolic
labor among specialized tissues and organs. The
liver is the central distributing and processing
organ for nutrients. Sugars and amino acids
produced in digestion cross the intestinal
epithelium and enter the blood, which carries
them to the liver. Some triacylglycerols derived
from ingested lipids also make their way to the
liver, where the constituent fatty acids are used
in a variety of processes.
■ Glucose 6-phosphate is the key intermediate in
carbohydrate metabolism. It may be polymerized
into glycogen, dephosphorylated to blood
glucose, or converted to fatty acids via acetyl-
CoA. It may undergo oxidation by glycolysis,
the citric acid cycle, and respiratory chain to
yield ATP, or enter the pentose phosphate
pathway to yield pentoses and NADPH.
■ Amino acids are used to synthesize liver and
plasma proteins, or their carbon skeletons are
converted to glucose and glycogen by
gluconeogenesis; the ammonia formed by
deamination is converted to urea.
■ The liver converts fatty acids to triacylglycerols,
phospholipids, or cholesterol and its esters, for
transport as plasma lipoproteins to adipose
tissue for storage. Fatty acids can also be
oxidized to yield ATP or to form ketone bodies,
which are circulated to other tissues.
■ Skeletal muscle is specialized to produce and
use ATP for mechanical work. During strenuous
muscular activity, glycogen is the ultimate fuel,
supplying ATP through lactic acid fermentation.
During recovery, the lactate is reconverted
(through gluconeogenesis) to glycogen and
glucose in the liver. Phosphocreatine is an
immediate source of ATP during active
contraction.
■ Heart muscle obtains nearly all its ATP from
oxidative phosphorylation.
■ The neurons of the brain use only glucose and
H9252-hydroxybutyrate as fuels, the latter being
important during fasting or starvation. The
brain uses most of its ATP for the active
transport of Na
H11001
and K
H11001
and maintenance of
the electrical potential across the neuronal
membrane.
■ The blood carries nutrients, waste products,
and hormonal signals among the organs.
23.3 Hormonal Regulation of Fuel
Metabolism
The minute-by-minute adjustments that keep the blood
glucose level near 4.5 mM involve the combined actions
of insulin, glucagon, epinephrine, and cortisol on meta-
bolic processes in many body tissues, but especially in
liver, muscle, and adipose tissue. Insulin signals these
tissues that blood glucose is higher than necessary; as
a result, cells take up excess glucose from the blood and
convert it to the storage compounds glycogen and tria-
cylglycerol. Glucagon signals that blood glucose is too
low, and tissues respond by producing glucose through
glycogen breakdown and (in liver) gluconeogenesis and
by oxidizing fats to reduce the use of glucose. Epi-
nephrine is released into the blood to prepare the mus-
cles, lungs, and heart for a burst of activity. Cortisol me-
diates the body’s response to longer-term stresses. We
discuss these hormonal regulations in the context of
three normal metabolic states—well-fed, fasted, and
starving—and look at the metabolic consequences of di-
abetes mellitus, which results from derangements in the
signaling pathways that control glucose metabolism.
The Pancreas Secretes Insulin or Glucagon
in Response to Changes in Blood Glucose
When glucose enters the bloodstream from the intes-
tine after a carbohydrate-rich meal, the resulting in-
crease in blood glucose causes increased secretion of
insulin (and decreased secretion of glucagon). Insulin
release by the pancreas is largely regulated by the level
of glucose in the blood supplied to the pancreas. The
peptide hormones insulin, glucagon, and somatostatin
are produced by clusters of specialized pancreatic cells,
the islets of Langerhans (Fig. 23–24). Each cell type of
the islets produces a single hormone: H9251 cells produce
glucagon; H9252 cells, insulin; and H9254 cells, somatostatin.
When blood glucose rises, GLUT2 transporters carry
glucose into the H9252 cells, where it is immediately con-
verted to glucose 6-phosphate by hexokinase IV (gluco-
kinase) and enters glycolysis (Fig. 23–25). The in-
creased rate of glucose catabolism raises [ATP], causing
the closing of ATP-gated K
H11001
channels in the plasma
membrane. Reduced efflux of K
H11001
depolarizes the mem-
brane, thereby opening voltage-sensitive Ca
2H11001
channels
in the plasma membrane. The resulting influx of Ca
2H11001
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism902
8885d_c23_881-919 3/1/04 1:26 PM Page 902 mac76 mac76:385_reb:
triggers the release of insulin by exocytosis. Stimuli from
the parasympathetic and sympathetic nervous systems
also stimulate and inhibit insulin release, respectively.
A simple feedback loop limits hormone release: insulin
lowers blood glucose by stimulating glucose uptake by
the tissues; the reduced blood glucose is detected by
the H9252 cell as a diminished flux through the hexokinase
reaction; this slows or stops the release of insulin. This
feedback regulation holds blood glucose concentration
nearly constant despite large fluctuations in dietary
intake.
23.3 Hormonal Regulation of Fuel Metabolism 903
H9251 cell (glucagon)
H9252 cell
(insulin)
H9254 cell
(somatostatin)
Pancreas
Blood
vessels
FIGURE 23–24 The endocrine system of the pancreas. In addition to
the exocrine cells (see Fig. 18–3b), which secrete digestive enzymes
in the form of zymogens, the pancreas contains endocrine tissue, the
islets of Langerhans. The islets contain H9251, H9252, and H9254 cells (also known
as A, B, and D cells, respectively), each cell type secreting a specific
polypeptide hormone.
Glucose
Glucose 6-phosphate
Glucose transporter
GLUT2
Extracellular space
Pancreatic b cell
hexokinase IV
(glucokinase)
oxidative
phosphorylation
glycolysis
citric acid cycle
[Ca
2+
]
Ca
2+
[ATP]
ATP-gated
K
+
channel
+
+
+
+
+
+
–
–
–
–
–
–
depolarization
Voltage-dependent
Ca
2+
channel
Insulin
granules
Insulin
secretion
Nucleus
K
+
K
+
V
m
Glucose
2
4
3
5
1
FIGURE 23–25 Glucose regulation of insulin secretion by pancreatic H9252 cells. When the blood
glucose level is high, active metabolism of glucose in the H9252 cell raises intracellular [ATP],
which leads to closing of K
H11001
channels in the plasma membrane, depolarizing the membrane.
In response to the change in membrane potential, voltage-gated Ca
2H11001
channels in the plasma
membrane open, allowing Ca
2H11001
to flow into the cell; this raises the cytosolic [Ca
2H11001
] enough to
trigger insulin release by exocytosis.
8885d_c23_881-919 3/1/04 1:26 PM Page 903 mac76 mac76:385_reb:
Insulin Counters High Blood Glucose
Insulin stimulates glucose uptake by muscle and adipose
tissue (Table 23–3), where the glucose is converted to
glucose 6-phosphate. In the liver, insulin also activates
glycogen synthase and inactivates glycogen phosphory-
lase, so that much of the glucose 6-phosphate is chan-
neled into glycogen.
Insulin also stimulates the storage of excess fuel as
fat (Fig. 23–26). In the liver, insulin activates both the
oxidation of glucose 6-phosphate to pyruvate via gly-
colysis and the oxidation of pyruvate to acetyl-CoA. If
not oxidized further for energy production, this acetyl-
CoA is used for fatty acid synthesis in the liver, and the
fatty acids are exported as the TAGs of plasma lipopro-
teins (VLDLs) to the adipose tissue. Insulin stimulates
TAG synthesis in adipocytes, from fatty acids released
from the VLDL triacylglycerols. These fatty acids are ul-
timately derived from the excess glucose taken up from
the blood by the liver. In summary, the effect of insulin
is to favor the conversion of excess blood glucose to two
storage forms: glycogen (in the liver and muscle) and
triacylglycerols (in adipose tissue) (Table 23–3).
Glucagon Counters Low Blood Glucose
Several hours after the intake of dietary carbohydrate,
blood glucose levels fall slightly because of the ongoing
oxidation of glucose by the brain and other tissues. Low-
ered blood glucose triggers secretion of glucagon and
decreases insulin release (Fig. 23–27).
Glucagon causes an increase in blood glucose con-
centration in several ways (Table 23–4). Like epineph-
rine, it stimulates the net breakdown of liver glycogen
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism904
FIGURE 23–26 The well-fed state: the lipogenic liver. Immediately
after a calorie-rich meal, glucose, fatty acids, and amino acids enter
the liver. Insulin released in response to the high blood glucose
concentration stimulates glucose uptake by the tissues. Some glucose
is exported to the brain for its energetic needs, and some to fat and
muscle tissue. In the liver, excess glucose is oxidized to acetyl-CoA,
which is used to synthesize fatty acids for export as triacylglycerols in
Pancreas
Intestine
Glucose
Pyruvate
ATP
Acetyl-
CoA
TAG
Amino acids
Urea
a-Keto acids
Protein
synthesis
Glycogen
Lymphatic
system
VLDL
TAG
Adipose
tissue
Muscle
Fatty acids
ATP
CO
2
ATP
CO
2
CO
2
Liver
Brain
to brain, adipose, muscle
Glucose
Amino
acids
Fats
Glucose
Glucose
Insulin
Insulin
TAG
VLDLs to fat and muscle tissue. The NADPH necessary for lipid
synthesis is obtained by oxidation of glucose in the pentose phosphate
pathway. Excess amino acids are converted to pyruvate and acetyl-
CoA, which are also used for lipid synthesis. Dietary fats move via the
lymphatic system, as chylomicrons, from the intestine to muscle and
fat tissues.
8885d_c23_881-919 3/1/04 1:26 PM Page 904 mac76 mac76:385_reb:
by activating glycogen phosphorylase and inactivating
glycogen synthase; both effects are the result of phos-
phorylation of the regulated enzymes, triggered by
cAMP. Glucagon inhibits glucose breakdown by glycol-
ysis in the liver and stimulates glucose synthesis by
gluconeogenesis. Both effects result from lowering the
concentration of fructose 2,6-bisphosphate, an allosteric
inhibitor of the gluconeogenic enzyme fructose 1,6-bis-
phosphatase (FBPase-1) and an activator of phospho-
fructokinase-1. Recall that [fructose 2,6-bisphosphate]
is ultimately controlled by a cAMP-dependent protein
phosphorylation reaction (see Fig. 15–23). Glucagon
also inhibits the glycolytic enzyme pyruvate kinase (by
promoting its cAMP-dependent phosphorylation), thus
blocking the conversion of phosphoenolpyruvate to
pyruvate and preventing oxidation of pyruvate via the
citric acid cycle. The resulting accumulation of phos-
phoenolpyruvate favors gluconeogenesis. This effect is
augmented by glucagon’s stimulation of the synthesis of
the gluconeogenic enzyme PEP carboxykinase. By stim-
ulating glycogen breakdown, preventing glycolysis, and
promoting gluconeogenesis in hepatocytes, glucagon
23.3 Hormonal Regulation of Fuel Metabolism 905
Metabolic effect Target enzyme
↑ Glucose uptake (muscle, adipose) ↑ Glucose transporter (GLUT4)
↑ Glucose uptake (liver) ↑ Glucokinase (increased expression)
↑ Glycogen synthesis (liver, muscle) ↑ Glycogen synthase
↓ Glycogen breakdown (liver, muscle) ↓ Glycogen phosphorylase
↑ Glycolysis, acetyl-CoA production (liver, muscle) ↑ PFK-1 (by ↑ PFK-2)
↑ Pyruvate dehydrogenase complex
↑ Fatty acid synthesis (liver) ↑ Acetyl-CoA carboxylase
↑ Triacylglycerol synthesis (adipose tissue) ↑ Lipoprotein lipase
TABLE 23–3 Effects of Insulin on Blood Glucose: Uptake of Glucose by Cells and
Storage as Triacylglycerols and Glycogen
Pancreas
Amino
acids
TAG
Adipose
tissue
Muscle
Liver
Brain
Glucagon
ATP
Ketone
bodies
CO
2
CO
2
Fatty
acids
Fatty
acids
ATP
Protein
Ketone
bodies
ATP
Proteins
CO
2
CO
2
Ketone
bodies
Glycogen
Pyruvate
gluconeogenesis
Glucose Glucose
Glycerol
Glucose 6-phosphate
FIGURE 23–27 The fasting state: the gluco-
genic liver. After some hours without a meal,
the liver becomes the principal source of
glucose for the brain. Liver glycogen is
broken down, and the glucose 1-phosphate
produced is converted to glucose 6-phos-
phate, then to free glucose, which is
released into the bloodstream. Amino acids
from the degradation of proteins and
glycerol from the breakdown of TAGs in
adipose tissue are used for gluconeogenesis.
The liver uses fatty acids as its principal fuel,
and excess acetyl-CoA is converted to
ketone bodies for export to other tissues for
fuel; the brain is especially dependent on
this fuel when glucose is in short supply.
8885d_c23_881-919 3/1/04 1:26 PM Page 905 mac76 mac76:385_reb:
enables the liver to export glucose, restoring blood glu-
cose to its normal level.
Although its primary target is the liver, glucagon
(like epinephrine) also affects adipose tissue, activating
TAG breakdown by causing cAMP-dependent phosphor-
ylation of perilipin and triacylglycerol lipase. The acti-
vated lipase liberates free fatty acids, which are ex-
ported to the liver and other tissues as fuel, sparing
glucose for the brain. The net effect of glucagon is there-
fore to stimulate glucose synthesis and release by the
liver and to mobilize fatty acids from adipose tissue, to
be used instead of glucose as fuel for tissues other than
the brain (Table 23–4). All these effects of glucagon are
mediated by cAMP-dependent protein phosphorylation.
During Fasting and Starvation, Metabolism Shifts
to Provide Fuel for the Brain
The fuel reserves of a healthy adult human are of three
types: glycogen stored in the liver and, in relatively small
quantities, in muscles; large quantities of triacylglycerols
in adipose tissues; and tissue proteins, which can be de-
graded when necessary to provide fuel (Table 23–5).
In the first few hours after a meal, the blood glucose
level is diminished slightly, and tissues receive glucose
released from liver glycogen. There is little or no syn-
thesis of lipids. By 24 hours after a meal, blood glucose
has fallen further, insulin secretion has slowed, and glu-
cagon secretion has increased. These hormonal signals
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism906
Caloric equivalent Estimated survival
Type of fuel Weight (kg) (thousands of kcal (kJ)) (months)
*
Normal-weight, 70 kg man
Triacylglycerols (adipose tissue) 15 141 (589)
Proteins (mainly muscle) 6 24 (100)
Glycogen (muscle, liver) 0.225 0.90 (3.8)
Circulating fuels (glucose, fatty acids, 0.023 0.10 (0.42)
triacylglycerols, etc.)
Total 166 (694) 3
Obese, 140 kg man
Triacylglycerols (adipose tissue) 80 752 (3,140)
Proteins (mainly muscle) 8 32 (134)
Glycogen (muscle, liver) 0.23 0.92 (3.8)
Circulating fuels 0.025 0.11 (0.46)
Total 785 (3,280) 14
TABLE 23–5 Available Metabolic Fuels in a Normal-Weight 70 kg Man and in an Obese 140 kg Man
at the Beginning of a Fast
TABLE 23–4 Effects of Glucagon on Blood Glucose: Production and Release of Glucose by the Liver
Metabolic effect Effect on glucose metabolism Target enzyme
↑ Glycogen breakdown (liver) Glycogen ?→ glucose ↑ Glycogen phosphorylase
↓ Glycogen synthesis (liver) Less glucose stored as glycogen ↓ Glycogen synthase
↓ Glycolysis (liver) Less glucose used as fuel in liver ↓ PFK-1
↑ Gluconeogenesis (liver) Amino acids ↑ FBPase-2
Glycerol
?→ glucose ↓ Pyruvate kinase
Oxaloacetate ↑ PEP carboxykinase
↑ Fatty acid mobilization (adipose tissue) Less glucose used as fuel by liver, muscle ↑ Triacylglycerol lipase
Perilipin phosphorylation
↑ Ketogenesis Provides alternative to glucose as ↑ Acetyl-CoA carboxylase
energy source for brain
*
Survival time is calculated on the assumption of a basal energy expenditure of 1,800 kcal/day.
?
?
?
?
?
8885d_c23_881-919 3/1/04 1:26 PM Page 906 mac76 mac76:385_reb:
mobilize triacylglycerols, which now become the primary
fuel for muscle and liver. Figure 23–28 shows the re-
sponses to prolonged fasting. 1 To provide glucose for
the brain, the liver degrades certain proteins—those
most expendable in an organism not ingesting food.
Their nonessential amino acids are transaminated or
deaminated (Chapter 18), and 2 the extra amino
groups are converted to urea, which is exported via the
bloodstream to the kidney and excreted.
Also in the liver, 3 the carbon skeletons of gluco-
genic amino acids are converted to pyruvate or inter-
mediates of the citric acid cycle. 4 These intermedi-
ates, as well as the glycerol 5 derived from triacyl-
glycerols in adipose tissue, provide the starting materi-
als for gluconeogenesis in the liver, yielding glucose for
the brain. Eventually the use of citric acid cycle inter-
mediates for gluconeogenesis depletes oxaloacetate, in-
hibiting entry of acetyl-CoA into the citric acid cycle. 6
Acetyl-CoA produced by fatty acid oxidation now accu-
mulates, favoring 7 the formation of acetoacetyl-CoA
and ketone bodies in the liver. After a few days of fast-
ing, the levels of ketone bodies in the blood rise (Fig.
23.3 Hormonal Regulation of Fuel Metabolism 907
Acetyl-CoA
Protein degradation
yields glucogenic
amino acids.
1
Citric acid cycle
intermediates are
diverted to
gluconeogenesis.
3
Acetyl-CoA accumulation
favors ketone body synthesis.
7
Lack of oxaloacetate
prevents acetyl-CoA
entry into the citric acid cycle;
acetyl-CoA accumulates.
6
Hepatocyte
(2H11003)
P
i
Oxaloacetate
Protein
Amino
acids
NH
3
Glucose
6-phosphate
Phosphoenol-
pyruvate
Citrate
Ketone bodies are
exported via the
bloodstream
to the brain, which
uses them as fuel.
8
Ketone bodies
Acetoacetyl-CoA
Fatty acids
(imported from
adipose tissue)
are oxidized as fuel,
producing acetyl-CoA.
5
Glucose is
exported to the
brain via the
bloodstream.
4
Glucose
Urea is
exported
to the kidney
and excreted
in urine.
2
Urea
Fatty acids
FIGURE 23–28 Fuel metabolism in the liver during prolonged fast-
ing or in uncontrolled diabetes mellitus. After depletion of stored car-
bohydrates, 1 to 4 proteins become an important source of glu-
cose, produced from glucogenic amino acids by gluconeogenesis. 5
to 8 Fatty acids imported from adipose tissue are converted to ke-
tone bodies for export to the brain. Broken arrows represent reactions
with reduced flux under these conditions. The steps are further de-
scribed in the text.
8885d_c23_881-919 3/1/04 1:26 PM Page 907 mac76 mac76:385_reb:
23–29) as these fuels are exported from the liver to the
heart, skeletal muscle, and brain, which use them in-
stead of glucose ( 8 ).
Acetyl-CoA is a critical regulator of the fate of pyru-
vate; it allosterically inhibits pyruvate dehydrogenase
and stimulates pyruvate carboxylase (see Fig. 15–20).
In these ways acetyl-CoA prevents it own further pro-
duction from pyruvate while stimulating the conversion
of pyruvate to oxaloacetate, the first step in gluconeo-
genesis.
Triacylglycerols stored in the adipose tissue of a
normal-weight adult could provide enough fuel to main-
tain a basal rate of metabolism for about three months;
a very obese adult has enough stored fuel to endure a
fast of more than a year (Table 23–5). When fat reserves
are gone, the degradation of essential proteins begins;
this leads to loss of heart and liver function, and even-
tually death. Stored fat can provide adequate energy
(calories) during a fast or rigid diet, but vitamins and
minerals must be provided, and sufficient dietary gluco-
genic amino acids are needed to replace those being
used for gluconeogenesis. Rations for those on a weight-
reduction diet are therefore commonly fortified with vi-
tamins, minerals, and amino acids or proteins.
Epinephrine Signals Impending Activity
When an animal is confronted with a stressful situation
that requires increased activity—fighting or fleeing, in
the extreme case—neuronal signals from the brain trig-
ger the release of epinephrine and norepinephrine from
the adrenal medulla. Both hormones dilate the respira-
tory passages to facilitate the uptake of O
2
, increase the
rate and strength of the heartbeat, and raise the blood
pressure, thereby promoting the flow of O
2
and fuels to
the tissues (Table 23–6).
Epinephrine acts primarily on muscle, adipose, and
liver tissues. It activates glycogen phosphorylase and in-
activates glycogen synthase by cAMP-dependent phos-
phorylation of the enzymes, thus stimulating the con-
version of liver glycogen to blood glucose, the fuel for
anaerobic muscular work. Epinephrine also promotes
the anaerobic breakdown of muscle glycogen by lactic
acid fermentation, stimulating glycolytic ATP formation.
The stimulation of glycolysis is accomplished by raising
the concentration of fructose 2,6-bisphosphate, a potent
allosteric activator of the key glycolytic enzyme phos-
phofructokinase-1 (see Figs 15–22, 15–23). Epinephrine
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism908
Glucose
Fatty acids
Ketone bodies
0
8
7
6
5
4
3
2
1
Plasma concentration (m
M
)
24
Days of starvation
68
FIGURE 23–29 Concentrations of fatty acids, glucose, and ketone
bodies in the plasma during the first week of starvation. Despite the
hormonal mechanisms for maintaining the level of glucose in the
blood, it begins to diminish after two days of fasting. The level of ke-
tone bodies, almost immeasurable before the fast, rises dramatically
after 2 to 4 days of fasting. These water-soluble ketones, acetoacetate
and H9252-hydroxybutyrate, supplement glucose as an energy source dur-
ing a long fast. Fatty acids cannot serve as a fuel for the brain; they
do not cross the blood-brain barrier.
TABLE 23–6 Physiological and Metabolic Effects of Epinephrine: Preparation for Action
Immediate effect Overall effect
Physiological
↑ Heart rate
↑ Blood pressure Increase delivery of O
2
to tissues (muscle)
↑ Dilation of respiratory passages
Metabolic
↑ Glycogen breakdown (muscle, liver)
↓ Glycogen synthesis (muscle, liver) Increase production of glucose for fuel
↑ Gluconeogenesis (liver)
↑ Glycolysis (muscle) Increases ATP production in muscle
↑ Fatty acid mobilization (adipose tissue) Increases availability of fatty acids as fuel
↑ Glucagon secretion
↓ Insulin secretion
Reinforce metabolic effects of epinephrine
?
?
?
?
?
?
?
?
?
?
?
?
?
8885d_c23_881-919 3/1/04 1:26 PM Page 908 mac76 mac76:385_reb:
CH
3
C
O
CH
2
COO
H11002
H11001 H
2
O CH
3
C
O
CH
3
H11001HCO
3
H11002
Acetoacetate Acetone
also stimulates fat mobilization in adipose tissue, acti-
vating (by cAMP-dependent phosphorylation) both peri-
lipin and triacylglycerol lipase (see Fig. 17–3). Finally,
epinephrine stimulates glucagon secretion and inhibits
insulin secretion, reinforcing its effect of mobilizing
fuels and inhibiting fuel storage.
Cortisol Signals Stress, Including Low Blood Glucose
A variety of stressors (anxiety, fear, pain, hemorrhage,
infections, low blood glucose, starvation) stimulate re-
lease of the corticosteroid hormone cortisol from the
adrenal cortex. Cortisol acts on muscle, liver, and adi-
pose tissue to supply the organism with fuel to with-
stand the stress. Cortisol is a relatively slow-acting hor-
mone that alters metabolism by changing the kinds and
amounts of certain enzymes synthesized in its target
cell, rather than by regulating the activity of existing en-
zyme molecules.
In adipose tissue, cortisol leads to an increase in the
release of fatty acids from stored TAGs. The fatty acids
are exported to serve as fuel for other tissues, and the
glycerol is used for gluconeogenesis in the liver. Corti-
sol stimulates the breakdown of muscle proteins and the
export of amino acids to the liver, where they serve as
precursors for gluconeogenesis. In the liver, cortisol
promotes gluconeogenesis by stimulating synthesis of
the key enzyme PEP carboxykinase (see Fig. 14–17b);
glucagon has the same effect, whereas insulin has the
opposite effect. Glucose produced in this way is stored
in the liver as glycogen or exported immediately to tis-
sues that need glucose for fuel. The net effect of these
metabolic changes is to restore blood glucose to its nor-
mal level and to increase glycogen stores, ready to sup-
port the fight-or-flight response commonly associated
with stress. The effects of cortisol therefore counter-
balance those of insulin.
Diabetes Mellitus Arises from Defects in Insulin
Production or Action
Diabetes mellitus, caused by a deficiency in
the secretion or action of insulin, is a relatively
common disease: nearly 6% of the United States popu-
lation shows some degree of abnormality in glucose
metabolism that is indicative of diabetes or a tendency
toward the condition. There are two major clinical
classes of diabetes mellitus: type I diabetes, or insulin-
dependent diabetes mellitus (IDDM), and type II dia-
betes, or non-insulin-dependent diabetes mellitus
(NIDDM), also called insulin-resistant diabetes.
In type I diabetes, the disease begins early in life and
quickly becomes severe. This disease responds to insulin
injection, because the metabolic defect stems from a
paucity of pancreatic H9252 cells and a consequent inability
to produce sufficient insulin. IDDM requires insulin ther-
apy and careful, lifelong control of the balance between
dietary intake and insulin dose. Characteristic symptoms
of type I (and type II) diabetes are excessive thirst and
frequent urination (polyuria), leading to the intake of
large volumes of water (polydipsia) (“diabetes mellitus”
means “excessive excretion of sweet urine”). These
symptoms are due to the excretion of large amounts of
glucose in the urine, a condition known as glucosuria.
Type II diabetes is slow to develop (typically in
older, obese individuals), and the symptoms are milder
and often go unrecognized at first. This is really a group
of diseases in which the regulatory activity of insulin is
defective: insulin is produced, but some feature of the
insulin-response system is defective. These individuals
are insulin-resistant. The connection between type II di-
abetes and obesity (discussed below) is an active area
of research.
Individuals with either type of diabetes are unable
to take up glucose efficiently from the blood; recall that
insulin triggers the movement of GLUT4 glucose trans-
porters to the plasma membrane of muscle and adipose
tissue (see Fig. 12–8). Another characteristic metabolic
change in diabetes is excessive but incomplete oxida-
tion of fatty acids in the liver. The acetyl-CoA produced
by H9252 oxidation cannot be completely oxidized by the cit-
ric acid cycle, because the high [NADH]/[NAD
H11001
] ratio
produced by H9252 oxidation inhibits the cycle (recall that
three steps convert NAD
H11001
to NADH). Accumulation of
acetyl-CoA leads to overproduction of the ketone bod-
ies acetoacetate and H9252-hydroxybutyrate, which cannot
be used by extrahepatic tissues as fast as they are made
in the liver. In addition to H9252-hydroxybutyrate and ace-
toacetate, the blood of diabetics also contains acetone,
which results from the spontaneous decarboxylation of
acetoacetate:
Acetone is volatile and is exhaled, and in uncon-
trolled diabetes, the breath has a characteristic odor
sometimes mistaken for ethanol. A diabetic individual
who is experiencing mental confusion due to high blood
glucose is occasionally misdiagnosed as intoxicated, an
error that can be fatal. The overproduction of ketone
bodies, called ketosis, results in greatly increased con-
centrations of ketone bodies in the blood (ketonemia)
and urine (ketonuria).
The ketone bodies are carboxylic acids, which ion-
ize, releasing protons. In uncontrolled diabetes this acid
production can overwhelm the capacity of the blood’s
bicarbonate buffering system and produce a lowering of
blood pH called acidosis or, in combination with ketosis,
ketoacidosis, a potentially life-threatening condition.
Biochemical measurements on blood and urine
samples are essential in the diagnosis and treatment of
diabetes. A sensitive diagnostic criterion is provided by
23.3 Hormonal Regulation of Fuel Metabolism 909
8885d_c23_881-919 3/1/04 1:26 PM Page 909 mac76 mac76:385_reb:
the glucose-tolerance test. The patient fasts over-
night, then drinks a test dose of 100 g of glucose dis-
solved in a glass of water. The blood glucose concen-
tration is measured before the test dose and at 30 min
intervals for several hours thereafter. A healthy indi-
vidual assimilates the glucose readily, the blood glucose
rising to no more than about 9 or 10 mM; little or no glu-
cose appears in the urine. Diabetic individuals assimi-
late the test dose of glucose poorly; their blood glucose
level far exceeds the kidney threshold (about 10 mM),
causing glucose to appear in the urine. ■
SUMMARY 23.3 Hormonal Regulation
of Fuel Metabolism
■ The concentration of glucose in blood is
hormonally regulated. Fluctuations in blood
glucose level (normally 60 to 90 mg/100 mL, or
about 4.5 mM) due to dietary intake or vigorous
exercise are counterbalanced by a variety of
hormonally triggered changes in the
metabolism of several organs.
■ High blood glucose elicits the release of insulin,
which speeds the uptake of glucose by tissues
and favors the storage of fuels as glycogen and
triacylglycerols, while inhibiting fatty acid
mobilization in adipose tissue.
■ Low blood glucose triggers release of glucagon,
which stimulates glucose release from liver
glycogen and shifts fuel metabolism in liver and
muscle to fatty acid oxidation, sparing glucose
for use by the brain. In prolonged fasting,
triacylglycerols become the principal fuel; the
liver converts the fatty acids to ketone bodies
for export to other tissues, including the brain.
■ Epinephrine prepares the body for increased
activity by mobilizing blood glucose from
glycogen and other precursors.
■ Cortisol, released in response to a variety of
stressors (including low blood glucose), stimu-
lates gluconeogenesis from amino acids and
glycerol in the liver, thus raising blood glucose
and counterbalancing the effects of insulin.
■ In diabetes, insulin is either not produced or
not recognized by the tissues, and the uptake
of blood glucose is compromised. When blood
glucose levels are high, glucose is excreted.
Tissues then depend on fatty acids for fuel
(producing ketone bodies) and degrade cellular
proteins to provide glucogenic amino acids for
glucose synthesis. Uncontrolled diabetes is
characterized by high glucose levels in the
blood and urine and the production and
excretion of ketone bodies.
23.4 Obesity and the Regulation
of Body Mass
In the United States population, 30% of adults
are obese and another 35% are overweight. (Obe-
sity is defined in terms of body mass index (BMI): BMI
H11005 weight in kg/(height in m)
2
. A BMI below 25 is con-
sidered normal; 25 to 30 is overweight, and greater than
30, obese.) Obesity is life-threatening. It significantly in-
creases the chances of developing type II diabetes as
well as heart attack, stroke, and cancers of the colon,
breast, prostate, and endometrium. Consequently, there
is great interest in understanding how body mass and
the storage of fats in adipose tissue are regulated. ■
To a first approximation, obesity is the result of tak-
ing in more calories in the diet than are expended by
the body’s energy-consuming activities. The body can
deal with an excess of dietary calories in three ways:
(1) convert excess fuel to fat and store it in adipose tis-
sue, (2) burn excess fuel by extra exercise, and (3)
“waste” fuel by diverting it to heat production (ther-
mogenesis) in uncoupled mitochondria. In mammals,
a complex set of hormonal and neuronal signals act to
keep fuel intake and energy expenditure in balance, so
as to hold the amount of adipose tissue at a suitable
level. Dealing effectively with obesity requires under-
standing these various checks and balances under nor-
mal conditions, and how these homeostatic mechanisms
fail in obesity.
The Lipostat Theory Predicts the Feedback
Regulation of Adipose Tissue
The lipostat theory postulates a mechanism that in-
hibits eating behavior and increases energy consump-
tion whenever body weight exceeds a certain value (the
set point); the inhibition is relieved when body weight
drops below the set point (Fig. 23–30). This theory pre-
dicts that a feedback signal originating in adipose tissue
influences the brain centers that control eating behav-
ior and activity (metabolic and motor). The first such
factor, leptin, was discovered in 1994, and several oth-
ers are now known.
Leptin (Greek leptos, “thin”) is a small protein (167
amino acids) that is produced in adipocytes and moves
through the blood to the brain, where it acts on recep-
tors in the hypothalamus to curtail appetite. Leptin was
first identified as the product of a gene designated OB
(obese) in laboratory mice. Mice with two defective
copies of this gene (ob/ob genotype; lowercase letters
signify a mutant form of the gene) show the behavior
and physiology of animals in a constant state of starva-
tion: their serum cortisol levels are elevated; they are
unable to stay warm, they grow abnormally, do not re-
produce, and exhibit unrestrained appetite. As a con-
sequence of the last effect, they become severely obese,
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism910
8885d_c23_881-919 3/1/04 1:26 PM Page 910 mac76 mac76:385_reb:
weighing as much as three times more than normal mice
(Fig. 23–31). They also have metabolic disturbances
very similar to those of diabetic animals, and they are
insulin-resistant. When leptin is injected into ob/ob
mice, they lose weight and increase their locomotor ac-
tivity and thermogenesis.
A second mouse gene, designated DB (diabetic),
has also been found to have a role in appetite regula-
tion. Mice with two defective copies (db/db) are obese
and diabetic. The DB gene encodes the leptin recep-
tor. When the leptin receptor is defective, the signaling
function of leptin is lost.
The leptin receptor is expressed primarily in regions
of the brain known to regulate feeding behavior—
neurons of the arcuate nucleus of the hypothalamus
(Fig. 23–32a). Leptin carries the message that fat re-
serves are sufficient, and it promotes a reduction in fuel
intake and increased expenditure of energy. Leptin-
receptor interaction in the hypothalamus alters the re-
lease of neuronal signals to the region of the brain that
affects appetite. Leptin also stimulates the sympathetic
nervous system, increasing blood pressure, heart rate,
and thermogenesis by uncoupling electron transfer from
ATP synthesis in the mitochondria of adipocytes (Fig.
23–32b). Recall that thermogenin, also called uncou-
pling protein (UCP), forms a channel in the inner mi-
tochondrial membrane that allows protons to reenter
the mitochondrial matrix without passing through the
ATP synthase complex (see Fig. 19–30). This permits
continual oxidation of fuel (fatty acids in an adipocyte)
without ATP synthesis, dissipating energy as heat and
consuming dietary calories or stored fats in potentially
very large amounts.
23.4 Obesity and the Regulation of Body Mass 911
FIGURE 23–30 Set-point model for maintaining constant mass.
When the mass of adipose tissue increases, released leptin inhibits
feeding and fat synthesis and stimulates oxidation of fatty acids. When
the mass of adipose tissue decreases, a lowered leptin production fa-
vors a greater food intake and less fatty acid oxidation.
Food
fat synthesis
fat b oxidation
Energy, heat
Adipose
tissue
FIGURE 23–31 Obesity caused by defective leptin production. Both
these mice, which are the same age, have defects in the OB gene. The
mouse on the right was provided with purified leptin by daily injec-
tion, and weighs 35 g. The mouse on the left got no leptin, conse-
quently ate more food and was less active, and weighs 67 g.
Nucleus
Increased
expression
of UCP gene
TAG
Lipid droplet Perilipin
Adenylyl
cyclase
Adipocyte
cAMP
H
+
G protein
Protein kinase A
Uncoupling
protein (UCP)
Fatty
acids
oxidation
H9252
H9252
Hormone-
sensitive
lipase
Norepinephrine
Heat
3
Adrenergic
receptor
H9252
Hypothalamus
Ventromedial
nucleus
Paraventricular
nucleus
Arcuate
nucleus
Posterior
pituitary
Anterior
pituitary
Leptin
(via blood)
Adipose
tissue
Neuronal signal
via sympathetic
neuron
(a)
(b)
FIGURE 23–32 Hypothalamic regulation of food intake and energy
expenditure. (a) Anatomy of the hypothalamus. (b) Interactions be-
tween the hypothalamus and an adipocyte, described later in text.
8885d_c23_881-919 3/1/04 1:26 PM Page 911 mac76 mac76:385_reb:
Leptin Stimulates Production of Anorexigenic
Peptide Hormones
Two types of neurons in the arcuate nucleus control fuel
intake and metabolism (Fig. 23–33). The orexigenic
(appetite-stimulating) neurons stimulate eating by pro-
ducing and releasing neuropeptide Y (NPY), which
causes the next neuron in the circuit to send the signal
to the brain, Eat! The blood level of NPY rises during
starvation, and is elevated in both ob/ob and db/db mice.
The high NPY concentration presumably underlies the
obesity of these mice, who eat voraciously.
The anorexigenic (appetite-suppressing) neurons
in the arcuate nucleus produce H9251-melanocyte-stimulating
hormone (H9251-MSH), formed from its polypeptide pre-
cursor pro-opiomelanocortin (POMC; Fig. 23–6). Release
of H9251-MSH causes the next neuron in the circuit to send
the signal to the brain, Stop eating!
The amount of leptin released by adipose tissue de-
pends on both the number and the size of adipocytes.
When weight loss decreases the mass of lipid tissue, lep-
tin levels in the blood decrease, the production of NPY
is diminished, and the processes in adipose tissue shown
in Figure 23–32 are reversed. Uncoupling is diminished,
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism912
Ghrelin
Stomach
Colon
Adipose
tissue
Pancreas
Leptin
Insulin
PYY
3-36
Neuron
Muscle, adipose
tissue, liver
Arcuate
nucleus
“Eat more;
metabolize
less”
“Eat less;
metabolize
more”
Anorexigenic
(a-MSH)
PYY
3-36
receptor
Melanocortin
receptor
Melanocortin
receptor
NPY receptor
Leptin receptor
or insulin receptor
Ghrelin receptor
Food
intake
Energy
expenditure
Orexigenic
(NPY)
FIGURE 23–33 Hormones that control eating. In the arcuate nucleus,
two sets of neurosecretory cells receive hormonal input and relay neu-
ronal signals to the cells of muscle, adipose tissue, and liver. Leptin
and insulin are released from adipose tissue and pancreas, respec-
tively, in proportion to the mass of body fat. The two hormones act on
anorexigenic neurosecretory cells (red) to trigger release of H9251-MSH;
this produces neuronal signals to eat less and metabolize more fuel.
Leptin and insulin also act on orexigenic neurosecretory cells (green)
to inhibit the release of NPY, reducing the “eat” signal sent to the tis-
sues. As described later in the text, the gastric hormone ghrelin stim-
ulates appetite by activating the NPY-expressing cells; PYY
3–36
, re-
leased from the colon, inhibits these neurons and thereby decreases
appetite. Each of the two types of neurosecretory cells inhibits hor-
mone production by the other, so any stimulus that activates orexi-
genic cells inactivates anorexigenic cells, and vice versa. This strength-
ens the effect of stimulatory inputs.
8885d_c23_881-919 3/1/04 1:26 PM Page 912 mac76 mac76:385_reb:
slowing thermogenesis and saving fuel, and fat mobi-
lization slows in response to reduced signaling by cAMP.
Consumption of more food combined with more efficient
utilization of fuel results in replenishment of the fat re-
serve in adipose, bringing the system back into balance.
Leptin Triggers a Signaling Cascade That Regulates
Gene Expression
The leptin signal is transduced by a mechanism also
used by receptors for interferon and growth factors, the
JAK-STAT system (Fig. 23–34; see Fig. 12–9). The lep-
tin receptor, which has a single transmembrane seg-
ment, dimerizes when leptin binds to the extracellular
domain of two monomers. Both monomers are phos-
phorylated on a Tyr residue of the intracellular domain
by a Janus kinase (JAK). The P –Tyr residues be-
come docking sites for three proteins that are signal
transducers and activators of transcription (STATs 3,
5, and 6, sometimes called fat-STATS). The docked
STATs are then phosphorylated on Tyr residues by the
same JAK. After phosphorylation, the STATs dimerize
then move to the nucleus, where they bind to specific
DNA sequences and stimulate the expression of target
genes, including the gene for POMC, from which H9251-MSH
is produced.
The increased catabolism and thermogenesis trig-
gered by leptin are due in part to increased synthesis
of the mitochondrial uncoupling protein UCP in
adipocytes. Leptin stimulates the synthesis of this un-
coupling protein by altering synaptic transmissions from
neurons in the arcuate nucleus to adipose and other tis-
sues. In these tissues, leptin causes increased release of
norepinephrine, which acts through H9252
3
-adrenergic re-
ceptors to stimulate transcription of the gene for UCP.
The resulting uncoupling of electron transfer from ox-
idative phosphorylation consumes fat and is thermo-
genic (Fig. 23–32).
Might human obesity be the result of insufficient
leptin production, and therefore treatable by the injec-
tion of leptin? Blood levels of leptin are in fact usually
much higher in obese animals (including humans) than
in animals of normal body mass (except, of course, in
ob/ob animals, which cannot make leptin). Some down-
stream element in the leptin response system must be
defective in obese individuals, and the elevation in lep-
tin is the result of an (unsuccessful) attempt to over-
come the leptin resistance. In those very rare humans
with extreme obesity who have a defective leptin gene
(OB), leptin injection does result in dramatic weight
loss. In the vast majority of obese individuals, however,
the OB gene is intact. In clinical trials, the injection of
leptin did not have the weight-reducing effect observed
in obese ob/ob mice. Clearly, most cases of human obe-
sity involve one or more factors in addition to leptin.
The Leptin System May Have Evolved to Regulate
the Starvation Response
Although much of the initial interest in leptin resulted
from its possible role in preventing obesity, the leptin
system probably evolved to adjust an animal’s activity
and metabolism during periods of fasting and starvation,
not to restrict weight. The reduction in leptin level trig-
gered by nutritional deficiency reverses the thermo-
genic processes illustrated in Figure 23–32, allowing fuel
conservation. Leptin activates AMP-dependent pro-
tein kinase (AMPK), which regulates many aspects
of fuel metabolism. Leptin also triggers decreased pro-
duction of thyroid hormone (slowing basal metabolism),
decreased production of sex hormones (preventing
reproduction), and increased production of glucocorti-
coids (mobilizing the body’s fuel-generating resources).
By minimizing energy expenditures and maximizing
the use of endogenous reserves of energy, these leptin-
mediated responses may allow an animal to survive
periods of severe nutritional deprivation.
23.4 Obesity and the Regulation of Body Mass 913
Cytosol
Plasma membrane
Leptin receptor
monomer
Nucleus
Leptin
Leptin
P
JAK
STAT
P
STAT
P
STAT
P
STAT
DNA
Neuropeptide
(POMC, etc.)
mRNA
P
STAT
STAT
JAK
JAK
JAK
P
FIGURE 23–34 The JAK-STAT mechanism of leptin signal transduc-
tion in the hypothalamus. Leptin binding induces dimerization of the
leptin receptor, followed by phosphorylation of Tyr residues of the
receptor, catalyzed by Janus kinase (JAK). STATs bound to the phos-
phorylated leptin receptor through their SH2 domains are now phos-
phorylated on Tyr residues by a separate activity of JAK. The STATs
dimerize, binding each other’s P –Tyr residues, and enter the nucleus.
Here, they bind specific regulatory regions in the DNA and alter the
expression of certain genes. The products of these genes ultimately in-
fluence the organism’s feeding behavior and energy expenditure.
8885d_c23_881-919 3/1/04 1:26 PM Page 913 mac76 mac76:385_reb:
Insulin Acts in the Arcuate Nucleus to Regulate
Eating and Energy Conservation
Insulin secretion reflects both the size of fat reserves
(adiposity) and the current energy balance (blood glu-
cose level). Insulin acts on insulin receptors in the hy-
pothalamus to inhibit eating (Fig. 23–33). Insulin re-
ceptors in the orexigenic neurons of the arcuate nucleus
inhibit the release of NPY, and insulin receptors in the
anorexigenic neurons stimulate H9251-MSH production,
thereby decreasing fuel intake and increasing thermo-
genesis. By mechanisms discussed in Section 23.3, in-
sulin also signals muscle, liver, and adipose tissues to
increase catabolic reactions, including fat oxidation,
which results in weight loss.
Leptin makes the cells of liver and muscle more sen-
sitive to insulin. One hypothesis to explain this effect
suggests cross-talk between the protein tyrosine kinases
activated by leptin and those activated by insulin (Fig.
23–35); common second messengers in the two signal-
ing pathways allow leptin to trigger some of the same
downstream events that are triggered by insulin,
through insulin receptor substrate-2 (IRS-2) and phos-
phoinositide 3-kinase (PI-3K) (Chapter 12).
Adiponectin Acts through AMPK
Adiponectin is a peptide hormone (224 amino acids)
produced almost exclusively in adipose tissue. It circu-
lates in the blood and powerfully affects the metabolism
of fatty acids and carbohydrates in liver and muscle.
Adiponectin increases the uptake of fatty acids from the
blood by myocytes and the rate at which fatty acids un-
dergo H9252 oxidation in the muscle. It also blocks fatty acid
synthesis and gluconeogenesis in hepatocytes, and it
stimulates glucose uptake and catabolism in muscle and
liver (Fig. 23–36). These effects of adiponectin occur
indirectly, through activation of the key regulatory en-
zyme AMPK by increased cytosolic [AMP]. Increased
[AMP] also results from ATP consumption during in-
tense muscular activity, but it can be brought about by
adiponectin through other, unknown mechanisms.
When activated, AMPK phosphorylates a number of tar-
get proteins critical to the metabolism of fatty acids and
carbohydrates, with profound effects on the metabolism
of the whole animal.
One enzyme regulated by AMPK is acetyl-CoA car-
boxylase, which produces malonyl-CoA, the first inter-
mediate committed to fatty acid synthesis. Malonyl-CoA
is a powerful inhibitor of the enzyme carnitine acyl-
transferase I, which starts the process of H9252 oxidation by
transporting fatty acids into the mitochondrion (see Fig.
17–6). By phosphorylating and inactivating acetyl-CoA
carboxylase, AMPK inhibits fatty acid synthesis while
relieving the inhibition (by malonyl-CoA) of H9252 oxidation
(Fig. 23–37).
Mice with defective adiponectin genes are less sen-
sitive to insulin than those with normal adiponectin, and
they show poor glucose tolerance; ingestion of dietary
carbohydrate causes a long-lasting rise in their blood
glucose. These metabolic defects resemble those of
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism914
Leptin
receptor
Insulin
receptor
Plasma
membrane
Cytosol
P
P
P
P
Leptin
JAKJAK
IRS-2
PI-3K
Inhibition
of food
intake
FIGURE 23–35 A possible mechanism for cross-talk between recep-
tors for insulin and leptin. The insulin receptor has intrinsic Tyr ki-
nase activity (see Fig. 12–6), and the leptin receptor, when occupied
by its ligand, is phosphorylated by a soluble Tyr kinase (JAK). One
possible explanation for the observed interaction between leptin and
insulin is that both may phosphorylate the same substrate—in the case
shown here, insulin receptor substrate-2 (IRS-2). When phosphory-
lated, IRS-2 activates PI-3K, which has downstream consequences that
include inhibition of food intake. IRS-2 serves here as an integrator of
the input from two receptors.
AMPK
Muscle
Fatty acid uptake
Glucose uptake
b Oxidation
Liver
Glycolysis
Fatty acid synthesis
Gluconeogenesis
Adiponectin
FIGURE 23–36 Effects of adiponectin on muscle and adipose tissue.
By interacting with its receptors on the surface of myocytes and he-
patocytes, adiponectin activates their AMPK. The activated kinase
phosphorylates key metabolic enzymes (see Fig. 23–37, for example),
shifting metabolism toward oxidation of fatty acids and away from
lipid and glucose synthesis.
8885d_c23_881-919 3/1/04 1:26 PM Page 914 mac76 mac76:385_reb:
humans with type II diabetes, who also are insulin-
insensitive and clear glucose from the blood only
slowly. Indeed, individuals with obesity or type II dia-
betes have lower blood adiponectin levels than nondia-
betic controls. Moreover, the drugs used in treatment of
type II diabetes—the thiazolidinediones, such as rosigli-
tazone (Avandia) and pioglitazone (Actos) (p. 807)—
increase the expression of adiponectin mRNA in adipose
tissue and increase blood adiponectin levels in experi-
mental animals; they also activate AMPK. It appears that
adiponectin, acting through AMPK, modulates the sen-
sitivity of cells and tissues to insulin. Perhaps this hor-
mone will prove to be one of the links between type II
diabetes and its most important predisposing factor,
obesity.
Three factors improve the health of individuals with
type II diabetes: regular exercise, use of thiazolidine-
diones, and dietary restriction. We have seen that ex-
ercise activates AMPK, as does adiponectin, and that
thiazolidinediones increase the concentration of
adiponectin in plasma, increasing insulin sensitivity. Di-
etary restriction may act by regulating the expression
of genes that encode proteins involved in fatty acid ox-
idation and in energy expenditure via thermogenesis.
Diet Regulates the Expression of Genes Central
to Maintaining Body Mass
Proteins in a family of ligand-activated transcription fac-
tors, the peroxisome proliferator-activated recep-
tors (PPARs), respond to changes in dietary lipid by
altering the expression of genes involved in fat and car-
bohydrate metabolism. These transcription factors
were first recognized for their roles in peroxisome
synthesis—thus their name. Their normal ligands are
fatty acids or fatty acid derivatives, but they can also
bind synthetic agonists and can be activated in the lab-
oratory by genetic manipulation. PPARH9251, PPARH9254, and
PPARH9253 are members of the nuclear receptor superfam-
ily. They act in the nucleus by forming heterodimers
with another nuclear receptor, RXR (retinoid X recep-
tor), binding to regulatory regions of DNA near the
genes under their control and changing the rate of tran-
scription of those genes (Fig. 23–38).
PPARH9253, expressed primarily in liver and adipose
tissue, is involved in turning on genes necessary to the
differentiation of fibroblasts into adipocytes and genes
that encode proteins required for lipid synthesis and
storage in adipocytes. PPARH9253 is activated by drugs of
the thiazolidinedione class, which are used to treat type
II diabetes. PPARH9251 in hepatocytes turns on the genes
necessary for H9252 oxidation of fatty acids and formation
of ketone bodies during fasting.
PPARH9254 is a key regulator of fat oxidation, which
acts by sensing changes in dietary lipid. It acts in liver
and muscle, stimulating the transcription of at least nine
genes encoding proteins for H9252 oxidation and for energy
dissipation through uncoupling of mitochondria. Normal
mice overfed on high-fat diets accumulate massive
amounts of both brown and white fat, and fat droplets
accumulate in the liver. But when the same overfeeding
experiment is done with mice that have a genetically
23.4 Obesity and the Regulation of Body Mass 915
[AMP]
Acetyl-CoA
+ CO
2
Malonyl-CoA
Fatty acid
synthesis
Cytosol
Mitochondrion
Fatty
acid Fatty
acid
Acetyl-CoA
b oxidation
Inactive
Carnitine
acyltransferase I
AMPK
AMPK
ACC ACC
P
FIGURE 23–37 Regulation of fatty acid
synthesis and H9252 oxidation by AMPK
action on acetyl-CoA carboxylase. When
activated by elevated 5H11541-AMP, AMPK
phosphorylates a Thr residue on acetyl-
CoA carboxylase (ACC), inactivating it.
This prevents the synthesis of malonyl-
CoA, the first intermediate in fatty acid
synthesis, and reduction in [malonyl-CoA]
relieves the inhibition of carnitine acyl-
transferase I, allowing fatty acids to enter
the mitochondrial matrix to undergo
H9252 oxidation.
8885d_c23_881-919 3/1/04 1:26 PM Page 915 mac76 mac76:385_reb:
altered, always active PPARH9254, this fat accumulation is
prevented. In mice with a nonfunctioning leptin recep-
tor (db/db), activated PPARH9254 prevents the development
of obesity that would otherwise occur (see Fig. 23–31).
By stimulating fatty acid breakdown in uncoupled mi-
tochondria, PPARH9254 causes fat depletion, weight loss, and
thermogenesis. Seen in this light, thermogenesis is both
a means of keeping warm and a defense against obesity.
Clearly, PPARH9254 is a potential target for drugs to treat
obesity.
Short-Term Eating Behavior Is Set by Ghrelin
and PYY
3–36
Ghrelin is a peptide hormone (28 amino acids) pro-
duced in cells lining the stomach. It was originally rec-
ognized as the stimulus for the release of growth hor-
mone (ghre is the Proto-Indo-European root of “grow”),
then subsequently shown to be a powerful appetite
stimulant that works on a shorter time scale (between
meals) than leptin and insulin. Ghrelin receptors are lo-
cated in the pituitary gland (presumably mediating
growth hormone release) and in the hypothalamus (af-
fecting appetite), as well as in heart muscle and adipose
tissue. The concentration of ghrelin in the blood varies
strikingly between meals, peaking just before a meal and
dropping sharply just after the meal (Fig. 23–39). In-
jection of ghrelin into humans produces immediate sen-
sations of intense hunger. Individuals with Prader-Willi
syndrome, whose blood levels of ghrelin are exception-
ally high, have an uncontrollable appetite, leading to ex-
treme obesity that often results in death before the age
of 30.
PYY
3–36
is a peptide hormone (34 amino acids) se-
creted by endocrine cells in the lining of the small in-
testine and colon in response to food entering from the
stomach. The level of PYY
3–36
in the blood rises after a
meal and remains high for some hours. It is carried in
the blood to the arcuate nucleus, where it acts on orex-
igenic neurons, inhibiting NPY release and reducing
hunger (Fig. 23–33). Humans injected with PYY
3–36
feel
little hunger and eat less than normal amounts for about
12 hours.
This interlocking system of neuroendocrine controls
of food intake and metabolism presumably evolved to
protect against starvation and to eliminate counterpro-
ductive accumulation of fat (extreme obesity). The dif-
ficulty most people face in trying to lose weight testi-
fies to the remarkable effectiveness of these controls.
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism916
Insulin (pmol/L)
Grehlin (pg/mL)
Breakfast
800
700
600
500
400
300
200
600
500
400
300
200
100
0
DinnerLunch
(a)
(b)
6 a.m. 12 noon 6 p.m.
Time of day
12 midnight
Thiazolidinedione
PPARg
PPARg
RXR
Cytosol
Plasma
membrane
Nuclear
envelope
DNA
Regulated gene
Response element
(DNA sequence that
regulates transcription
in response to PPARg
ligand T)
RXR
T
T
T
T
FIGURE 23–38 Mode of action of PPARs. PPARs, when bound to their
cognate ligand, form heterodimers with the nuclear receptor RXR. The
dimer binds specific regions of DNA, response elements, stimulating
transcription of genes in those regions.
FIGURE 23–39 Variations in ghrelin and insulin relative to meal
times. (a) Plasma levels of ghrelin rise sharply just before the normal
time for meals (7 a.m. breakfast, 12 noon lunch, 5:30 p.m. dinner)
and drop precipitously just after meals, paralleling the subjective feel-
ings of hunger. (b) Insulin levels rise immediately after each meal, in
response to the increase in blood glucose concentration.
8885d_c23_881-919 3/1/04 1:26 PM Page 916 mac76 mac76:385_reb:
Chapter 23 Further Reading 917
Key Terms
neuroendocrine
system 882
radioimmunoassay
(RIA) 884
Scatchard analysis 884
endocrine glands 886
paracrine 886
autocrine 886
insulin 887
epinephrine 888
norepinephrine 888
catecholamines 888
eicosanoid
hormones 888
steroid hormones 888
vitamin D hormone 889
retinoid hormones 889
thyroid hormones 889
nitric oxide (NO) 889
NO synthase 889
hypothalamus 889
posterior pituitary 890
anterior pituitary 890
tropic hormone 890
tropin 890
hepatocyte 893
adipocyte 897
myocyte 898
erythrocyte 901
leukocyte 901
lymphocyte 901
platelets 901
blood plasma 901
plasma proteins 901
cortisol 909
diabetes mellitus 909
type I diabetes 909
type II diabetes 909
glucosuria 909
ketosis 909
acidosis 909
ketoacidosis 909
glucose-tolerance
test 910
thermogenesis 910
leptin 910
thermogenin (uncoupling
protein) 911
orexigenic 912
neuropeptide Y
(NPY) 912
anorexigenic 912
H9251-melanocyte-stimulating
hormone (H9251-MSH) 912
Janus kinase (JAK) 913
STAT (signal transducer
and activator of
transcription) 913
thermogenic 913
AMP-dependent protein
kinase (AMPK) 913
adiponectin 914
PPAR (peroxisome
proliferator-activated
receptor) 915
ghrelin 916
PYY
3–36
916
Terms in bold are defined in the glossary.
SUMMARY 23.4 Obesity and the Regulation
of Body Mass
■ Obesity is increasingly common in the
developed countries and predisposes people
toward several life-threatening conditions.
■ Adipose tissue produces leptin, a hormone that
regulates feeding behavior and energy expendi-
ture so as to maintain adequate reserves of fat.
Leptin production and release increase with
the number and size of adipocytes.
■ Leptin acts on receptors in the arcuate nucleus
of the hypothalamus, causing the release of
anorexigenic peptides, including H9251-MSH, that
act in the brain to inhibit eating. Leptin also
stimulates sympathetic nervous system action
on adipocytes, leading to uncoupling of
mitochondrial oxidative phosphorylation, with
consequent thermogenesis.
■ The signal-transduction mechanism for leptin
involves phosphorylation of the JAK-STAT
system. On phosphorylation by JAK, STATs can
bind to regulatory regions in nuclear DNA and
alter the expression of genes for the proteins
that set the level of metabolic activity and
determine feeding behavior. Insulin acts on
receptors in the arcuate nucleus, with results
similar to those caused by leptin.
■ The hormone adiponectin stimulates fatty acid
uptake and oxidation and inhibits fatty acid
synthesis. Its actions are mediated by AMPK.
■ Ghrelin, a hormone produced in the stomach,
acts on orexigenic neurons in the arcuate
nucleus to produce hunger before a meal.
PYY
3–36
, a peptide hormone of the intestine,
acts at the same site to lessen hunger after
a meal.
Further Reading
General Background and History
Crapo, L. (1985) Hormones: The Messengers of Life, W. H.
Freeman and Company, New York.
Short, entertaining account of the history and results of
hormone research.
Litwack, G. & Norman, A.W. (1997) Hormones, 2nd edn,
Academic Press, Orlando, FL.
An excellent, authoritative, well-illustrated, clear introduction
to hormone structure, synthesis, and action.
Yalow, R.S. (1978) Radioimmunoassay: a probe for the fine
structure of biologic systems. Science 200, 1236–1245.
History of the development of radioimmunoassays; the author’s
Nobel lecture.
Hormone Structure and Function
Litwack, G. & Schmidt, R.J. (2001) Biochemistry of hormones I:
polypeptide hormones. In Textbook of Biochemistry with Clinical
Correlations, 5th edn (Devlin, T.M., ed.), pp. 905–958. John Wiley
& Sons, Inc., New York.
8885d_c23_881-919 3/1/04 1:26 PM Page 917 mac76 mac76:385_reb:
Chapter 23 Hormonal Regulation and Integration of Mammalian Metabolism918
1. ATP and Phosphocreatine as Sources of Energy for
Muscle During muscle contraction, the concentration of
phosphocreatine in skeletal muscle drops while the concen-
tration of ATP remains fairly constant. However, in a classic
experiment, Robert Davies found that if he first treated mus-
cle with 1-fluoro-2,4-dinitrobenzene (p. 97), the concentra-
tion of ATP declined rapidly while the concentration of
phosphocreatine remained unchanged during a series of con-
tractions. Suggest an explanation.
2. Metabolism of Glutamate in the Brain Brain tissue
takes up glutamate from the blood, transforms it into gluta-
mine, then releases it into the blood. What is accomplished
by this metabolic conversion? How does it take place? The
amount of glutamine produced in the brain can actually ex-
ceed the amount of glutamate entering from the blood. How
does this extra glutamine arise? (Hint: You may want to re-
view amino acid catabolism in Chapter 18; recall that NH
H11001
4
is
very toxic to the brain.)
Problems
Litwack, G. & Schmidt, R.J. (2001) Biochemistry of hormones
II: steroid hormones. In Textbook of Biochemistry with Clinical
Correlations, 5th edn (Devlin, T.M., ed.), pp. 959–989. John Wiley
& Sons, Inc., New York.
Tissue-Specific Metabolism
Arias, I.M., Boyer, J.L., Chisari, F.V., Fausto, N., Schachter,
D., & Shafritz, D.A. (2001) The Liver: Biology and Pathobiol-
ogy, 4th edn, Lippincott, Williams & Wilkins, Philadelphia.
Advanced text; includes chapters on the metabolism of
carbohydrates, fats, and proteins in the liver.
Nosadini, R., Avogaro, A., Doria, A., Fioretto, P., Trevisan,
R., & Morocutti, A. (1989) Ketone body metabolism: a physiolog-
ical and clinical overview. Diabetes Metab. Rev. 5, 299–319.
Randle, P.J. (1995) Metabolic fuel selection: general integration at
the whole-body level. Proc. Nutr. Soc. 54, 317–327.
Hormonal Regulation of Fuel Metabolism
Attie, A.D. & Raines, R.T. (1995) Analysis of receptor-ligand
interactions. J. Chem. Ed. 72, 119–123.
Carling, D. (2004) The AMP-activated protein kinase cascade—
a unifying system for energy control. Trends Biochem. Sci. 29,
[in press].
Elia, M. (1995) General integration and regulation of metabolism
at the organ level. Proc. Nutr. Soc. 54, 213–234.
Holloszy, J.O. & Kohrt, W.M. (1996) Regulation of carbohydrate
and fat metabolism during and after exercise. Annu. Rev. Nutr.
16, 121–138.
Pilkis, S.J. & Claus, T.H. (1991) Hepatic gluconeogenesis/
glycolysis: regulation and structure/function relationships of
substrate cycle enzymes. Annu. Rev. Nutr. 11, 465–515.
Advanced review.
Snyder, S.H. (1985) The molecular basis of communication
between cells. Sci. Am. 253 (October), 132–141.
Introductory discussion of the human endocrine system.
Zammit, V.A. (1996) Role of insulin in hepatic fatty acid
partitioning: emerging concepts. Biochem. J. 314, 1–14.
Control of Body Mass
Auwerx, J. & Staels, B. (1998) Leptin. Lancet 351, 737–742.
Brief overview of the leptin system and JAK-STAT signal
transductions.
Elmquist, J.K., Maratos-Flier, E., Saper, C.B., & Flier, J.S.
(1998) Unraveling the central nervous system pathways underlying
responses to leptin. Nat. Neurosci. 1, 445–450.
Short, excellent review.
Flier, J.S. & Maratos-Flier, E. (1998) Obesity and the
hypothalamus: novel peptides for new pathways. Cell 92, 437–440.
Freake, H.C. (1998) Uncoupling proteins: beyond brown adipose
tissue. Nutr. Rev. 56, 185–189.
Review of the structure, function, and role of uncoupling
proteins.
Friedman, J.M. (2002) The function of leptin in nutrition, weight,
and physiology. Nutr. Rev. 60, S1–S14.
Intermediate-level review of all aspects of leptin biology.
Inui, A. (1999) Feeding and body-weight regulation by
hypothalamic neuropeptides—mediation of the actions of leptin.
Trends Neurosci. 22, 62–67.
Short, intermediate-level review of the leptin system.
Jequier, E. & Tappy, L. (1999) Regulation of body weight in
humans. Physiol. Rev. 79, 451–480.
Detailed review of the role of leptin in body-weight regulation,
the control of food intake, and the roles of white and brown
adipose tissues in energy expenditure.
Korner, J. & Aronne, L.J. (2003) The emerging science of body
weight regulation and its impact on obesity treatment. J. Clin.
Invest. 111, 565–570.
Marx, J. (2003) Cellular warriors at the battle of the bulge.
Science 299, 846–849.
Short review of biochemistry of weight control, and
introduction to several papers in the same issue of Science on
obesity in humans.
Unger, R.H. (2003) The physiology of cellular liporegulation.
Annu. Rev. Physiol. 65, 333–347.
Advanced review.
Wang, Y.-X., Lee, C.-H., Tiep, S., Yu, R.T., Ham, J., Kang, H.,
& Evans, R.M. (2003) Peroxisome-proliferator-activated receptor
activates fat metabolism to prevent obesity. Cell 113, 159–170.
Describes the role of PPARH9254.
Wood, S.C., Seeley, R.J., Porte, D., Jr., & Schwartz, M.W.
(1998) Signals that regulate food intake and energy homeostasis.
Science 280, 1378–1383.
Review of the roles of leptin, insulin, and other neuropeptides
in the regulation of feeding and catabolism.
8885d_c23_881-919 3/1/04 1:26 PM Page 918 mac76 mac76:385_reb:
Chapter 23 Problems 919
3. Absence of Glycerol Kinase in Adipose Tissue
Glycerol 3-phosphate is required for the biosynthesis of tria-
cylglycerols. Adipocytes, specialized for the synthesis and degra-
dation of triacylglycerols, cannot use glycerol directly, be-
cause they lack glycerol kinase, which catalyzes the reaction
Glycerol H11001 ATP 88n glycerol 3-phosphate H11001 ADP
How does adipose tissue obtain the glycerol 3-phosphate
necessary for triacylglycerol synthesis?
4. Oxygen Consumption during Exercise A sedentary
adult consumes about 0.05 L of O
2
in 10 seconds. A sprinter,
running a 100 m race, consumes about 1 L of O
2
in 10 sec-
onds. After finishing the race, the sprinter continues to
breathe at an elevated (but declining) rate for some minutes,
consuming an extra 4 L of O
2
above the amount consumed
by the sedentary individual.
(a) Why does the need for O
2
increase dramatically dur-
ing the sprint?
(b) Why does the demand for O
2
remain high after the
sprint is completed?
5. Thiamine Deficiency and Brain Function Individu-
als with thiamine deficiency show some characteristic neu-
rological signs and symptoms, including loss of reflexes, anx-
iety, and mental confusion. Why might thiamine deficiency be
manifested by changes in brain function?
6. Potency of Hormones Under normal conditions, the
human adrenal medulla secretes epinephrine (C
9
H
13
NO
3
) at
a rate sufficient to maintain a concentration of 10
H1100210
M in cir-
culating blood. To appreciate what that concentration means,
calculate the diameter of a round swimming pool, with a wa-
ter depth of 2.0 m, that would be needed to dissolve 1.0 g
(about 1 teaspoon) of epinephrine to a concentration equal
to that in blood.
7. Regulation of Hormone Levels in the Blood The
half-life of most hormones in the blood is relatively short. For
example, when radioactively labeled insulin is injected into
an animal, half of the labeled hormone disappears from the
blood within 30 min.
(a) What is the importance of the relatively rapid inac-
tivation of circulating hormones?
(b) In view of this rapid inactivation, how is the level of
circulating hormone kept constant under normal conditions?
(c) In what ways can the organism make rapid changes
in the level of a circulating hormone?
8. Water-Soluble versus Lipid-Soluble Hormones On
the basis of their physical properties, hormones fall into one
of two categories: those that are very soluble in water but rel-
atively insoluble in lipids (e.g., epinephrine) and those that
are relatively insoluble in water but highly soluble in lipids
(e.g., steroid hormones). In their role as regulators of cellu-
lar activity, most water-soluble hormones do not enter their
target cells. The lipid-soluble hormones, by contrast, do en-
ter their target cells and ultimately act in the nucleus. What
is the correlation between solubility, the location of recep-
tors, and the mode of action of these two classes of hormones?
9. Metabolic Differences between Muscle and Liver
in a “Fight or Flight” Situation During a “fight or flight”
situation, the release of epinephrine promotes glycogen
breakdown in the liver, heart, and skeletal muscle. The end
product of glycogen breakdown in the liver is glucose; the
end product in skeletal muscle is pyruvate.
(a) What is the reason for the different products of
glycogen breakdown in the two tissues?
(b) What is the advantage to an organism that must fight
or flee of these specific glycogen breakdown routes?
10. Excessive Amounts of Insulin Secretion: Hyperin-
sulinism Certain malignant tumors of the pancreas cause
excessive production of insulin by the H9252 cells. Affected indi-
viduals exhibit shaking and trembling, weakness and fatigue,
sweating, and hunger.
(a) What is the effect of hyperinsulinism on the metab-
olism of carbohydrates, amino acids, and lipids by the liver?
(b) What are the causes of the observed symptoms?
Suggest why this condition, if prolonged, leads to brain
damage.
11. Thermogenesis Caused by Thyroid Hormones
Thyroid hormones are intimately involved in regulating the
basal metabolic rate. Liver tissue of animals given excess thy-
roxine shows an increased rate of O
2
consumption and in-
creased heat output (thermogenesis), but the ATP concen-
tration in the tissue is normal. Different explanations have
been offered for the thermogenic effect of thyroxine. One is
that excess thryroxine causes uncoupling of oxidative phos-
phorylation in mitochondria. How could such an effect ac-
count for the observations? Another explanation suggests
that the thermogenesis is due to an increased rate of ATP uti-
lization by the thyroxine-stimulated tissue. Is this a reason-
able explanation? Why?
12. Function of Prohormones What are the possible ad-
vantages in the synthesis of hormones as prohormones?
13. Sources of Glucose during Starvation The typical
human adult uses about 160 g of glucose per day, 120 g of
which is used by the brain. The available reserve of glucose
(~20 g of circulating glucose and ~190 g of glycogen) is ad-
equate for about one day. After the reserve has been depleted
during starvation, how would the body obtain more glucose?
14. Parabiotic ob/ob mice By careful surgery, researchers
can connect the circulatory systems of two mice so that the
same blood circulates through both animals. In these parabi-
otic mice, products released into the blood by one animal reach
the other animal via the shared circulation. Both animals are
free to eat independently. If an ob/ob mouse (both copies of
the OB gene are defective) and a normal OB/OB mouse (two
good copies of the OB gene) were made parabiotic, what would
happen to the weight of each mouse?
8885d_c23_881-919 3/1/04 1:26 PM Page 919 mac76 mac76:385_reb:
The bacterial RNA polymerase
8885d_c24_920 2/11/04 3:10 PM Page 920 mac76 mac76:385_reb: