Understanding The Control of Metabolism
Understanding The Control of Metabolism
net/publication/317369476
CITATIONS READS
804 11,634
1 author:
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by David Andrew Fell on 06 June 2017.
David Fell
School of Biological & Molecular Sciences,
Oxford Brookes University
Preface v
i
Contents
ii
Contents
iii
Contents
8 Conclusion 261
8.1 A new view of the physiological control of flux . . . . . . . . . 263
8.1.1 Evidence for multisite modulation in cells . . . . . . . . 265
8.1.2 Implications of multisite modulation . . . . . . . . . . . 274
8.2 Limitations of Control Analysis . . . . . . . . . . . . . . . . . . 274
8.2.1 Objections to Control Analysis . . . . . . . . . . . . . . 275
8.2.2 Extensions of Control Analysis . . . . . . . . . . . . . . 277
8.3 Applications of Metabolic Control Analysis . . . . . . . . . . . 280
8.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
References 285
iv
Preface
There is little doubt that the central areas of traditional biochemistry (such as
metabolism, metabolic regulation and enzymology) are currently unfashion-
able and generally regarded as problems long solved. They have been eclipsed
by molecular genetics, which is not only able to identify the genes for protein
components and their regulatory elements, but which also promises the means
to change the genes and their control elements at will, to the point where it
could truly be called biological engineering. In the press and on television,
there are already debates about how we should use, and not abuse, the power
we are expected to posess tomorrow to manipulate even complex multigenic,
and environmentally–influenced, characteristics such as human intelligence
and resistance to disease. How simple it should be, then, to engineer a trait
such as metabolism, where there are such clear and well–understood links
from the genes to the enzymes to the metabolic pathways.
Why then, have I written a book now about metabolism and its regula-
tion which returns to these central, unfashionable concerns of biochemistry
and which only mentions molecular genetics in a subordinate role? The an-
swer is that even 20 years ago, a small number of researchers were convinced
that there were fundamental flaws in the explanations that biochemists gave
for the regulation and control of metabolism. Together, these critics created
Metabolic Control Analysis as an alternative means to understand and ex-
plain these problems, and they have gradually been persuading more and
more biological scientists to design and interpret their experiments in a differ-
ent way. However, their criticisms and reworkings of traditional biochemical
explanations are still not widely known, and have as yet had little influence on
the contents of the standard biochemistry textbooks, which still cling to the
rejected concepts, usually without even mentioning the doubts and problems.
The purpose of my book is therefore to present an introduction to Metabolic
Control Analysis that places it in the context of the biochemical results and
phenomena to which it must be applied. My hope is that this will provide a
more accessible account of the major features of Metabolic Control Analysis
than has previously been available, since most of the existing literature con-
sists of reviews and original articles in research journals, or brief outlines of a
few selected aspects.
My full case for why Metabolic Control Analysis deserves to be more widely
known is contained in the rest of the book, but a representative illustration
v
Preface
comes from genetic engineering itself. There have already been attempts to
manipulate metabolism, and these have involved successfully changing the
amounts or kinetic properties of certain target enzymes. These enzymes have
been selected because traditional biochemical theory, and the textbooks, iden-
tify them as the key steps in the control of a metabolic process. Virtually none
of these experiments has had the expected outcome: the technology exists,
but traditional metabolic biochemistry does not provide the understanding
needed to use it purposefully. Nor, it should be said, is Metabolic Control
Analysis certain to be able to predict the outcome, though it can provide
a framework in which the possibility of achieving a specific modification of
metabolism by a particular genetic intervention can be assessed.
One factor that has undeniably worked against the wide acceptance of
Metabolic Control Analysis is that it inevitably involves more mathematical
concepts and numerical computation than is usual in metabolic biochem-
istry. This is inevitable because it is the essence of the Control Analysis case
against the traditional concepts that they cannot be rigorously justified and
are not adapted to the quantitative testing that would be necessary to discrim-
inate between alternative explanations. Nevertheless, I have tried to limit the
amount of mathematics that is used in this book, and to explain the meaning
and interpretation of the essential equations. Sometimes I have placed addi-
tional mathematical material in boxes or appendices if it is not essential to
the main lines of the argument. I have not, as a result, presented Metabolic
Control Analysis in its fullest or most mathematically rigorous form, nor have
I provided derivations of all its theorems; these can be found elsewhere. How-
ever, it would be a delusion for metabolic biochemists to imagine that the
detailed behaviour of systems as complex as metabolic pathways can be pre-
dicted with the aid of a few qualitative, verbal principles. Other biological
scientists studying comparably complex phenomena, such as population dy-
namics and genetics, or physiology, have long known this, and biochemistry
is remarkable for its limited use of mathematical and theoretical analysis.
The ideas I present in this book draw on the work of a great many re-
searchers and are only partly my own. Even the ideas I would claim as my
own owe much to the influences of other scientists and colleagues, some of
whom I wish specifically to identify.
Eric Newsholme, first stimulated my interest in the control of metabolism,
with his excellent lectures and tutorials at the University of Oxford, even
though I have not followed his approach nor, in the end, agreed with his
conclusions.
Henrik Kacser’s article The control of flux 119 (with Jim Burns) made me
realize that a more rational approach to metabolic regulation was possible.
Later, when I first stated to publish my results in Metabolic Control Analysis,
Henrik’s encouragement gave me confidence to continue. His death in March
1995 whilst still actively involved in research on Metabolic Control Analysis,
even though officially in retirement, was a great loss to the field of Metabolic
Control Analysis of which he was a cofounder.
Herbert Sauro’s unsuppressible curiosity when he was my research student
vi
Preface
led us to make our first discoveries in the theory of Metabolic Control Analysis.
He also then began the development of the metabolic computer simulation
package SCAMP, which has continued to be a useful tool in my research
group ever since and which has contributed some of the simulations reported
in this book.
My next research student, Rankin Small, helped continue the momentum
of the work begun with Herbert and began the development of the Control
Analysis of covalent modification systems described in Chapter 7. Some of
his subsequent work with Henrik Kacser influenced the ideas I put forward in
the final chapter.
More recently, Simon Thomas has worked with me, first as a research
student and then as a Leverhulme Special Research Fellow, on some of the
ideas presented in the final two chapters and he drew my attention to some
of the examples I have used there.
The elasticity calculus presented in the Appendix to Chapter 5 has been
developed further and applied in computer programs by Simon Thomas, John
Woods and Herbert Sauro.
A number of colleagues helped me by supplying various materials for the
book. Kevin Brindle provided me with an original copy of Fig. 2.3, which
appeared first as Fig. 1 in Biochim. Biophys. Acta, 847, 285–292 (1985) and
which is reproduced with permission of the publishers. Jean–Pierre Mazat
and Thierry Letellier provided me with the data and fitting functions for their
inhibitor titrations in Fig. 6.16, but also, more importantly, much–appreciated
hospitality in Bordeaux on several occasions for discussions about Control
Analysis. Christoph Giersch sent me a manuscript before publication and the
data that appears in Figs. 6.18 and 6.19; we also had useful discussions about
the related text.
I am grateful to all those who have commented on various parts of the
book. Keith Snell, as editor of this series, read the whole text and made
useful suggestions about presentation. Athel Cornish–Bowden corrected my
errors in Chapter 3 on enzyme kinetics and also had helpful ideas about the
presentation. Simon Thomas read much of the draft and helped to improve
the text in several places. Others who have commented on the text include
Jofao Pedro Moniz Barreto, Stefan Schuster, John Woods and my wife, Mary.
Many other friends and colleagues have made an indirect contribution to
this book through conversations, discussions and even critical scrutiny of my
ideas, particularly at conferences. I cannot list them all, but three who have
had a direct influence on material in the book are Athel Cornish–Bowden,
Jannie Hofmeyr and Mark Stitt. Athel and Jannie have provided me with
advance copies of several of their papers that relate particularly to material
in the first chapter and the last two. Mark drew my attention to some of the
illustrative material I have used from plant biochemistry.
I am also grateful to the undergraduates of Oxford Brookes University who
were the unwitting guinea pigs whilst I rehearsed the presentation of parts of
this book.
In addition to those mentioned previously, the following authors and pub-
vii
Preface
lishers have kindly given me permission to use their diagrams in this book:
• J. Katz et al. , J. Biol. Chem. 253, 4530–4536, (1977) for Fig. 2.4;
David Fell
Oxford
January 1996
viii
1
1
2 1 Introduction: regulation and control
• Aiding the search for drugs that affect metabolism by identifying ap-
propriate target sites for bringing about the required change.
in terms of the performance of the metabolic system rather than the existence
of particular mechanisms in the system. They stated that the effectiveness
of metabolic regulation should be assessed from the way a metabolic system
responds to environmental changes. For example, metabolic regulation could
be a sensitive response of the rate of a metabolic pathway to environmental
conditions whilst metabolite concentrations are kept near constant. In other
circumstances, the notable feature of the performance might be the mainte-
nance of near-constant metabolic rates (for example, of ATP synthesis) in the
face of fluctuations in the environment. This underlines the problem that it
might not be completely clear what aspect of a metabolic system is being
regulated; this might be assumed or presumed, whereas in a technological
system, this aspect of function would be an explicit feature of the design.
– Measuring metabolites
– Measuring metabolic fluxes
– Measuring enzyme contents
(Note that I have used metabolic fluxes rather than metabolic rates. Although
the two expressions mean essentially the same, it is conventional in Metabolic
Control Analysis to use flux to refer to a rate in a multicomponent system
such as a metabolic pathway and to reserve rate for individual component
steps, such as enzymes. I will maintain this distinction in the rest of this
book.)
At the molecular level, the properties of individual molecular components
of metabolic systems are characterized. Examples from later in the book
include:
• Enzyme kinetics.
• Feedback inhibition.
Researchers in the areas I have just crudely allocated into these two groups
may resent my implication that, by studying a topic from one of them, they
are distanced from biochemistry at the other level. They might point out
that system–level studies have established that particular pathways have some
means of responding to certain physiological or environmental stimuli, and
that molecular–level studies have identified molecules that relay the signals
and the enzymes on which they act. However, I still believe that to build
a bridge between these two levels, to ensure that the knowledge and under-
standing we have at each can be combined as a consistent whole, is extremely
difficult and has not received sufficient attention until recently. Even though
the discovery of control molecules and signals has required work at the two
levels (e.g. Chapter 7.4), I do not think I am proved wrong because there have
been many unresolved disagreements about the physiological significance and
role of many of the molecular events that have been discovered to have the
potential for regulation and control. The relative lack of attention given to
relating the two levels, and to developing an adequate theory able to resolve
disputes, probably reflects the optimistic tendency in science to exaggerate
the extent of our islands of knowledge and to stress the importance of work at
their margins, whilst minimizing the significance of the relatively uncharted
seas of ignorance that separate them. To explain how this applies to metabolic
regulation, I will consider in the following sections three types of approach
to explanation of metabolic control and regulation. Although the aim of the
book is to promote the advantages of the third type of approach, examples
of the other two will be found, sometimes to show how current thinking has
evolved from previous ideas, sometimes to criticize ideas that are still current,
and sometimes because the third type of approach is incomplete and does not
yet have a view to offer.
nately, most of these interactions do not occur, and the most important ones
are obvious: first of all, an enzyme must interact with its substrates and
products, and then perhaps with certain other metabolites from its own path-
way. Even so, for an enzyme such as glutamine synthase, which has three
substrates, three products and nine significant effectors, the problem of ob-
taining an accurate description of its behaviour for all concentrations of these
fifteen metabolites is huge. Michael Savageau217 pointed out that it would be
necessary to make 415 or ≈ 109 measurements in order to investigate all the
possible interactions, even if only four different concentrations of each of the
fifteen metabolites would be needed to characterize its response. Obviously,
it is unlikely that complete descriptions of complex enzymes will ever exist,
though the principal characteristics of many of the common enzymes, such
as those in glycolysis and the tricarboxylic acid cycle, are known for several
different cell types. Indeed, there have been attempts to build this infor-
mation into computer models to simulate metabolism, particularly by David
Garfinkel and his group working on the carbohydrate catabolism of mam-
malian heart muscle,78 and by Barbara Wright and her co–workers on the
carbohydrate metabolism of slime moulds2, 278 (Chapter 6.2). Whilst these
models have had some success, in order to make their behaviour realistic it
is usually necessary to change some of the values for some of the enzymes’
properties, since the experimental values seem unsuitable, often for no very
clear reason.279
In terms of my previous metaphor, reductionists are not worried about a
sea of ignorance between the molecular level and system level of explanation,
because they believe that in principle it will eventually be crossed with the
boat of mathematical modelling and computer simulation, even though the
record of successful crossings is not very good. It seems we are still a long way
from being able to simulate metabolism easily and routinely on the basis of
our present biochemical knowledge. If we cannot do this, can we be justified in
claiming that our knowledge of the molecular properties explains the system–
level behaviour in biochemistry? Up to a point we probably can, because
the types of behaviour observed with real metabolic pathways are seen in
computer simulations of metabolic systems composed of model enzymes with
properties typical of real enzymes, even though we cannot easily reproduce
the exact behaviour of a specific metabolic pathway.
the world map, so that we have a good impression of the global geography
even though we accept that there is a lack of very specific detail.
Mechanisms for the first of these are known in great molecular detail in a
number of cases, beginning with the work of Jacob and Monod on the control
of the lac operon of E. coli. Much less is known about the mechanisms of the
second, though it is known that in mammals, for example, some enzymes have
lifetimes of only a few minutes, whilst other enzymes in the same cells can
last for days. Proteolytic breakdown of enzymes can be selectively targeted,
for example with the ubiquitin system,102 and can be initiated in response to
specific stimuli. The dynamic range of these mechanisms can be very large;
that is, the amount of enzyme protein in a cell can be varied between nothing
and the level of the major metabolic enzymes such as those of glycolysis. This
book will not deal with the molecular details of these mechanisms, examples
1.3 An overview of mechanisms of regulation and control 9
The Nobel Laureates E G Krebs and E Fischer first demonstrated this in mam-
malian muscle for the activation of glycogen phosphorylase by phosphoryla-
tion (Chapter 7.4.1); since then, many eukaryotic enzymes have been found
to be affected by phosphorylation and dephosphorylation, and more recently
the process has been discovered in bacteria. There are also other types of
covalent modification apart from phosphorylation (see Chapter 7.4.2, p. 231).
In animals, the modification reactions may respond to hormonal or nervous
signals, and are in fact the means whereby these signals can have relatively
rapid effects on metabolism. The potential degree of change from the lowest
enzymic activity to the highest (the dynamic range) is very large with this
mechanism. This is because the inactive form of the enzyme can be almost
completely devoid of activity, and virtually all the enzyme can potentially be
converted to this state; on the other hand, when activation occurs, there can
be essentially complete conversion of all the enzyme present in the cell to the
active form. Examples of cyclic covalent modification schemes are considered
later in this book (Chapter 7.4.3 & 7.4.4), together with an examination of
the types of behaviour they can exhibit (Chapter 7.4.5).
• Short time scales i.e. seconds or less. The mechanisms on this time
scale are presumed to be the reversible binding of metabolites to en-
zymes, causing effects such as allosteric inhibition and activation (Chap-
ter 3.5, p. 71).
The forces holding metabolites to enzymes are usually non–covalent and there-
fore relatively weak, but as the links can easily break and reform, the equilibria
are established extremely rapidly. On the other hand, the dynamic range of
these effects on enzyme activity is limited, since inhibition and activation
involving non–covalent binding are rarely 100% effective. The maximum acti-
vation is also limited by the amount of enzyme available, but it is not possible
to have a large reserve of enzyme to compensate for this partial activation
because then the inhibition effects would not be strong enough to slow the
enzyme when it is not needed. These effects are considered in several places
in this book; in the chapter on enzyme kinetics (Chapter 3.5) the molecular
mechanisms of the effects are considered, whereas in Chapter 7.2 examples
are given of the ways in which they implement feedback control systems.
10 1 Introduction: regulation and control
1- 2- 3- -
Glucose Glc6P Fru6P Ethanol + CO2
tion (Fig. 1.1). This also requires that the flow through the pathway remains
constant for the reasons explained in the figure legend. It is not inevitable
that any sequence of successive enzymic reactions will always be able to reach
a steady state in these circumstances, but the consequence of not reaching
a steady state would be that one or more intermediate metabolites would
continue to accumulate in ever increasing amounts. Such an increase would
start to give a cell severe osmotic problems, and would be at odds with the
general impression of dynamic equilibrium. Indeed, biochemists must have an
expectation that a metabolic system tends to a reproducible state, since sim-
ilar experimental results could only be obtained from day to day in this case.
It would not generally be possible with non–steady states, since the amounts
of metabolites would depend very sensitively on the previous history of the
sample to such an extent that it would be difficult to repeat any observation.
(An exception to this would be the metabolic oscillations that are sometimes
observed reproducibly; however, although these show regular fluctuations in
concentrations and fluxes, their average values are constant.)
Nevertheless, an exact steady state is a mathematical abstraction. One
reason for this is that as a pathway comes closer and closer to steady state, its
rate of approach becomes ever slower, so that in theory it will only arrive there
after an infinite amount of time. In practice this does not matter, because the
limited accuracy of methods of biochemical measurement mean that a system
that is more than 95% of the way to steady state is virtually indistinguishable
from the steady state itself, and 95% attainment of the steady state could
well occur in the preincubation phase of a metabolic experiment.
Even so, there is another limitation to the concept of a steady state: the
environment, whether inside some laboratory glassware or in nature itself,
is never completely constant, not least because of the action of metabolism.
It is inevitable that the source materials of metabolic pathways will slowly
deplete, and that excretory products will accumulate. Also, as mentioned in
the previous section, there are other events that come into effect on longer
time scales, such as induction or repression of enzyme synthesis. Another
possibility is that minor metabolic pathways, such as those for the synthesis
of adenine or nicotinamide nucleotides, cause slow changes in the amounts
of coenzymes like ATP and NAD+ . If there are very slow changes in the
concentrations of metabolites or the pathway flux caused by any of these
factors, the pathway may still be regarded as being in quasi steady state
provided the time scale of the changes is very much longer than the time
taken by the pathway to approach steady state. In other words, if there
are slow changes in one or more metabolites, this must mean that the rates
of synthesis and degradation are not quite in balance, but if the degree of
imbalance is only a small fraction of the rate of synthesis, then it may still be
a good approximation to regard it as being at steady state. Fig. 1.2 illustrates
the way in which the quasi steady state is relative to the time scale of the
experimental observations. In the chapter on enzyme kinetics (Chapter 3),
it is shown that exactly the same concept of quasi steady state is routinely
accepted as a basis for interpreting experiments.
12 1 Introduction: regulation and control
.....Enzyme
.........................synthesis
............................
A B C D
..................................................................
....................... . . . . . . . . . ........................
... . . . . . . . ..... . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. .
. . . Fast processes. . . . . . . . Slow processes . . . .
... ..... Processes ...
. . . at dynamic . . . . . under study . . . . . .ignored ....
. . . equilibrium . . . . . . . . . . . . . . . . . . . . . .. . ..
....................... . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . .. . ..
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Fig. 1.2: Metabolic time scales and the quasi steady state.
For different types of experiment, the experimental viewing frame is placed over the
time axis in different positions, masking off the regions that are not being observed. For
example, a fast reaction enzyme kineticist would centre the window of the frame above
the letter A, a metabolic biochemist working on isolated cells or tissues might center
it in the range B to C, and a biochemist working on whole animals, or a nutritionist,
might centre it above D. The time scales of various biochemical events are indicated as
very approximate ranges.
1.4 Basic theory of metabolism 13
going at the same rate in a steady state, there is no requirement that they
all have the same free energy change, since this has no direct link with the
kinetics. In fact, it is a matter of observation (see Chapter 4.2, p. 88) that
the degrees of displacement from equilibrium of the reactions of a pathway
are extremely variable. Before we consider the significance of this, let us first
define the degree of displacement more precisely.
Consider the reaction:
A + B*
)C + D
The actual free energy of this reaction for a given set of concentrations of the
substrates and products is ∆G0 (where the prime indicates the (biological)
standard condition of pH 7), related to the standard free energy ∆G◦0 by the
expression:
CD
∆G0 = ∆G◦0 + RT ln (1.1)
AB
where R is the gas constant, T the absolute temperature and ln is the natural
logarithm. The ratio of product concentrations to substrate concentrations is
defined as the mass action ratio, Γ, i.e. :
CD
Γ= (1.2)
AB
0 = ∆G◦0 + RT ln Keq
or
∆G◦0 = −RT ln Keq (1.3)
kept within reasonable bounds is for reactions with large negative standard
free energy changes to be far from equilibrium, and for reactions with small
standard free energy changes to be near to equilibrium. In addition, as Atkin-
son7 pointed out, chemical energy can only be conserved by maintaining a
reaction such as ATP hydrolysis far from equilibrium; free energy coupling in
metabolism requires the existence of non–equilibrium reactions.
If these thermodynamic considerations are now combined with chemical
kinetics, it is possible to show why non–equilibrium reactions are more likely
than near–equilibrium reactions to have a determining influence on the rate of
a metabolic pathway, using arguments developed by a number of biochemists
in the 1960s.
Consider first the reversible reaction:
f v
A −→
←− B
vr
kf A = k r B
Alternatively,
vp
=1−ρ
vf
With vp > 0, this shows that the reaction can only be near–equilibrium (i.e.
with ρ ' 1) if vf À vp so that vf ≈ vr . A non–equilibrium reaction has
a small ρ, and therefore the pathway rate, vp is comparable in size to the
forward reaction rate vf , and the rate of the reverse reaction, vr , is relatively
very small. (Although these relationships between ρ and the ratio of the
reverse and forward reaction rates have been derived here using first–order
chemical kinetics, the same expression is obtained for the ratio of the reaction
rates if proper enzyme kinetic equations are used.201 )
Such calculations convinced biochemists that a near–equilibrium reaction
is catalyzed by an enzyme that is present in a great excess over the amount
required to deliver the overall flux of the pathway, and that therefore it is
not limiting the rate of the pathway. The non–equilibrium reaction, on the
other hand, is a potentially limiting factor on the overall pathway flux, be-
cause the catalyzed reaction rate is only slightly greater. Furthermore, many
biochemists gained the impression from these equations for the ratios of rates
that the reverse and net rates of a non–equilibrium reaction would not be
significantly affected by changes in the product concentration. Unfortunately
this is not true for enzyme–catalyzed reactions because the values of the net
and reverse rates do in general depend on product concentration, though as
the proportional effect is the same on both, the ratio does not show this. As
a result, product inhibition effects have been wrongly discounted unless they
are unusually strong.
A second argument (originally put forward by another Nobel Laureate,
Hans Krebs140 ) reinforced the view that non–equilibrium enzymes can be
rate–limiting, but near–equilibrium enzymes cannot. Suppose more enzyme
is added to catalyze a reaction that is already near–equilibrium; this will tend
to increase the amount of product and decrease the amount of substrate to
bring the reaction closer to equilibrium (as Eqn. [1.6] shows for an increase
in vf ). However, since the reaction is already close to equilibrium, if it is
brought closer to equilibrium, there will not be much change in the metabolite
concentrations, nor will the reaction using the product be stimulated to go
much faster so there is no reason to expect a great change in the pathway rate.
On the other hand, for a non–equilibrium reaction, increasing the amount
of enzyme could bring the reaction much closer to equilibrium, causing the
product concentration to rise significantly, and stimulating the next reaction
to go faster. Thus the change could propagate through the pathway to cause
a significant change in pathway flux.
These arguments may seem persuasive and in many cases they probably
have some validity. A potential weakness though is that they focus too closely
on a single reaction in the pathway, and give insufficient weight to how the
metabolite concentrations will change as the whole system responds to an
attempt to change the rate at one step. Later we will see why these fears are
justified (Chapter 5.5.1).
18 1 Introduction: regulation and control
This concept has been described as a ‘truism’ and ‘self–evident’, and versions
of it continue to be stated in biochemistry textbooks up to the present day.
In many ways this is odd, because it is difficult to give a literal interpretation
of the concept. As explained in the earlier section on the steady state (p. 10),
a requirement of successive steps along a linear metabolic pathway is that
they are operating at the same rate in the steady state, so none of them
could be termed the ‘slowest’ step in the normal sense of the word. What was
meant by the many advocates of the ‘rate–limiting step’ was that the rate of
a pathway could be altered only by changing the activity of one particular
enzyme in the pathway. On this basis, the study of the control of a pathway
entailed the identification of the rate–limiting step. From time to time, doubts
were expressed: why should the fact that the alteration of the activity of one
enzyme can change the rate of a pathway mean that this cannot be true of
another enzyme in the pathway? Attempts to analyse the kinetics of multi–
reaction systems did not support the rate–limiting step; it was possible to
show that the overall rate depended in principle on the kinetics of every
reaction in the system. Until recently, these criticisms had little impact, but
this is changing with the growing acceptance of Metabolic Control Analysis.
From Chapter 5 on, the theoretical and experimental grounds for rejecting
the rate–limiting step concept will be presented. In summary, it has been
found that more than one enzyme can affect the rate of a pathway; although
it is possible to imagine conditions where only one step affects the rate of a
pathway, experimental studies show that this is not the usual case.
In studying the control of metabolism, we therefore have a problem. Much
of the work that has been done has been interpreted in terms of the concept
of the rate–limiting enzyme. This work cannot be disregarded, because it
has given us methodologies for studying metabolic pathways, has provided
evidence about what does happen in pathways as their rates change and has
led to the discovery of many mechanisms that appear to have a role in the
control of pathway rates. Since it is necessary to look back at this work, as
1.4 Basic theory of metabolism 19
1.5 Summary
1. Regulation is homoeostasis, or the maintenance of constant conditions
in the face of external perturbations.
2. Control is the ability to make changes as necessary.
3. Regulation and control are properties of metabolic systems of great com-
plexity and the unsolved challenge is to link our knowledge of molecular
details to system–level explanations in a convincing yet feasible manner.
1.6 Appendix: free energy changes 21
6. Enzymes that catalyze reactions that are far from equilibrium are thought
to be more important in regulation and control than those whose reac-
tions are near to equilibrium.
Further reading
1. Atkinson, D. E. Cellular Energy Metabolism and its Regulation. Aca-
demic Press, New York (1977).
2. van Dam, K. Biochemistry is a Quantitative Science. Trends Biochem.
Sci. 11, 13–14 (1986).
3. Kacser, H. Control of Metabolism, Chap. 2 (sections I & IIA–C, pp. 39–
44) in The Biochemistry of Plants, Vol. 11, D. D. Davies (ed.) Academic
Press, New York (1987).
Problems
1. Phosphofructokinase catalyses the reaction:
Fru-6-P + ATP *
) Fru-1,6-bisP + ADP
The following substrate and product concentrations were measured in
the perfused rat heart: Rat heart contains 5.67 cm3 of H2 O/g dry
Glu-6-P *
) Fru-6-P
2.1 Introduction
Although my aim in this book is to describe a new approach to the inter-
pretation of control and regulation in metabolism, the experimental methods
that are needed are exactly the same as those traditionally used in biochem-
istry. This chapter therefore introduces the experimental measurements that
have been used in the past to develop explanations of control and regulation
and that continue to be used in the growing body of experimental studies in
Metabolic Control Analysis.
There are four main issues that have to be addressed before carrying out
most experiments in metabolism:
23
24 2 Methods for studying metabolism and its regulation
reasons for this, though, are varied. Some will relate to the extrinsic motiva-
tions for studying a particular metabolic pathway in the first instance. For
example, if the aim is to explore the differences between an aspect of human
metabolism in health and disease, and if suitable human cell or tissue samples
cannot readily be obtained ethically, then it will almost certainly be necessary
to use animal material. The animal would need to be no larger than necessary
to supply sufficient material for study (since larger animals are more expen-
sive to feed and house humanely), have a metabolism sufficiently similar to
humans for the results to be relevant, and have been sufficiently studied in
the past that the results can be placed in context. It is also an advantage if
inbred strains are available that have a restricted range of individual varia-
tion. In most cases, this points to the small rodents. Another aim might be
to study the metabolism of an organism of economic importance, for example
an animal or plant used as food, or a microorganism that produces a valuable
chemical such as a drug, and in this case there is no real choice. What I
am implying, of course, is that the extrinsic motivation to study a particular
organism arises from the existence of a organization prepared to fund bio-
chemists to carry out the experiments. (Unfortunately, there are few sources
of funding for carrying out basic research propelled by an intrinsic motiva-
tion to understand metabolic regulation better, in spite of the evidence that
the funded applied research is handicapped by the limitations in our basic
understanding.)
Given, then, that a choice of organism has been made, or forced, there
still remains the choice of how it, or parts of it, are to be used. Here there
are a range of options (not all of which are appropriate to all organisms):
After death of the organism, the full range of analytical biochemical tech-
niques can be applied to tissue samples, but this obviously terminates the
experiment with that individual, and different individuals must be analysed
to determine the effects of different experimental conditions.
The advantage of whole organisms studies is that metabolism is taking
place under physiologically normal conditions.
The disadvantages are that:
• physiologically normal conditions for multicellular organisms means that
there are metabolic interactions between tissues, which usually have dif-
ferent metabolic specializations. Observations will show the net result of
these interactions, as modulated by the effects of hormonal and nervous
stimuli (which may be difficult to control).
• metabolizable substrates must reach the site of their metabolism, and
the products escape, by crossing the natural permeability barriers. Since
these are selective, this reduces the scope for introducing artifical sub-
strates or changing the amounts of intermediates that are normally con-
fined within the cells.
• the natural biological variability of the organisms on which the exper-
iments are performed means that it is necessary to carry out replicate
experiments to determine the mean behaviour and to perform statisti-
cal tests to establish the significance of any observed differences between
control and experimental conditions. This increases the number of ex-
perimental subjects that have to be used, which increases the time taken
and the cost, and, in the case of animal experiments raises the ethical
dilemma of minimizing the number of individuals used whilst adequately
verifying the findings.
one organ per organism, then the numbers needed are reduced because
control and test experiments can be performed on matched organs from
the same organism.
• the procedure is technically simpler than organ perfusion, since the slices
just have to be incubated in oxygenated nutrient solution as the nutrient
supply is by diffusion, and
• fewer individuals are used in animal studies because many slices can
be taken from one organ and this also reduces the effects of biological
variability since control and test treatments can be compared on slices
from the same organ.
• there is inevitably cell damage at the cut surfaces, and substances re-
leased from the damaged cells may affect the undamaged ones, and
lives and can often be grown as cell suspensions. In plants, even differenti-
ated cells may dedifferentiate and grow well, whereas differentiation is largely
irreversible in animal cells.
Techniques have also been developed for isolating cells from animal organs.
For example, in 1969 Berry and Friend16 in Australia devised the method
commonly used these days for obtaining individual cells, hepatocytes, from
mammalian liver in high yield. Their innovation was to perfuse the liver
with an oxygenated solution containing proteolytic enzymes that attack the
intercellular ground substance and the adhesion factors on the cell surfaces.
After a short while, the tissue loses its integrity and can be dispersed to yield
free cells. Similar techniques have been applied to other tissues, for example,
giving fat cells (adipocytes) from fat tissue. In most cases, these isolated
cells will not grow in culture and can only be maintained for relatively short
periods.
Compared to the other systems examined so far, the advantages of using
isolated cells include:
• diffusion is entirely adequate for nutrient supply provided the cell sus-
pension is kept adequately mixed and there is sufficient surface area for
gas exchange (where necessary). The equipment required is therefore
relatively simple.
• many samples can be taken from the same batch of cells, and since
each sample might typically contain 105 –107 cells, inter–sample varia-
tion should be minimal. This should reduce the number of experiments
needed to establish statistically significant findings. Variability between
different experiments, and between different laboratories, can be reduced
if clones of cells are used.
• possible uncertainty about the extent of damage to the cells and the
degree of loss of enzymes etc.
• the possible dilution of key metabolites and enzymes if the cells are lysed
or homogenized (as is usual) in a liquid medium.
detection techniques can be linked to the enzyme assay. One that pro-
vides particularly high sensitivity for measuring metabolites in very low
concentrations is luminescence. This involves the detection of the weak
pulse of light produced by the light–emitting enzymes, the luciferases.
Luciferase derived from fireflies acts on ATP, so detects it and reactions
that produce it whilst bacterial luciferase reacts with NADH and so can
be used with coupled enzyme assays involving NAD–linked dehydroge-
nases.
2.3.3 Compartmentation
Tissues and cells contain a number of liquid compartments separated by se-
lectively permeable membranes. Most metabolites will not be present at a
uniform concentration in every compartment, and whilst average values may
show significant changes with experimental conditions, they cannot accurately
2.3 Measurement of metabolites 35
Metabolic Control Analysis were to emerge a few years later. In their method,
a tube for a small centrifuge that accelerates very rapidly is filled with a small
volume of perchloric acid solution overlaid by a layer of inert silicone oil. On
top of this is laid a medium containing the metal–chelating agent EDTA and
the detergent digitonin. Digitonin attacks cell membranes according to their
cholesterol content, which is much higher in plasma membranes than in mi-
tochondrial membranes. In a short period of exposure, therefore, the outer
membranes of the cell lyse, but the mitochondria remain intact. The whole
tube is cooled on ice before the cell sample is added to the top layer and
mixed. 10–15 seconds later, the centrifuge is started and the mitochondria
are rapidly spun down through the oil into the perchloric acid. The centrifuge
is stopped and the top layer containing the cytoplasm is acidified. The lower
layer is assayed to give the metabolite concentrations in the mitochondria and
the top layer those in the cytoplasm. The separation of the mitochondrial and
cytoplasmic compartments is 80–90% efficient, and changes in the cytosolic
adenine nucleotide levels by continuing metabolism are minimized by the low
temperature and the EDTA, which removes essential Mg2+ ions.
infer the pathway usage by calculating a pattern of pathway fluxes that ac-
counts for the mass balances between the inputs and outputs and ensures that
all the internal metabolites (including coenzymes such as NAD+ ) have rates
of formation equal to their rates of consumption. However, this does require
very extensive and accurate measurements of all the inputs and outputs of the
system, as well as full knowledge of the pathways that are potentially active.
The most common method of determining internal fluxes experimentally
is by isotopic labelling. The system under investigation is incubated for a
while to allow it to reach a metabolic steady state. One or more of the input
2.5 Measurement of metabolic fluxes 41
and molecular fragments on the basis of molecular mass with sufficient sensi-
tivity to distinguish molecules containing the rare isotope from those contain-
ing the natural one, so that the relative proportions of labelled and unlabelled
molecules can be measured in the same spectrum. Often the mass spectrom-
eter is used as a detector after a chromatographic separation of the sample
by gas liquid chromatography or HPLC, though it is not essential to purify
the target compound completely since the mass spectrum will separate the
impurities. Furthermore, since compounds break down in a reproducible way
into fragments in the mass spectrometer, the extent of labelling at particular
positions in the molecule can be determined from the mass spectrum of the
fragments. For example, the relative fluxes of gluconeogenesis and glucose
phosphorylation in rats after 17 hours fasting were estimated by the relative
rates of incorporation of deuterium (2 H) from heavy water ([2 H]H2 O) onto the
C–6 and C–2 carbon atoms of glucose respectively.227 Mass spectrometry, like
radioactivity measurements, has much higher sensitivity and precision than
NMR measurements of isotopic labelling. The major disadvantages compared
with radioisotope studies are the much higher cost and greater complexity of
the apparatus.
Nuclear magnetic resonance can be used to detect the spread of 13 C
through a pathway. Since individual carbon atoms give separate lines in
the spectrum, the position of labelling in the molecule is determined at the
same time (Fig. 2.5). As mentioned previously, the method can be applied to
live, metabolizing specimens, though the low sensitivity of detection of 13 C
means that it takes a long time to collect a measurement — a clear disadvan-
tage when the degree of labelling is changing. For these purposes, it is better
to collect samples as for other types of metabolic experiment and then take
the spectra of the extracts. Then the measurement time is unimportant, and
better accuracy can be achieved. The results are best interpreted by fitting
to a computer model of the metabolism. For example, citric acid cycle fluxes
in perfused rat hearts were measured and interpreted in this way by Chance
et al.36 in the experiment illustrated in Fig. 2.5. The kinetics of the pentose
phosphate pathway in human red blood cells were also analyzed by 13 C NMR
by Kuchel’s group in Sydney, Australia, combined with computer modelling of
the enzyme reactions involved.17 Computer modelling is not always necessary,
as in Sheila Cohen’s measurements of the relative rates of the pyruvate kinase
reaction to gluconeogenic flux in perfused livers of normal and diabetic rats
using in vivo NMR studies of the redistribution of label from the substrate
[3– 13 C] alanine.46
NMR can also measure the rates of individual steps in metabolism in vivo
in favourable circumstances by magnetic labelling using a technique called
magnetization transfer. The sample is irradiated with radio waves of the
specific frequency that is absorbed by certain atoms of a particular compound,
for example the terminal phosphate of ATP. This causes the magnetic axis of
the nuclei to change their alignment in the magnetic field of the instrument,
and if the irradiation is strong enough, most of them will become aligned
in the same direction. If one of these atoms is transferred by a reaction
2.5 Measurement of metabolic fluxes 43
Fig. 2.5: 13 C NMR spectrum of a perchloric acid extract of rat heart perfused
with glucose and [2-13 C]acetate.
GluC2 , (GC2 ) etc are the groups of lines corresponding to carbon atoms 2, 3 and 4 of
glutamate (i.e. oxoglutarate), taken from Fig. 3, Chance et al. 36 The spectrum is the
average of 2000 scans collected over more than an hour. T 1 and T2 are carbon atoms
1 and 2 of naturally abundant taurine.
into another chemical grouping, it takes its alignment with it for a while,
until it spontaneously flips back. If enough of the atoms are transferring to
this other compound at a rate comparable to the rate at which the magnetic
alignment is lost, they affect the spectral line recorded from the new compound
in a manner that allows the rate of transfer to be recorded. There are a
number of ways the measurements can be taken, but one of the simplest
measures the rate at which the atoms transfer from the compound whose
spectrum is being measured to the compound being irradiated. (Perhaps
rather counterintuitively, since at first sight it looks as though it is the transfer
44 2 Methods for studying metabolism and its regulation
2.6 Summary
1. Methodologies exist to study metabolism in systems at all levels of or-
ganization:
Further reading
1. Denton, R. M. & Pogson, C. I. Metabolic Regulation Chapman Hall
(1976).
3.1 Introduction
A description of metabolic regulation at the molecular level must include an
account of how the rates of the enzyme–catalyzed reactions vary with the
concentrations of the metabolites in the cell. The metabolites that most
obviously affect any enzyme are its substrates, and the best–known model for
enzyme action on a substrate gives rise to the Michaelis–Menten equation.
The dominance of this model in the context of metabolism is unfortunate,
for, as I will show, it is over–simplified for our purposes and can too easily
be misapplied. This chapter starts by reviewing Michaelis–Menten kinetics
and then presents some of the other aspects of kinetics relevant to metabolic
regulation, in particular:
47
48 3 Enzyme activity: the molecular basis for its regulation
• the majority of enzymes have more than one substrate (it is only amongst
the hydrolases and isomerases that single substrates are the norm), and
SV
v= (3.1)
S + Km
1 k k2
E + S −→
←− ES −→ E + P
k−1
They assumed that the reactions between E, S and ES were so rapid that the
process was virtually at equilibrium (the rapid equilibrium assumption), with
the position of equilibrium essentially undisturbed by the much slower reaction
whereby ES broke down to E and P . If we draw an analogy with the problem
of understanding what determines the rate of a whole metabolic pathway,
then Michaelis and Menten were effectively declaring the breakdown of ES
to product to be the rate–limiting step of catalysis. Under these conditions,
Km in Eqn. [3.1] is equal to the equilibrium constant for the dissociation of
ES to E and S.
3.2 Michaelis–Menten kinetics 49
1.0
Rate, v
0.5
S = Km
0.0
0 5 10
Substrate concentration, S
Briggs and Haldane later showed that the same equation could be derived
without any assumption about the relative rates of the reactions or their close-
ness to equilibrium. They used what they called the steady state assumption;
that is, they assumed that the rate of formation of the ES complex becomes
equal to its rate of breakdown shortly after mixing the enzyme and substrate
(c.f. Chapter 1.4.1). In Scheme 3.1 above, there is a single reaction form-
ing ES, which by the normal laws of chemical kinetics takes place at a rate
vf = k1 E S (i. e. rate equals a rate constant times the concentration of each
reactant). There are two routes of breakdown, one to E and S, the other to
E and P ; thus the breakdown rate, vb , is given by vb = k−1 ES + k2 ES. By
equating the rates of formation and breakdown, rearranging the equation and
making a few substitutions, the Michaelis–Menten equation can be derived, as
shown in most biochemistry texts. The effects of their change in assumptions
relative to those of Michaelis and Menten is buried in the interpretation of
Km : in the Michaelis-Menten derivation, Km equals the dissociation constant
for ES, or k−1 /k1 ; in the Briggs-Haldane derivation, Km = (k−1 + k2 )/k1 .
I have highlighted the distinction between these two derivations to empha-
size that neither of them is a factual description of enzyme activity; they are
conceptual and mathematical models. Any particular experimental observa-
tion that the rate of an enzyme-catalyzed reaction depends on the substrate
concentration in the way predicted by the Michaelis–Menten equation merely
demonstrates that either model is an adequate mathematical description of
the phenomenon for our purposes. We must look to other lines of evidence
(such as the existence of enzyme-substrate complexes) to support the molec-
50 3 Enzyme activity: the molecular basis for its regulation
ular details of the theory. Another characteristic of the models is that they
are mathematical approximations; both derivations assume that the substrate
concentration is very much higher than that of the enzyme, which is generally
true in laboratory experiments, but not always true in the intracellular envi-
ronment. Additionally, it is clear in the case of the Briggs–Haldane derivation
that the equation is an approximation because the steady state assumption
is not true on very short time scales (before the steady state concentration
of ES builds up). Indeed it is only approximately true the rest of the time
because there is a continual slow decline in the amount of ES, correspond-
ing to the gradual slowing down of the rate of reaction as the substrate is
exhausted, which in turn means that the rate of formation of ES must be
very slightly less than its rate of breakdown. Generally we can expect that
the errors introduced by these approximations will be much smaller than the
inherent observational inaccuracies in our experiments. This should not make
us forget that we have chosen to use a simple model that is adequate for our
purposes rather than a more elaborate one that would accurately reflect the
underlying molecular mechanisms. It is also worth noting that the derivation
involves the same assumptions of a time hierarchy of events that is necessary
in considering metabolic pathways, along with the same decision to ignore
events on very short and very long timescales (c.f. Chapter 1.3, p. 8).
For a more realistic molecular model of enzyme action, we could take into
account the reversibility of enzyme catalysis: the enzyme can catalyze the
conversion of product to substrate, so it must be able to form an EP complex
as well as an ES complex. The pathway from ES to EP passes through a
transient, intermediate form, the transition state ET , so if we also included
this, in initial rate conditions where product is absent, Scheme 3.1 would
become:
Scheme 3.2
1 k 2 3 kk4 k
E + S −→
←− ES −→
←− ET −→
←− EP −→ E + P
k−1 k−2 k−3
There is a lot more work involved in solving the initial rate equation, and the
result is to some extent disappointing because it turns out to be in the same
form as the Michaelis–Menten equation, except that Km is now an expression
involving all seven rate constants.
in this form is not a very practicable method for determining the parameters
because an enzyme kinetics experiment does not give the curve directly; in-
stead it gives a small set of rate measurements made at particular starting
concentrations of substrate. Furthermore, these discrete points on the graph
include experimental errors which move them off the curve, and it is not pos-
sible to sketch an accurate rectangular hyperbola that passes as closely as
possible to the points.
In most branches of science until recently, the usual solution to the problem
of analysing a non–linear relationship like the Michaelis–Menten equation was
to rearrange the equation to give a linear relationship, so that the points could
be plotted as a graph and the best straight line drawn through them. It is
very unfortunate that the most popular linearization for enzyme kinetics has
been that proposed by Lineweaver and Burk:
µ ¶
1 Km 1 1
= + (3.2)
v V S V
This corresponds to the equation of a straight line, y = mx + c, with y = 1/v,
x = 1/S, the gradient m = Km /V and the y-intercept c = 1/V (Fig. 3.2).
The problem with this method is that using the reciprocals of v and S causes
extreme stretching of the axes at low substrate concentrations. This is il-
lustrated in Fig. 3.2 by showing what happens to a set of equally–spaced
5
1/v
0
0 1 2 3 4 5
1/S
S/v 4
0
0 1 2 3 4 5
S 1 Km
= S+ (3.3)
v V V
This equation corresponds to a line of slope 1/V , with y–intercept Km /V
and x-intercept of −Km (Fig. 3.3). There is no distortion of the substrate
concentration spacing, since this is used unaltered as the x–axis, and the
errors in the velocity measurements are not unduly exaggerated or suppressed
in any part of the graph. The use of linear regression to determine the best-fit
straight line is therefore justifiable. This graph can equally as well be used
3.2 Michaelis–Menten kinetics 53
1.0
0.8
2
0.6
Rate, v
0.4
0.2
0.0
0 1 2 3 4
Substrate concentration, S
where vnet is the net rate (positive for formation of P ), Vf is the limiting rate
in the forward direction, Km,S and Km,P are the Km values for substrate and
product and Keq is the equilibrium constant for the reaction. (This equation
can be found in other forms in some books because there is also a Vr for
the limiting rate in the reverse direction; however one of the parameters can
always be eliminated because of the constraint imposed by the equilibrium
constant known as the Haldane relationship: Keq = Vf Km,P /(Vr Km,S ).) An
illustration of the way the net rate of the reaction varies with both S and P is
shown in the three-dimensional plot of Fig. 3.5. The exact details of the form
56 3 Enzyme activity: the molecular basis for its regulation
would depend on the relative values of the constants and the equilibrium con-
stant. It is clear though, that once both substrate and product are present,
the response of an enzyme cannot necessarily be visualized by thinking about
where, on a rectangular hyperbola, the substrate concentration would fall rel-
ative to the Km,S . Later we will look at how we can characterize the position
on this 3-D graph that corresponds to particular intracellular concentrations
of S and P when we look at the concept of elasticity (Chapter 5.3). For the
moment, note how the diagram shows that a substrate concentration above
the Km value does not guarantee that the enzyme rate is approaching V ;
that will depend on the product concentration and the equilibrium constant.
An example where this has been overlooked is at the enzyme aldolase in the
glycolytic pathway. In certain cells, the concentration of substrate (fruct-
ose 1,6 bisphosphate) is greater than the Km , and this has been erroneously
interpreted to suggest that the enzyme is working near to its limiting rate.
7.5
Rate
5.0
2.5
0.0
-2.5
10
10 8
8 6
P 6 4 S
4 2
2 0
0
Fig. 3.5: Simultaneous dependence of enzyme rate on both substrate and product.
The parameters in Eqn. [3.4] have been set to: Km,S = 1; Vf = 10; Km,P = 2, and
Keq = 4.
The compulsory order mechanism would occur when the enzyme molecule
has a binding site for one of the substrates (say A), but not the other; when
this substrate binds, a conformational change in the enzyme (as proposed in
Koshland’s induced fit model of enzyme action) generates the binding site for
the second substrate, B. The ternary complex EAB therefore forms by an
obligatory sequence and converts to the ternary product complex, EP Q, by
the catalytic action of the enzyme. The product complex then breaks down
to release products, thus:
Scheme 3.4
E+A ←−
−→ EA
EA + B ←−
−→ EAB
EAB ←−
−→ EP Q
EP Q −→ E+P +Q
The random order mechanism also forms the ternary complex EAB in two
stages, but occurs when the enzyme has binding sites for both substrates so
that either one may bind before the other (though not necessarily with equal
probability).
Scheme 3.5
EA
%. -&
E EAB
-& %.
EB
EAB ←−
−→ EP Q
EP Q −→ E+P +Q
The double displacement mechanism differs from the previous two in that
a ternary complex is not formed. Instead, a chemical entity that is transferred
between the two substrates is temporarily parked on the enzyme molecule by
the first substrate, and then collected by the second substrate to complete the
catalytic cycle:
Scheme 3.6
E+A ←−
−→ EA
EA −→ EX + P
EX + B ←−
−→ EXB
EXB −→ E+Q
The scheme shows that B and P are related in that they both must have an
attachment point for X, and A and Q are related in that they both consist
of something linked to X (A = P + X and Q = B + X).
3.3 Two–substrate enzymes 59
V AB
v= (3.5)
KiA KB + KB A + KA B + AB
The meanings of the different parameters in this equation are as follows. V
is, just as for single substrate enzymes, the limiting rate approached as the
concentrations (but this time of both the substrates) increase to very high
values. Suppose the concentration B is fixed at a very high level and A is
varied; Eqn. [3.5] above can have both top and bottom lines of its right hand
side divided by AB to give:
V
v= (3.6)
KiA KB KB K
+ + A +1
AB B A
We can suppose that B can be raised high enough that the terms in the
denominator of the equation that are divided by B become very much smaller
than the others, so that we can ignore them and get the approximation:
V VA
v≈ = (3.7)
KA KA + A
+1
A
This is the same as the single substrate Michaelis–Menten equation (Eqn. [3.1]),
with A instead of S and KA in place of Km ; in other words, when B is fixed
at a high value and A is varied, the enzyme behaves like a single substrate en-
zyme, and KA is the Km for A under these conditions. By exchanging A and
B, we can conclude that at high concentrations of A, the rate will vary with
the concentration of B like a single substrate enzyme with KB being the Km
under these circumstances. Thus KA and KB can legitimately be regarded
as the Km values for these two substrates, except that we only know this to
be true when the concentration of the other substrate is very high, which is
unlikely to be the case in intracellular conditions for most enzymes. What
happens at lower concentrations of the other substrate? To illustrate this,
let us consider very low concentrations of B. As B decreases, the numerator
term of Eqn. [3.5] becomes smaller because the whole of it is multiplied by B;
the denominator, however, only shrinks to a certain extent. The denominator
terms mutiplied by B certainly become smaller and smaller, but the terms
without B are unaffected and, since the different terms are added together,
60 3 Enzyme activity: the molecular basis for its regulation
This means that the enzyme again behaves like a Michaelis–Menten enzyme
when A is varied. Comparison with Eqn. [3.1] shows that the apparent limiting
rate, Vapp is given by Vapp = V B/KB , and the Km is KiA . At intermediate
concentrations of B, it is possible to show that the apparent Km for A is
between KA and KiA . Thus although it is possible to study a two–substrate
enzyme by holding one substrate constant and varying the other, the results
have limited value. It is necessary to determine the values of all four param-
eters V , KA , KB and KiA to calculate how rapidly the enzyme will work at
any particular concentrations of A and B. How the parameters are found is
described next.
Concn. Concentration of A
of B A1 A2 A3 A4 A5
B1 v1,1 v1,2 · · ·
A1 A2
v1,1 v1,2
B2 v2,1 ···
A1
v2,1
B3 ···
B4 ···
B5 ··· v5,5
A5
v5,5
values of 1/Vapp . Two secondary plots are then constructed. The first of
these involves plotting each slope value times the value of B at which it was
measured against B (Fig. 3.7a). The slope of this plot gives 1/V , and the y-
axis intercept is KB /V . (The x-axis intercept is −KB .) The other secondary
plot is each intercept value from the primary plot times the value of B at
which it was measured against B (Fig. 3.7b). The slope of this plot is K A /V ,
and since V has already been determined, this gives KA . The intercept on the
y-axis is KiA KB /V , and since KB /V has been determined, this gives KiA .
The values of the parameters obtained by this graphical analysis can them-
selves be reported as the properties of the enzyme. These days, they can also
be used as the initial estimates in an iterative computer fitting procedure, and
the results from this are then taken as the best estimates of the parameters.
In this case, the ultimate worth of the graphical analysis is that it gives visual
confirmation that the results do fit the two substrate equation. (If they do
not, there will be systematic deviation of the points from a straight line.)
What can the results tell us about the mechanism of the enzyme? We
need to look at this for each mechanism in turn.
6
....... B1
intercept
.. .
...
..........
.
........ ........ B2
...
.
..
............. .................
. .
........ .......... ......
...
.
............... ................. .................... 4 B3
. . .. .
........ .......... ................ .....................?. B4
...
.
............................. ..............4 ...
...... ............ B
A/v
. .. ...
....... ......... ............................?............................. ◦ 5
. ........ ......... ........4 .. ..
... ..................................................................?....................................◦....
. ..
.... ..
.... .....
4. . ......
. . .....
◦.... . .
....
.....................................................?............................ ...
.................................
..
......................................4
.....
..
...
.........
.. ◦
...
.
..
... . . ..
..
.. ... . . . . . . .
. .
. . . . . .
.?
◦.
.... slope=1/V
. .
. ..
.............. .............. ........... . . app
. . . .
.
..
.
.
.
.
.
..
.
.. .
. .
. ..
. -
. .
... ..........
.
.
...
..... A
from the initial rate kinetics experiments described above. Furthermore, the
results can equally be used to calculate a Ki term (i.e. KiB ) for the other
substrate even though, if B is the second substrate to bind, the EB complex
is not formed when enzyme and B are mixed together, so that a dissociation
constant for the EB complex cannot be measured. Therefore if a purified
sample of the enzyme is available, a comparison of the kinetic results with
binding studies (see p. 65) may show that the enzyme obeys a compulsory
order mechanism. Another experiment that can be used to diagnose this
mechanism is possible if there are a number of different substrates on which
the enzyme can act. For example, alcohol dehydrogenases will act on alcohols
with carbon chains of different lengths, though generally at different rates
with different substrates. Suppose that there are a number of substrates of
the B–type: B1, B2 etc. If the kinetics are determined using A and each of
these substrates in turn, KiA is the same whichever B substrate is used. On
the other hand, if the kinetics are studied with a number of different variants
of the A substrate, KiB will vary with the nature of the A substrate. This is
because the binding of the first substrate cannot be affected by the nature of a
substrate that has not yet bound, whereas the binding of the second substrate
is obviously influenced by the molecule already at the active site.
Aspartate transcarbamylase catalyzes the reaction:
carbamoyl phosphate + aspartate *
) carbamoyl aspartate + phos-
phate
The enzyme has some very interesting properties that will be discussed later
in the chapter in connection with allosteric enzymes (section 3.5), but its ki-
netic mechanism is ordered. Carbamoyl phosphate binds first, followed by
3.3 Two–substrate enzymes 63
6 6
a) ¡ b)
B x intercepts
¡◦
B x slopes
?¡ ◦
¡ #
#
¡@I ?
#
@ # o
S
¡4 1/V 4
# S KA /V
¡ #
¡ #
¡ #
¡ #
¾ K /V #¾ KiA KB /V
¡ B
#
¡ - # -
6 B 6 B
−KB −KiB
that they obey a special form of the random order mechanism, but because
of cooperative binding that will be described later (p. 71). In practice, it
seems that most enzymes working by the random order mechanism do actu-
ally follow the two–substrate rate equation. One characteristic of the random
order mechanism is that binding of both substrates to the enzyme will be
demonstrable in binding studies, with the dissociation constants for the EA
and EB complexes being KiA and KiB respectively. Unlike the compulsory
order mechanism, these Ki terms would not be expected to be invariant if
different forms of the other substrate were used in the kinetic studies.
The effects of products in inhibiting the reaction can distinguish between
compulsory order and random mechanisms. Whether the inhibition by each
product with respect to each substrate is of the competitive or noncompetitive
type gives a set of results that differs between the two mechanisms.
One difficulty in distinguishing between the random order and compul-
sory order mechanisms is that in kinetic experiments, the dividing line may
be fuzzy. Although a random order enzyme may bind either substrate first,
it is unlikely that both routes will be equally probable. If one route is al-
ways dominant under the conditions investigated, then the enzyme is effec-
tively compulsory order, even though on the longer time scales of a binding
experiment, the binding of both substrates to the enzyme might be demon-
strable. This might explain why some enzymes have been difficult to classify,
with different laboratories reaching opposite conclusions. The hexokinase of
yeast is one example where contradictory conclusions were reached by dif-
ferent groups: some favoured an ordered mechanism in which glucose bound
first, whereas others thought that the mechanism was random. However,
the probable explanation is that the enzyme follows a steady state (rather
than rapid equilibrium) random mechanism, in which only the binary com-
plexes, enzyme–glucose and enzyme–ATP, are near to equilibrium with the
enzyme.202 (In these conditions, the random mechanism could show hybrid
characteristics with the compulsory order mechanism.)
second substrate, and the net reaction stops when all the enzyme has been
converted into the modified form. If larger amounts of enzyme are available,
it may be possible to detect and characterize the modified form of the enzyme.
However, even when the amount of enzyme is small, so that the net product
formation is also small, the equilibrium between substrate and product is
being catalyzed, and it is possible to detect this reaction by isotope exchange.
Thus, suppose the partial reaction with one substrate AX is:
←− EX + P
E + AX −→
This has two partial reactions: in the first, ATP and HCO− 3 react with the
enzyme to form an enzyme–CO2 complex (in which the CO2 is attached to the
prosthetic group biotin), and in the second, the enzyme–CO2 complex reacts
with pyruvate to form oxaloacetate. The enzyme–CO2 complex has been
isolated after incubation with [14 C]HCO− 3 to demonstrate the first partial
reaction, and the exchange of [14 C] from pyruvate into oxaloacetate has been
detected to show the second partial reaction.252
P + L ←−
−→ P L
[P ][L]
Kd =
[Lb ]
where Kd has the units of concentration. Furthermore, when the total pro-
tein Pt is divided equally between free protein, P , and protein–ligand complex,
P L = Lb , Kd equals the free ligand concentration [L]. Although the disso-
ciation constant has this advantage that its value corresponds to the concen-
tration of ligand needed to fill half the binding sites, it has the disadvantage
that strong binding (or high ligand affinity) corresponds to low values of K d ,
and weak binding corresponds to high values. (The equilibrium defined in
the opposite direction is the association constant, Ka ; Ka = 1/Kd , and large
values correspond to strong binding, but the units of Ka are the reciprocal of
concentration.) The equation for Kd can be rearranged with substitution of
[Pt ] = [P ] + [Lb ] to give:
[Pt ].[L]
[Lb ] =
Kd + [L]
This is analogous to the Michaelis–Menten equation (not surprisingly, as it was
the origin of their derivation) and therefore defines a rectangular hyperbolic
binding curve.
Many enzymes are composed of a number of identical subunits (most com-
monly four). Binding will involve the formation of complexes with 1, 2 . . . n
molecules of L at various stages of the binding curve:
P + L ←−
−→ P L ←−
−→ P L2 · · · ←−
−→ P Ln
If the binding sites do not interact in any way, then defining the concentration
of bound ligand, Lb as Lb = P L + 2P L2 · · · + nP Ln leads to essentially the
same equation:
n.Pt .L
Lb = (3.9)
Kd + L
Thus the binding curve can be characterized by the values of n and Kd , and
this is done by measuring Lb at various concentrations of L. Sometimes it
is easier to determine the fraction of binding sites occupied, the fractional
3.4 Binding characteristics of enzyme sites 67
2.0
1.5
b
Lb /(Pt L)
1.0
0.5 c
0.0
0.0 1.0 2.0 3.0 4.0
Lb /Pt
saturation, Y , where
Lb L
Y = = (3.10)
n.Pt Kd + L
in which case it is still possible to determine Kd .
The analysis of the results of binding experiments has parallels with that
of enzyme kinetics. There are graphical methods based on the linearization of
the binding equation. For example, the Hughes–Klotz plot (an analogue of the
Lineweaver–Burk plot) consists of plotting 1/Lb against 1/Lf ; however this
has not been favoured and it suffers from the same defects as the Lineweaver–
Burk plot. The plot used most frequently is known as the Scatchard plot. It
consists of plotting Lb /(Pt .L) against Lb /Pt (Fig. 3.8) and is an analogue of
the enzyme kinetics plot of v/S against v, known as the Eadie–Hofstee plot.
The slope of the graph (line a) is −1/Kd and the intercept on the x-axis is
the number of binding sites, n.
However, the method of preference for analysing binding results is com-
puterized fitting of the binding equation. Since this is essentially the same
as the Michaelis–Menten equation, a computer program suitable for one task
may be satisfactory for the other.
possible reasons for this. One is that there is a specific high–affinity site for a
particular ligand, but the ligand can also bind with lower affinity and speci-
ficity to another part of the protein through hydrophobic or ionic interactions.
Another possibility is that the protein sample contains a mixture of proteins
of different binding affinities, either because the protein exists in a number
of variant isoforms in vivo or because partial damage during preparation has
modified the binding properties of some of the molecules.
If the different sites do not interact in any way, they behave independently
and the total binding observed is just the sum of the binding at the separate
sites. Thus for a protein molecule with two types of site, of which there are
n1 and n2 per molecule, with dissociation constants Kd,1 and Kd,2 , the total
binding is:
µ ¶
n1 .L n2 .L
Lb = P t +
Kd,1 + L Kd,2 + L
In this case, the linear graphical plots are curved (e.g. Fig. 3.8, curve b).
There has been a general belief that the two components can be estimated by
drawing tangents to both ends of the curve on a Scatchard plot. In spite of the
many instances of this continuing practice in the biochemical and physiological
research literature, it has long been known that such a procedure gives grossly
incorrect estimates of the two sets of parameters.134, 176 For example, no part
of curve b in Fig. 3.8 can be extrapolated to 2 on the x–axis — the correct
value of the number of higher–affinity sites. In such cases computer fitting to
the equation above can often yield the parameters of the two components.
is:
O2n
Y = (3.11)
Kdn + O2n
This equation does give rise to a sigmoid curve if n is greater than 1; for
2
log [Y/(1 - Y)]
-1
-2 B
-3
A
-4
log S
before, and so on at each stage. This requires that the binding of oxygen at
one site transmits an influence to the other sites to aid their binding; this
effect is known as cooperativity. However, for much of the century, little was
known about the molecular mechanism by which this could occur.
On the basis of energetic considerations, it was realized that the Hill equa-
tion was not a feasible description of cooperative binding, since it would re-
quire an infinite interaction energy to ensure that the affinity of all the sites
increased to such an extent that they immediately filled up after the first
oxygen had bound. With a finite interaction energy, the binding curve would
give a sigmoid curve in the Hill plot (see Fig. 3.9), having slopes of 1 at both
ends (i.e. at very low saturations and very high saturations). This is because
at very low concentrations of oxygen, only the reaction
←− HbO2
Hb + O2 −→
is being seen. Similarly, when the oxygen concentration is very high, the only
reaction being observed is
←− Hb(O2 )4
Hb(O2 )3 + O2 −→
Furthermore, the Hill coefficient could only have a maximum value equal to
the number of subunits, n, for an infinite interaction energy, and with a finite
interaction energy, nh would be less than n.
Thus the original basis of the Hill equation is no longer accepted. Nev-
ertheless, many proteins in addition to haemoglobin exhibit cooperativity in
their ligand binding, and the Hill plot is still widely used to characterize the
cooperativity. This is partly because the other forms of binding curve analy-
sis, such as the Scatchard plot, do not give linear plots for cooperative binding
and cannot therefore be used to analyse the results (e.g. see Fig. 3.8, curve
c). The Hill plot, on the other hand, tends to have a substantial linear por-
tion in the centre; the curvature towards a line with a slope of one is only
observable if the extreme ends of the curve can be measured. Furthermore,
the Hill coefficient, nh , can be used to diagnose cooperativity even when the
degree of sigmoidicity of the binding curve is difficult to detect by eye. If the
value of nh is greater than 1, the binding can be said to be cooperative (or
even positively cooperative). A value of 1 corresponds to ordinary hyperbolic
binding that should be analysable by methods such as the Scatchard plot.
However, it is possible to obtain values of nh of less than 1. For example, a
mixture of independent binding sites of different affinity gives this result. It is
also possible to conceive of the case where the first ligand to bind to a protein
causes a lowering of affinity of the other sites, so they become progressively
more difficult to fill. Since this is the opposite effect to the cooperativity
described above for haemoglobin, it is termed negative cooperativity. It gives
rise to non–hyperbolic binding curves and Hill coefficients of less than 1. The
difficulty is that the binding experiments themselves do not distinguish be-
tween these two causes for nh being less than 1. As a result there is sometimes
3.5 Allosteric enzymes 71
0
0 5 10 15 20 25
[Aspartate], mM
Fig. 3.10: The kinetics of the allosteric enzyme aspartate transcarbamylase (AT-
Case).
The rate of reaction is shown with respect to one of its substrates, aspartate (¦). The
enzyme is inhibited in a feedback manner by the pathway end–product CTP (+). ATP
acts as an activator ( ). (Based on results of Pigiet, Yang & Schachmann. 82 )
• In the majority of cases, the enzyme rate gives a sigmoid curve when
plotted against the concentration of at least one of the substrates, cor-
responding to a Hill coefficient of greater than one. The effectors work
by changing the S0.5 of the enzyme for one of its substrates. Activators
decrease the S0.5 so that, at a given substrate concentration, the velocity
is higher in the presence of the activator (Fig. 3.10). On the other hand,
inhibitors increase S0.5 so that, at a given substrate concentration, the
enzyme rate is lower (Fig. 3.10). In both cases, the Hill coefficient is
altered by the effectors. Enzymes whose activities are regulated in this
way are classed as ‘K systems’. There is a minority class, the ‘V sys-
tems’, where the effectors alter the rate without changing the S0.5 , and
these do not necessarily exhibit sigmoid rate curves with respect to their
substrates.
..........
..... ........
.... .. R
.... . .
............... .
∆G◦0 L ......... .
..... ........
.
......... ..
........ ....
..
.... .......... ... KR
... ..
... .
... T .........
... ..............
... ....... .... ......
.. . .... S
... ........KT ...... .... .. RS
.. . .................. ..................
...................
S ........
TS ....
....
........ ...
.
.
......... .........
...........
..... ......
TS2
S S ............................................ .... S S .... RS2
....
...................
.......... ........
.
......... ....
...
...
S S ..
TS3 ........
S ...................
................ ............
.
........ ...... ..... .....
. .... S S ....
........ ....S .... RS3
..............
S S .......
TS4 ....
S S ..
........... ...
...... ...
...... .
......
....... ........
.
..... ...........
..... ......
.... S S .... RS4
.... S S ..
.................
Fig. 3.11: The concerted allosteric enzyme model of Monod, Wyman & Changeux.
The T conformation is represented by squares and the R by circles. The different states
of the protein are placed vertically according to their relative standard free energy for
this specimen case. The exact relative relationships will vary with the values of the
equilibrium constants L, KR and KT .
where the denominator calculates the total concentration of ligand sites. Deriving
an equation like Eqn. [3.12] can be done by taking the diagram of the model (such
as Fig. 3.11) assigning one of the forms the concentration 1, calculating all the other
concentrations relative to it by using the equilibrium constants, and inserting these
concentrations into Eqn. [3.13].
The Monod, Wyman and Changeux equation (Eqn. [3.12]) therefore describes
cooperative binding using three parameters, L, c and KR ; it is assumed that n is
already known by other means. If c = 1 (corresponding to equal substrate affinity of
the R and T states) or L = 0 (corresponding to no observable T state) the equation
can be shown to reduce to an ordinary hyperbolic binding function like Eqn. [3.10]:
α S
Y = =
1+α KR + S
KR determines the left–right positioning of the top of the curve on the Hill plot
(Fig. 3.9). L and c determine the shape of the Hill plot curve. It can also be shown
that the maximum value of the Hill coefficient, nh , depends on the values of L and
c, but never exceeds the number of binding sites, n.
The model also accounts for the action of allosteric effectors. An inhibitor
is a molecule that binds preferentially to the T state and stabilizes it, just
as the substrate stabilizes the R state. Because inhibitors stabilize the T
state, they make it more difficult for the substrate to convert the molecule to
the R state. An extremely effective inhibitor would stabilize the T state so
much that the protein could not escape from it. The substrate would then
always bind to low affinity T –state sites, and there would be no cooperativity
in the substrate binding. An activator works the other way round by bind-
ing preferentially to the R state and therefore stabilizing it. This will favour
the transition to the R-state earlier in the substrate binding curve. Again,
a highly effective activator would stabilize the protein almost entirely in the
R state, so that the binding of substrate would be to the high affinity sites
and the cooperativity in substrate binding would be lost. This explanation
76 3 Enzyme activity: the molecular basis for its regulation
requires that the binding of the inhibitor and the activator both exhibit co-
operativity. Indeed, the interactions between the different ligands, known as
heterotropic effects, depend for their occurrence on the cooperativity of the
substrate and/or the effector binding, the homotropic effects.
@ µ
¡
@
I ¡¡
@@ R
@ ¡¡
»
@ ª
¡
S
½
S
Fig. 3.12: The sequential allosteric enzyme model of Koshland, Nemethy & Filmer.
The A and B conformations are represented as squares and circle respectively. A tetramer
with square interaction geometry is illustrated. There are two possible distinct arrange-
ments of the tetramer with two substrates bound. 2 bars at the subunit interface show
an AA interaction, 1 an AB interaction and none a BB interaction.
worked out using the principles described above in relation to Eqn. [3.13]. However,
in this case, different equations are obtained depending on the number of subunits
and the interaction geometry, whereas a single equation (containing the number of
subunits as a parameter) suffices for the concerted model. The original derivations
considered the equilibrium constant (or free energy change) at each step in a scheme
such as Fig 3.12 to be made up from a number of components: KS , the intrinsic
dissociation constant for a subunit in the B state; Kt , the equilibrium constant for
the tertiary conformational change of the A subunit into a B subunit; KAA , the
equilibrium constant for the interaction of two adjacent A subunits (i.e. part of the
quaternary conformational energy); KAB , the equilibrium constant for the quater-
nary interaction between an A subunit and a B subunit, and KBB the equilibrium
constant for the interaction of an adjacent pair of B subunits. At each step, the
change in the subunit–subunit interactions, brought about by the binding of S, has
to be taken into account, and this can be different if different spatial arrangments
of S on the protein subunits are possible (as in the case of two S molecules bound
in Fig. 3.12). Finally, it is found that in the resulting binding equation, these basic
equilibrium constants always occur in certain combinations, so that they cannot be
independently determined. For example, KS and Kt always occur multiplied to-
gether, so that any pair of values that give the same product would give the same
binding curve. In fact, for similar reasons, the final binding equation for a square
tetramer can be expressed in terms of just two constants, formed of collections of
the five constants. Unfortunately, the way the constants are collected proves to be
different for each case of number of subunits and interaction geometry. Thus the
assumed molecular interactions underlying the binding behaviour are not experi-
mentally accessible from the binding curves.
1.0
0.8
0.4
0.2
0.0
0 5 10 15 20
3.5.3.1 Haemoglobin
Haemoglobin is an oxygen binding protein rather than an enzyme, but other-
wise it has been regarded as an example of an allosteric protein. Myoglobin
and haemoglobin were the two first proteins to have their 3-dimensional struc-
tures determined by x–ray crystallography by Max Perutz and his colleagues
at Cambridge. Myoglobin is an oxygen storage protein that is found in partic-
ularly high levels in the muscles of diving mammals such as the sperm whale.
It is relevant here because it contains a haem group and its polypeptide chain
is folded in a virtually identical manner to the subunits of haemoglobin. Myo-
globin, however, occurs as single subunits showing a hyperbolic binding curve
80 3 Enzyme activity: the molecular basis for its regulation
for oxygen, with a much greater affinity for oxygen than has haemoglobin. The
oxygen affinity of haemoglobin at the top end of its oxygen saturation curve,
where the fourth oxygen is binding, is comparable to the oxygen affinity of
myoglobin, so it appears that the assembly of haemoglobin into a tetrameric
structure in some way inhibits the oxygen binding, as would be expected for a
T –state structure in the concerted model of Monod, Wyman and Changeux.
The concept that there were two distinct conformational states of haemoglobin
was supported by the crystallography studies. Crystals of deoxyhaemoglobin
were grown and the protein structure determined in this state. Significantly,
if oxygen is allowed into contact with the crystals, they crack because of the
conformational change on oxygen binding. Crystals of oxyhaemoglobin had to
be grown separately for determining its structure; in fact, during the growth
and x–ray analysis of the crystals, the iron atom in the haem group becomes
oxidized, and met–haemoglobin (which does not bind oxygen) is formed, but
its conformation is believed to be the same as oxyhaemoglobin. The con-
formations of the subunits themselves scarcely differ between deoxy and oxy
haemoglobin; the big difference was found to be the relative positioning of
the subunits and the bonding between them. In accordance with the con-
certed model, there was more hydrogen bonding and salt linkages between
the subunits in the deoxy state than in the oxy. Perutz’s model of the binding
of oxygen to haemoglobin assumes that there is a concerted conformational
transition between the deoxy (T ) state and oxy (R) state because it is as if the
zig–zag surface of the interface has moved across by a complete notch, and any
intermediate position would not be tenable. Nevertheless, x–ray crystallogra-
phy gives a relatively static view of protein structures, which are known from
other evidence to be flexible and mobile, and there is little direct evidence of
the intermediate states in oxygen binding because they cannot be prepared
as crystals, so the concerted transition is inferred rather than observed. How-
ever, there are well over a hundred known mutant haemoglobins that differ
from normal haemoglobin by different substitutions of a single amino acid,
and the effects of these on the affinity and the cooperativity of the oxygen
binding can generally be interpreted in terms of the impact of the structural
change on the concerted model (see question 2 at the end of this chapter).
In most mammalian red blood cells, there are high concentrations of 2,3–
bisphosphoglycerate. This is normally only present in low concentrations in
cells, as a cofactor for the glycolytic enzyme phosphoglycerate mutase, but in
humans and many other mammals it acts as an allosteric inhibitor of oxygen
binding. Its concentration is around 5 mM, comparable to the concentration
of haemoglobin protein, but it varies slowly with conditions that affect the
oxygenation of the blood such as the low oxygen pressure found at high al-
titudes. It acts as an inhibitor of oxygen binding, and is responsible for the
oxygen affinity of blood being approximately half that of isolated haemoglobin,
though it also slightly increases the cooperativity of binding. X-ray crystal-
lography has shown that the binding site for 2,3–bisphosphoglycerate is in
the central cavity of deoxyhaemoglobin between the four subunits; this cavity
closes down in oxyhaemoglobin and would be too small to allow binding. Di-
3.5 Allosteric enzymes 81
3.6 Summary
1. The well–known Michaelis–Menten equation is not a complete descrip-
tion of the behaviour of single–substrate enzymes in vivo because of the
the effects of product inhibition and reversibility of the reaction.
• compulsory order;
• random order, and
• double displacement.
Further reading
Background to enzyme kinetics and allosteric enzymes: any biochemistry text
such as:
Computer programs
There are many programs available for fitting enzyme kinetics data. Hyper
by John Easterby is a Windows program for fitting Michaelis–Menten kinet-
ics, currently available as shareware on the Internet by anoymous FTP from
ftp.bio.indiana.edu in /molbio/ibmpc/hyper.zip, or from micros.hensa.ac.uk
in /mirrors/cica/win3/util/hyper.zip.
For fitting inhibition and two substrate kinetics, the program Leonora for
IBM PC-compatible microcomputers by Athel Cornish–Bowden is distributed
with his book explaining the theory behind the fitting procedures: Analysis
of enzyme kinetic data, Oxford University Press, Oxford, (1995).
Problems
1. A calcium–binding protein from bovine brain has a relative molecu-
lar mass of 17000. Its calcium–binding properties were investigated
by an equilibrium dialysis experiment, in which the protein solution (2
mg.cm−3 ) was placed in dialysis sacs that retain the protein and allowed
to come to equilibrium with outer solutions containing various amounts
of Ca2+ (in addition to buffer and salts). At equilibrium, the Ca2+
concentration is higher in the sacs than the outer solution since both
contain free Ca2+ at the same concentration, but the sacs also contain
protein–bound Ca2+ . The calcium concentrations at equilibrium were
measured by atomic absorption spectrophotometry, giving the results in
the table below. Determine the number of binding sites on the protein
for Ca2+ and their dissociation constant.
84 3 Enzyme activity: the molecular basis for its regulation
[Ca2+ ], µM
Outer solution Sac fluid
0.9 47
2.1 94
4.3 152
7.7 206
18.2 284
35.0 336
2. The following table contains data for the mid–regions of the oxygen bind-
ing curves of three haemoglobins measured in solution under identical
conditions. The oxygen concentration is reported as the partial pres-
sure of oxygen, PO2 , and the degree of oxygen binding by the fractional
saturation, S. The three haemoglobins are A (normal adult human),
Hiroshima (His 146 β → Asp) and Kansas (Asp 102 β → Thr). Plot a
graph to determine the Hill coefficient, nH , of each haemoglobin, and
speculate on how the effects on the cooperativity and oxygen affinity
could be interpreted in terms of the concerted model, given that the
mutations are at the subunit interfaces rather than the oxygen binding
sites.
A Hiroshima Kansas
P O2 S P O2 S P O2 S
0.67 0.044 0.133 0.045 1.33 0.193
1.33 0.243 0.267 0.179 2.67 0.371
2.00 0.500 0.667 0.620 4.00 0.500
2.67 0.691 1.33 0.822 5.33 0.592
4.00 0.874 2.00 0.948 6.67 0.660
The following rates of product formation (as µmol.s−1 .mg−1 ) were mea-
sured on the enzyme: Use the primary and secondary plot system to
estimate V and the other kinetic parameters of the two substrate rate
equation. Can you deduce anything about the possible mechanism of
this enzyme?
3.6 Summary 85
NAD+ , mM
isocitrate, mM 0.2 0.4 0.8 2
0.05 8.5 10.1 11.2 11.9
0.15 12.6 16.6 19.6 22.1
0.25 14.0 19.0 23.1 26.6
0.50 15.2 21.3 26.7 31.5
86 3 Enzyme activity: the molecular basis for its regulation
4
Traditional approaches to
metabolic regulation
87
88 4 Traditional approaches to metabolic regulation
of control and regulation. Of course, this assumes that we are able to identify
correctly the problems of constructing an efficient metabolism. The argument
applied in the search for regulatory enzymes is that it would be wasteful to
have ‘unnecessary’ metabolism, and if pathways were controlled near their
output ends, when they were slowed down, metabolites would still continue
to flow in at the input, causing an accumulation of metabolic intermediates.
Not only might this cause the expenditure of unnecessary metabolic energy,
it could also cause osmotic problems for the cell. Therefore, it would be much
better if a controlling enzyme were to be placed at the beginning of a pathway,
as suggested by Sir Hans Krebs.141
The argument seems seductively reasonable, and there are many appar-
ent examples in its favour. This is no doubt because the argument has been
influenced by the discovery of many enzymes with regulatory properties at
the points where a metabolic pathway branches off from the rest of metabol-
ism, as will be described later (Chapter 7.2). The problem is that there are
examples where the first enzyme in the pathway is not a regulatory enzyme
(e.g. mammalian serine metabolism, see Chapter 6.3.4.2, p. 193, in which the
regulatory enzyme is the last one in a three step pathway). In these cases, it is
explained that the regulatory enzyme is the first ‘committed step’. That is the
problem with teleological arguments: when they work, they seem reasonable;
when they don’t work, the story is changed slightly, and it’s now reasonable
that things should be different for this particular case.
A variant on this concept foreshadowed later developments: Rolleston 201
argued that the group of enzymes at the beginning of a pathway, including the
first non–equilibrium enzyme, could form a rate–limiting system, interacting
through substrate and coenzyme concentrations to form a functional unit.
10
.................................... ....................................
.................................... ....................................
10-4 .................................... ....................................
.................................... ....................................
.................................... ....................................
10-5 .................................... ....................................
.................................... ....................................
.................................... ....................................
10-6 ....................................
............
10-7
Ph Pg Tr He Gl Pf Pm En Py
Glycolytic reaction
perfused working rat heart, taken from experiments performed recently in the
laboratory of Veech and Passonneau in the United States.125 This study was
undertaken to examine the proposal that, in the absence of the hormone in-
sulin, transport of glucose into the heart muscle cells is the rate–limiting step
for glycolysis.162 Since insulin stimulates the movement of the glucose trans-
porter molecules from the endoplasmic reticulum to the plasma membrane,
it had also been held in the past that transport ceases to be ‘rate–limiting’
in the presence of insulin.170 The design of these recent experiments included
a quantitative analysis of the distribution of control by Metabolic Control
Analysis. Later in this book (Chapter 5.5.1), I will return to the results of
that analysis and how they relate to the disequilibrium ratios. My reason
for choosing this set of results as an illustration, even though they were not
used to find a rate–limiting step, is that they were undertaken with unusual
thoroughness. Thus all the equilibrium constants used in the calcuations have
been corrected to intracellular conditions, such as pH and the free cytoplas-
mic phosphate (which had been measured by 31 P NMR) and the free Mg2+
(which had been calculated from its effect on the citrate:isocitrate ratio).
The results show that three reactions, phosphorylase, hexokinase and
phosphofructokinase are very far from equilibrium, as is also pyruvate ki-
nase, though to a lesser degree. Phosphoglucomutase, glucose–6–phosphate
isomerase and phosphoglycerate mutase are close to equilibrium; indeed, the
measured disequilibrium ratio for the latter is greater than 1, though not sig-
nificantly so given the errors in the metabolite ratios. Glucose transport and
enolase are on the boundary between the two groups. (The enzymes between
aldolase and phosphoglycerate kinase could not have their disequilibrium ra-
tios measured. This is partly because aldolase binds sufficient glyceraldehyde–
3–phosphate to distort the results and also because 1,3–bis phosphoglycerate
is both difficult to measure and subject to binding by enzymes.)
In the past, some researchers attributed significance to differences in the
disequilibrium ratio at a step in different metabolic states. However, it is not
easy to draw any conclusions from such results, and the separate changes in
substrate and product concentrations were often thought to be more informa-
tive, as will be discussed in a later section of this chapter (4.6).
1.0 1.0
a) b)
0.8 0.8
Relative specific activity
0.4 0.4
0.2 0.2
0.0 0.0
0.0 1.0 2.0 3.0 4.0 5.0 0.0 1.0 2.0 3.0 4.0 5.0
Time Time
the example where the first and second intermediates are linked by a near–
equilibrium enzyme, the isotope is slower to accumulate in the first inter-
mediate because of the high rate of transfer to the second intermediate. The
amount of isotope in the second intermediate closely follows the amount in the
first, and isotope accumulates in this intermediate faster than it does when
the reaction is irreversible and non–equilibrium, as in the other example,
Fig. 4.2b). Computer fitting of results such as those in Fig. 4.2a) allows the
forward and reverse reaction rates at the equilibrium step to be determined,
as well as the pathway flux.
Many investigators have used such techniques to measure the fluxes through
near–equilibrium reactions, particularly those in the tricarboxylic acid cycle
and the glycolytic and gluconeogenic pathways in a wide range of cells and
tissues.10, 19, 58, 130, 192 In a study of the oxidation of [1–14 C]acetate by perfused
rat heart, Sir Philip Randle and colleagues in Bristol found192 that the specific
activities (relative to the acetate) of successive tricarboxylic acid cycle inter-
mediates after 60 seconds labelling were: acetyl CoA, 63%; citrate (in the 2
carbons derived from acetate), 43%; 2–oxoglutarate, 23 %, malate, 13.5% and
oxaloacetate (from the 4 carbons contained in citrate), 4.7%. This illustrates
the progressive spread of label through the intermediates in a similar manner
92 4 Traditional approaches to metabolic regulation
to the example shown in Fig. 4.2. The activities in glutamate and aspartate
were 20% and 6%, showing that these two intermediates were near equilibrium
with 2–oxoglutarate and oxaloacetate respectively (as in Fig. 4.2a) through
the action of transaminases. These conclusions were confirmed by computer
modelling of the results and estimation of the rates of the individual steps.
Fig. 4.3 shows the relative rates of the non–equilibrium steps and the
forward and reverse reactions of the near–equilibrium reactions of the tricar-
boxylic acid cycle in the slime mould Dictyostelium discoideum as determined
by Barbara Wright’s group130 from fitting experimental results on the spread
of 14 C.
citrate
1
0.42 0.42
w
oxaloacetate 2–oxoglutarate
±
U °
malate succinate
o 7
0.73 1.04
1.25 z 1.55
fumarate )
..............................
.......................
.......................
1000 .......................
.......................
.......................
....................... .......................
....................... ...............
................ ....................... .............................. .............. ...............................
100 ....................... .......
..................... ............... ............................... ............................... ............................ ...............................
V/Jgly
....................... . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
....................... ..................... ....................... ....................... ....................... ..................... ..............................
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
.. .. .. .. .. .. .. . ..................... ....................... ....................... ....................... ........ ..................... .......................
....................... ....................... ....... ....... .......
....................... ....................... ....................... ........ ............................ ............................... ............................... ............................... ............................... ............................ ...............................
10 ................ ....................... ....................... ....................... ....................... ..................... ....................... ....................... ....................... ....................... ..................... .......................
....................... ....................... .............. ....................... ....................... ....................... ..................... ....................... ....................... ....................... ....................... ..................... .......................
....................... ....................... ..................... ....................... ....................... ....................... ..................... ....................... ....................... ....................... ....................... ..................... .......................
....................... ....................... ..................... ....................... ....................... ....................... ..................... ....................... ....................... ....................... ....................... ..................... .......................
....................... ....................... ..................... ....................... ....................... ....................... ..................... ....................... ....................... ....................... ....................... ..................... .......................
....................... ....................... ..................... ....................... ....................... ....................... ..................... ....................... ....................... ....................... ....................... ..................... .......................
1
Ph Pg He Gl Pf Al Tr Ga Pg Pu En Py La
Glycolytic reaction
NADH end of the chain became more reduced relative to normal (i.e. there
was a build up of ‘substrates’ of the chain) whereas the carriers from the site
of inhibition towards oxygen became more oxidized (i.e. there was a reduc-
tion in the ‘substrate’).34 They then used this phenomenon of a cross-over
in the state of oxidation at the inhibition site to determine the sites of cou-
pling of electron transport to oxidative phosphorylation by comparing State
3 respiration (i.e. phosphorylating respiration in the presence of ADP, see
Chapter 6.1.4.2, p. 161) to State 4 respiration, which is inhibited by the lack
of ADP for phosphorylation. They formulated a number of cross-over the-
orems to guide the interpretation of these experiments35 and they validated
these theorems33 by computer simulations. The principal theorem was that a
cross–over from increased substrate (relatively more reduced carriers) to de-
creased substrate concentration (relatively more oxidized carriers) when the
pathway flux had decreased indicated the site of an inhibition. Subsidiary
theorems indicated that cross–overs in the opposite direction upon inhibition
were not significant, and dealt with identification of multiple sites of inhi-
bition (as obtained with the three coupling sites of electron transport upon
limitation of the phosphorylation rate).
Other investigators then applied the principle cross–over theorem to iden-
tifying the sites of action of regulatory signals, both inhibitors and activa-
200
150
Concentration, %
100
50
0
G6P F6P FBP DHAP GAP PGA PYR
Metabolite
Thus the cross–over plot cannot be regarded as reliable for metabolic path-
ways in general. In particular, the failure to observe a cross–over when there
has been a change in metabolic flux cannot be taken as proof that an effec-
tor did not act on an enzyme’s kinetics. Even when a cross–over has been
observed, as in the yeast glycolysis example above, the significance of the
finding needs careful thought. We shall see later in this chapter, section 4.8,
that it cannot be taken as proof that glycolysis has been accelerated by an
activation of phosphofructokinase alone, since experiments to increase that en-
zyme’s activity specifically and substantially have not resulted in any change
in glycolytic flux.
Sir Hans Krebs suggested in 1957141 that a site of regulatory action might
be identified by the observation of anomalous changes in the substrate con-
centration of a non–equilibrium enzyme when comparing two metabolic states
with different fluxes. That is, if the flux through the enzyme had decreased,
whereas its substrate concentration had increased (which would usually be
expected to cause an increase in the flux), then presumably the enzyme must
have been inhibited by some other factor. (Changes in product concentration
were rather summarily dismissed as a factor that could account for the change
in rate of a non–equilibrium enzyme on the grounds that its reverse reaction
rate is negligible; the inhibitory effect that a product can have even on an
irreversible enzyme reaction was, as usual, discounted.) This is effectively a
simplified version of the cross-over theorem and therefore shares some of its
limitations.
The thinking behind this assumption is that such mechanisms are not acci-
dental and must serve some function. This may well be true, though covalent
modification of enzymes by phosphorylation seems very widespread in eu-
karyotic cells, and it is not known in every case what effect this has on the
target enzyme — phosphofructokinase is an example of a glycolytic enzyme
that undergoes phosphorylation in many mammalian tissues, but in some of
these it is unclear what difference the phosphorylation makes to its proper-
ties. Furthermore, even if the molecular mechanism is brought into play in
some metabolic circumstances, this does not prove that the effect has had a
significant role in the situations under study. If an enzyme has its activity
changed by an effector, then the proof that this interaction is of regulatory
4.9 Conclusion 99
4.9 Conclusion
None of the lines of evidence described above has a single, unambiguous mean-
ing. In the search for regulatory enzymes and ‘rate–limiting steps’, it was
100 4 Traditional approaches to metabolic regulation
4.10 Summary
1. For most of the past half century, the main aim in the study of metabolic
control was to find the ‘rate–limiting’ enzyme of a pathway. Candidate
enzymes should show one or more of the following characteristics.
Further reading
1. Denton, R. M. & Pogson, C. I. Metabolic regulation, Chapman Hall,
London (1976).
6. Kwack, H. & Veech R. L., Citrate: its relation to free magnesium con-
centration and cellular energy, Curr. Top. Cell. Reg. 33, 185–207
(1992).
102 4 Traditional approaches to metabolic regulation
• The definition states that the rate–limiting step is the slowest step in a
pathway, but how is this to be interpreted? In a metabolic steady–state
(p. 10) all the steps along a linear pathway are going at the same rate. If
‘slowest step’ is interpreted as the one least able to go faster, the types
of observation described in Chapter 4 do not appear to measure this
inability.
• By the 1930s, it was known that the rate of a sequence of simple chem-
ical reactions could depend to varying degrees on the rate constants of
all the reactions. In 1964, Stephen Waley 260 showed that the rate of a
sequence of unsaturated enzymes (i.e. enzymes for which all the metabo-
103
104 5 Metabolic control analysis
xase % ydh
X0 −→ S1 −→ · · ·Y −→ S6 · · · → X1
&
source Jxase Jydh sink
∂Jydh .... c)
b)
.... .... .......... ∂lnJydh Jydh
...... ........................................
∂Exase =Cxase
.. . ∂lnExase
........ .... ..................
j
..
...
.......... .... ...
.............
...... j .... .......
........... .... . .................
................. ..... . ..
. .
Flux, Jydh
...............
ln(Flux)
.. ...................
...
.. .. .....
.
. . . ....
.. .
....
.
.... ....
. .
.
... .
.....
.
...
..
e e
Concentration of enzyme, Exase ln(Concentration of enzyme, Exase )
only requires measurements of flux and enzyme activity, which are standard
techniques in metabolic biochemistry. There were theoretical grounds, such
as Waley’s equation cited earlier, for expecting the relationship between flux
and enzyme activity to appear concave or hyperbolic as shown in Fig. 5.1b).
The few experimental studies available at the time were consistent with this
sort of relationship, and it has since been found in other investigations, some
of which are referred to later in this book. In addition, the increasing avail-
ability of computers had made possible the first attempts at simulating the
behaviour of sequences of enzymes, allowing theoretical investigation of their
behaviour.
The important feature of the dependence of flux on the concentration of
one particular enzyme (Fig. 5.1b) is that the response to a change in the
amount of enzyme will vary depending on the position of the starting point
of the change on the x-axis. Near the origin, there could be an almost pro-
portional increase in metabolic flux as the enzyme amount is increased; at
this point, there might be some justification for terming the enzyme ‘rate–
limiting’. If the starting point for increasing the enzyme is part way along
the x-axis, then the response of the flux to an increase in enzyme is notice-
able, but weaker than before. If the starting point is far to the right on the
x-axis, then an increase in the enzyme has a small or negligible effect on the
flux. There is continuous variation of the response of the flux to the amount
of enzyme between these extremes. Unless some suitable measurements have
already been made, we do not know where the content of a chosen enzyme in
a given cell type will map on this curve. Metabolic Control Analysis begins
by defining a coefficient that quantifies this variable response and then finds
means of measuring and using it.
5.2.1 Definition
In the pathway shown in Fig. 5.1a), suppose that a small change, δExase , is
made in the amount of enzyme Exase , and that this produces a small change,
δJydh , in the steady state pathway flux, J, measured at the step catalyzed
by ydh. If the change is made small enough, then the ratio δJydh /δExase
becomes equal to the slope of the tangent to the curve of Jydh against Exase
as shown in Fig. 5.1b). In mathematical notation this tangent is represented
as ∂Jydh /∂Exase . Obviously this represents the steepness of the response of
the flux to the amount of enzyme, but has the disadvantage that its numerical
value and its units will depend on the units used to measure the flux and the
enzyme. This problem can be avoided if we compare the fractional changes in
the enzyme and flux, i.e. δExase /Exase and δJydh /Jydh ; since the numerator
and denominator of each fraction are measured in the same units, the result is
dimensionless. (If multiplied by 100, each of these fractional changes can be
Jydh
regarded as percentage changes.) The flux control coefficient Cxase is given
5.2 Flux control coefficients 107
This can be rearranged so that the flux control coefficient is expressed as the
tangent to a curve such as that in Fig. 5.1b) times a scaling factor Exase /Jydh :
J
ydh ∂ ln Jydh
Cxase = (5.3)
∂ ln Exase
(Whether logarithms to base 10 or natural logarithms are used does not matter
as long as the same type are used for both axes.)
5.2.2 Interpretation
What values might we expect flux control coefficients to have, and what do
they signify? For the pathway shown in Fig. 5.1, typical values are shown
in Fig. 5.2. If the curve had flattened out so that flux could no longer be
increased by further addition of enzyme, as would have been seen if the graph
had been extended further to the right, the flux control coefficient would be
0. It therefore seems that the values can range between 0 and 1. A value of
1 would correspond to a proportional relationship between the pathway flux
and the amount of enzyme. (This can most easily be seen from the Eqn. [5.1],
108 5 Metabolic control analysis
..................
..
...
...
..................
..
............ ≈ 0.2
..
...
............
.
.
........
..
........
.
.
.....
.
......
.
.... ≈ 0.5
Flux, Jydh .....
..
...
.. .
.
..
....
.
...
...
.
..
... ≈ 1.0
Concentration of enzyme, Exase
which shows that in this case, a given fractional change in enzyme amount
produces an equal fractional change in the flux.)
It is also possible for a flux control coefficient to have a negative value. For
example, in a branched pathway where one metabolite can go to two different
end products, an increase in the activity of an enzyme in one branch might
increase the flux of the metabolite down its own branch, and thereby decrease
the amount available to flow down the other branch. Therefore the enzyme
could have a negative flux control coefficient on the flux in the other branch.
There is also a theoretical possibility that a flux control coefficient could
have a value greater than 1 in certain control structures found in metabolism
(e.g. Chap. 7.3.3, pp. 223 & 248). No experimental measurements have yet
demonstrated that such values can be found in metabolism but there is an
important piece of evidence from the genetics of diploid organisms that shows
that many enzymes have small flux control coefficients. Diploids generally
have two functioning copies in each cell of the gene for an enzyme, but if
this copy number is changed, the amount of enzyme is usually proportional
to the gene dose. In addition, there are many known examples of mutations
in genes for enzymes, both naturally occurring and artificially induced, that
result in the production of enzymically inactive protein. However, the impor-
tant finding is that these mutations are almost invariably recessive, that is, in
heterozygotes that contain one normal and one mutant copy of the gene, the
organism’s properties (or phenotype) appear normal. The normal gene is said
to be dominant to the mutant form. For example, in 1928, the geneticist R A
Fisher reported that 94% of mutations in the fruit fly Drosophila melanogaster
were recessive, 6% were semi–dominant, and none were dominant. Now en-
zymes affect the phenotype through the flux in metabolic pathways, so the
5.2 Flux control coefficients 109
observation that the wild–type is dominant effectively means that there has
been no discernable change in flux in the heterozygote, even though it contains
only half the enzyme present in the wild–type homozygote. In a number of
cases, it has indeed been confirmed by direct measurement that the metabolic
flux is only marginally reduced in heterozygotes. This is only possible if the
content of the enzyme in the normal homozygote is high enough for the flux–
enzyme curve to have flattened out (c.f. Fig. 5.2) to such an extent that
halving the enzyme content has little effect on flux, which is equivalent to
the flux control coefficient being much less than 1 and close to zero. Since
it is so common for mutations in enzymes to be recessive, the implication is
that most enzymes have small flux control coefficients. Furthermore, Kacser
& Burns121 also pointed out that the phenomenon of dominance has a simple
explanation in terms of the systems behaviour of biochemical pathways. Pre-
viously, dominance had been difficult to explain, and mechanisms such as the
one Fisher proposed assumed that it had been selected during the evolution
of the control mechanisms of the pathways. Recent experiments on artificial
diploids created from a normally haploid organism, the alga Chlamydomonas
reinhardtii, showed that these new diploids also exhibited dominance of the
wild type genes over mutants in 90% of cases177 . Since the recessive nature of
mutants in the diploid could not have been subject to evolution in this haploid
organism, the experiments are consistent with the interpretation devised by
Kacser & Burns, but not with the earlier idea that dominance was an evolved
property.
In the following subsection on the summation theorem, we shall see that
there are some further restrictions on the values that the flux control coeffi-
cients can have.
If the value of a flux control coefficient is known, approximate predictions
can be made about how the metabolic flux will change if the amount of enzyme
is changed. Such changes might be brought about by changes in genetic
expression or by covalent modification of inactive enzyme (Chap. 7.4, p. 229).
(Other methods of changing enzyme activity involve changing the catalytic
activity whilst the amount of enzyme is unchanged; the flux control coefficient
is also involved in predicting the response to these perturbations, as will be
explained in the section on the response coefficient, p. 124.) If the enzyme
concentration changes from Exase,1 to Exase,2 , by an amount small enough
Jydh
for Cxase to be effectively constant, Eqn. [5.2] can be integrated to give:
Jydh
∆ ln Jydh = Cxase ∆ ln Exase (5.4)
where ∆ ln Exase = ln Exase,2 − ln Exase,1 . That is, the change in the loga-
rithm of the flux is equal to the flux control coefficient times the change in
the logarithm of the amount of enzyme. As pointed out by Higgins104 this
corresponds to the power law form:
J = aE C (5.5)
where for simplicity, subscripts and superscripts have been omitted, so that
110 5 Metabolic control analysis
J
ydh
J = Jydh , E = Exase , C = Cxase , and where a is a constant that takes the
value needed to make the equation exactly true at the enzyme and flux values
for which the control coefficient was measured. Eqns. [5.4] and [5.5] both
describe the tangent in Fig. 5.1c), so any prediction made with the flux control
coefficient of the metabolic flux at other enzyme concentrations becomes less
and less accurate the further away the new enzyme content is from the one
at which the tangent was measured. What is less obvious is that the tangent
in Fig. 5.1c) is generally a curve on the linear scales of Fig. 5.1b), that lies
between the tangent shown on that graph and the true curve. Thus Eqns. 5.4
and 5.5 are always a better approximation to the flux–enzyme relationship
than the following linear equation (based on Eqn. [5.1]), which has often been
advocated:
∆Jydh Jydh ∆Exase
= Cxase . (5.6)
Jydh Exase
Although the equations above have limited predictive power for large
changes in enzyme content, in 1993 Rankin Small & Henrik Kacser233 de-
vised a surprisingly simple but improved predictor. Strictly, their result only
applies in cases where the flux–enzyme relationship, such as Fig. 5.1b), is rect-
angular hyperbolic (as seen in Michaelis–Menten enzyme kinetics; Chap. 3.2,
p. 47). That is to say, the flux and enzyme amount have to be related by an
equation of the form:
AExase
Jydh =
B + Exase
where A and B are constants; if this is the case, a plot of 1/Jydh v. 1/Exase
will be linear. As it happens, a number of experimentally–determined results
appear to obey this relationship, although there can be no guarantee that
this will always be true. Nevertheless, when it is true in a linear metabolic
pathway, measurement of the two values of the flux Jydh , J1 and J2 , at two
widely separated levels of the enzyme Exase , E1 and E2 , allows calculation of
the flux control coefficient at enzyme level E1 as:
J (J2 − J1 )E2
CE = (5.7)
(E2 − E1 )J2
This differs from the small change approximation to the flux control coefficient
given earlier as Eqn. [5.6] in that the weighting factor is E2 /J2 , and not E1 /J1 .
Oddly, if E1 /J1 is used as the weighting factor, the resulting control coefficient
is that at enzyme level E2 ! This equation had also been derived by Mark Stitt
in experiments on the control of photosynthesis that are described later in this
book.
Rankin Small and Henrik Kacser also noticed that, with these equations,
the flux control coefficient measured at one point allows prediction of the flux
at a markedly different enzyme concentration with good accuracy. Suppose it
is possible to increase the enzyme amount r times; the pathway flux will be
5.2 Flux control coefficients 111
This function is plotted in Fig. 5.3. What this shows is that by changing
the amount of a single enzyme, the effects on the pathway flux can be quite
limited, unless the flux control coefficient is greater than 0.5 to start with.
The maximum change in flux that can be obtained by a very considerable
increase in enzyme, when r À 1 (so that r ≈ r − 1), is:
1
f= ydh J
(5.9)
1 − Cxase
Jydh
Thus if Cxase = 0.5, the maximum flux that can be obtained is a factor of
Jydh
2.0 times greater. Only if Cxase is near to 1.0 are very large changes in flux
possible, but as mentioned previously, few measurements have been made of
enzymes with such large values for their flux control coefficients. This result
is very disappointing since it implies that biotechnologists who might have
hoped to increase significantly the flux through a chosen metabolic pathway
by engineering an increase in a single target enzyme, will rarely find it possible.
This limitation only applies, however, to increasing the metabolic flux; the
same restriction does not apply for a decrease in flux caused by a reduction
in the amount of active enzyme (i.e. r < 1), since this can always lead to a
substantial drop in flux.
Equation [5.8] and Fig. 5.3 also have important consequences for how cells
can control metabolic rate, and these will be discussed later (Chapter 8.1).
This relationship is known as the summation theorem for flux control coeffi-
cients, and in mathematical notation is expressed as:
n
X
CiJ = 1 (5.11)
i=1
Pn
where i=1 is the mathematical symbol for summing the n items indicated
by the following term. There are several ways of proving this theorem; an
112 5 Metabolic control analysis
10
Relative flux, f
6
50
20
4
10
5
2
2
0
0 0.2 0.4 0.6 0.8 1
Fig. 5.3: The relative change of flux for large changes in enzyme amount.
The increases in flux predicted by Eqn. [5.8] against the value of the flux control coef-
ficient. The degree of amplification, r, of the enzyme amount is shown on each curve.
The graph can be used in reverse to give a quick estimate of the control coefficient if
the relative change in flux has been measured for a known change in enzyme amount.
example is given in the Appendix to this chapter, More about flux control
coefficients, p. 130.
The summation theorem shows that the enzymes of the pathway can share
the control of flux. In a linear pathway consisting of enzymes with normal
kinetics (i.e. where substrates stimulate and products inhibit the reaction
rate), all the flux control coefficients must be zero or positive, so the maximum
value any enzyme’s coefficient could have is 1, when all the other enzymes
would necessarily have flux control coefficients of zero. In this case, this one
enzyme could be said to be ‘rate–limiting’, since a flux control coefficient
of 1 corresponds to a proportional relationship between the activity of an
enzyme and the pathway flux. The summation theorem shows that this is
not a necessary feature of the pathway, as it is also possible for some or all
of the enzymes to have values greater than zero but less than 1. Note that
the definition states that the sum is for the flux control coefficients of all
the enzymes in the metabolic system; for example, this potentially implies
the whole cell. In practice, we would expect a pathway flux to be influenced
mainly by enzymes in that pathway, and perhaps by a few closely–connected
pathways, and that distantly–connected enzymes would have little influence at
all. In other words, the flux control coefficients of hundreds, even thousands,
of enzymes in a cell on our chosen flux will be zero, so even though the control
of flux is shared, it is not shared evenly.
5.3 Elasticities 113
5.3 Elasticities
Although the last section has demonstrated the important conclusion that
the flux control coefficient of an enzyme is a system property that cannot be
related solely to the enzyme’s properties in isolation, there must of course
be links between the enzyme’s kinetic properties and its potential for flux
control. This can be illustrated by considering what happens in a pathway if
an enzyme activity is changed. For example, in the pathway:
114 5 Metabolic control analysis
Scheme 5.1
xase ydh zase
X0 −→ Y −→ Z −→ X1
suppose that an extra amount of ydh is added, to increase the rate of the
second step. What is the effect on the pathway? The increased amount of
ydh tends to lower the concentration of Y . The lower Y will:
The increased amount of ydh also tends to raise the concentration of Z. The
increased Z will:
In conclusion:
• The effects of the increased amount of ydh involve the relative sizes of
the responses of the enzymes to the pathway metabolites.
• The effects on the metabolites could tend to change the rates of neigh-
bouring enzymes to match the change in ydh.
This explanation has shown that the flux control coefficient of an enzyme
is likely to be linked to its kinetic responses to changed metabolite concen-
trations, as well as its ability to influence the concentrations of metabolites
in the pathway, a linkage that was shown first by Heinrich & Rapoport.99 To
be more specific about how the kinetics of enzymes relate to the flux control
coefficients, we need a means of describing their kinetic responses in a manner
that is compatible with the control coefficients. This measure is provided by
the elasticity coefficient.
a) ∂vxase .... b)
∂S .... ... ..........
..... ................................
∂lnvxase =εxase
............ ∂lnS S
. . ......
v .... ...... ....
.... ..
................
.... ...
..........
.
..... v .... .......
........... ..... ..............
. ..
Rate, vxase
. ............
ln (rate)
... .. ............
..... .
....
..
.
..
.....
.... .
....
... ......
..... ...
...
..
....
s s
Metabolite concentration, S ln (metabolite concentration, S )
∂vxase S
εxase
S = .
∂S vxase
(5.12)
∂ ln | vxase |
=
∂ ln S
(The term ln | vxase | in this equation means that the logarithm of the abso-
lute value of vxase is used in this equation. That is, if vxase is negative, as
will be seen in Fig. 5.8, its positive value is used because it is not possible
to take the logarithm of a negative number.) Elasticities have positive values
for metabolites that stimulate the rate of a reaction (substrates, activators)
and negative values for those, like products and inhibitors, that slow the reac-
tion. Obviously, there must be some link to the results from enzyme kinetics
described in Chap. 3. In fact, the typical range of values that elasticities
116 5 Metabolic control analysis
1.610
1.600
log v
S = 0.5 mM
1.590
1.580
-0.335 -0.315 -0.295 -0.275
log S
1.0 ...
...
...
... . ..... .... ..
... .... .... .... .... .... ...
0.8 .... .... ....
... ... . ....
... ....
... ...
... ... .
0.6 ... ..
v/V or εvS ... .
....
........
.. ......
0.4
.. .....
......
.. .......
.........
.. ...........
.
. ..............
0.2 ..................
.. ...........................
.......................
..
0.0 ..
0 2 4 6 8 10
Relative concentration, S/Km,S
Fig. 5.6: Elasticity of an irreversible enzyme with respect to its substrate at varying
substrate concentrations.
The enzyme is an irreversible single–substrate enzyme operating in the presence of
its product, according to the equation v = V(S/Km,S )/(1 + S/Km,S +P/Km,P ), i.e.
Eqn. [3.4] with Keq = ∞ and P/Km,P = 0.2. : εvS . — — — : fractional velocity,
v/V.
from a rate equation by use of differential calculus; if you are familiar with
this, you will find more details, including the result for the Michaelis–Menten
equation, in the Appendix to this chapter More about elasticities, p. 134.
Examples of the elasticity values obtained for some simple enzyme mech-
anisms are shown in the next few figures. Fig. 5.6 shows how the substrate
elasticity of an irreversible Michaelis–Menten enzyme varies with substrate
concentration; its value goes from 1 at substrate concentrations much below
Km to 0 as the enzyme approaches its maximal velocity. Note that the val-
ues have been calculated for the presence of a fixed concentration of product,
which binds at the active site even if there is no appreciable reverse reaction.
This inhibitory effect of a product, even with an irreversible enzyme, means
that the enzyme has an elasticity with respect to the product, as illustrated
in Fig. 5.7. The value is negative, going as low as -1 when the enzyme is
strongly inhibited, because increasing the product concentration reduces the
rate.
For enzyme reactions that are near to equilibrium, the elasticities with
respect to substrate and product are mostly determined by the degree of dis-
placement of the reaction from equilibrium rather than by the kinetic details
of the reaction, and this can lead to much larger elasticity values than we have
seen so far (as was first pointed out by Kacser and Burns,119 and later de-
118 5 Metabolic control analysis
veloped in more detail by Bert Groen and his colleagues90, 273 ). For example,
consider the reversible Michaelis–Menten equation for the conversion of S to
P , Eqn. [3.4] where v is the net rate (positive for formation of P ). Algebraic
calculation of the elasticity with respect to substrate by differentiation and
scaling of this equation gives:
1 S/Km,S
εvS = −
1 − ρ 1 + S/Km,S + P/Km,P
1 vf
= − (5.13)
1−ρ Vf
where ρ is the disequilibrium ratio (Eqn. [1.6]) and vf the total forward rate.
The first term on the right hand side depends on the degree of displacement
from equilibrium, going from 1 in the absence of product to infinity at equilib-
rium, and the second term represents the fractional saturation of the enzyme
with S, going from 0 in the absence of S to a maximum value of 1, giving the
overall result illustrated in Fig. 5.8. (Note that S À Km,S is not, in this case,
a sufficient condition for the saturation term to reach 1, because the value
of P counts, particularly if P À Km,P ; ‘saturated’ can be a vague concept
with a near–equilibrium reaction.) In many practical cases, the elasticity of a
near–equilibrium reaction will be given with sufficient accuracy by the term
0.6
...
...
....
0.3 ..... .
... ..... .....
..... ..... ..... ..... .....
..... ..... ..... ..... ..... ..... .....
0.0 ...
...
v/V or εvP
...
...
...
-0.3 ....
....
.....
.....
......
.........
-0.6 ...........
.............
................
.......................
....................................
.......
-0.9
2 4 6 8 10
Relative concentration, P/Km,P
Fig. 5.7: Elasticity of an irreversible enzyme with respect to its product at varying
product concentrations.
The enzyme is an irreversible single-substrate enzyme operating in the presence of
its product, according to the equation v = V(S/Km,S )/(1 + S/Km,S +P/Km,P ), i.e.
Eqn. [3.4] with Keq = ∞ and S/Km,S = 1. : εvP . — — — : fractional velocity,
v/V.
5.3 Elasticities 119
10 ... .... ..
...
... ..... ..... ....
... .. .....
... ..... ..... ...
... . .....
5 ... ...
.. ....
... ..
... ... ..
.........
... ...................................
20v/V or εvS . .. .........................................................................
0 .......... ...
..... ...
.....
... .....
... ....
. ...
-5 . ...
...
...
...
...
..
-10
0 1 2 3
Relative concentration, S/Km,S
Fig. 5.8: Elasticity of a reversible Michaelis Menten enzyme with respect to its
substrate at varying substrate concentrations.
The enzyme is a reversible single–substrate enzyme operating in the presence of its
product, according to Eqn. [3.4] with Km,S =1; P = 5; Km,P = 5; Keq = 10. The
elasticity is undefined at equilibrium (S = 0.5). : ε vS . — — : 20 × fractional
velocity, v/V.
1/(1 − ρ), which generates very large values as ρ approaches 1. The product
elasticity is given similarly by:
−ρ P/Km,P
εvP = −
1 − ρ 1 + S/Km,S + P/Km,P
−ρ vr
= − (5.14)
1 − ρ Vr
Where the saturation terms are negligible, it follows that the two elasticities
are related as:
εvS + εvP = 1 (5.15)
Groen90 also showed that similar results can be obtained with multisubstrate
rate equations having numerators of similar form to that of Eqn. [3.4].
Large elasticity values can also be obtained with allosteric enzymes, even
when (as is often the case) their reactions are not close to equilibrium. With
an allosteric enzyme that has cooperative kinetics with respect to a substrate,
the elasticity can be greater than 1, although it must be less than the Hill
coefficient (see Eqn. [3.11]). In fact, Herbert Sauro showed when he was
a research student with me that, if the enzyme is an irreversible enzyme
whose rate of reaction is proportional to its fractional saturation, Y , with the
substrate S, then:
εvS = (1 − Y )nh (5.16)
120 5 Metabolic control analysis
3.0
2.5
2.0
εvS or Y or h 1.5
1.0
0.5
0.0
0 5 10 15 20
In other words, the elasticity is similar to the Hill coefficient at low degrees of
fractional saturation, when Y is near zero, but becomes progressively smaller
as the degree of saturation increases. Since the Hill coefficient for an enzyme
showing positive cooperativity is 1 at low saturation, becomes perhaps as high
as 2 — 3 at mid–saturation, and then decreases to 1 again at high saturation,
the elasticity in the same range will start at 1, perhaps rise above 1 and
then fall away to zero (Fig. 5.9). Of course, if the Hill coefficient is greater
than 1, the enzyme will have a higher elasticity than it would if it were
non–cooperative at an equivalent degree of saturation. (For an irreversible
Michaelis–Menten enzyme, nh = 1, and the equation above gives the same
result for the elasticity as Fig. 5.6 and Eqn. [5.34] in the Appendix.)
These examples have shown how elasticity values can be obtained from
equations for enzyme kinetics. However, this can only be done for an enzyme
in a cell if:
• a complete kinetic equation is available that incorporates all the reac-
tants and effectors that occur in the cell and affect the enzyme;
• the intracellular concentrations of all these metabolites in the compart-
ment where the enzyme occurs are known, and
• all the kinetic parameters are known under intracellular conditions.
This is quite a demanding set of requirements, often difficult to satisfy because
5.3 Elasticities 121
of the patchy nature of the published results on enzymes. For instance, any
particular enzyme is likely only to have been studied in a few organisms, and
not necessarily in the one under study. Also, enzyme kinetics experiments
often do not report the right type of results. This is partly for reasons con-
sidered in more detail in the Appendix More about elasticities, p. 134. There
are also undiscovered reasons, examples of which have been given previously
(Chapter 4.4, p. 93). As an alternative to determination from kinetics, there
are methods for direct determination of elasticities in vivo (see Chapter 6.3,
p. 170).
This is equivalent to the previous equation, even though the sum is now formed
of all the n enzymes in our system; this is because the enzymes that are not
affected by S will have elasticities of zero and will therefore not contribute
to the sum; perhaps only the enzyme that produces the metabolite and the
one that consumes it will have non–zero elasticities. To see how this works in
practice, consider a short section of glycolysis:
Scheme 5.2
enolase pyruvate kinase
· · · −→ PEP −→ ···
hand side of Eqn. [5.18] for a single one of these metabolites is no longer equal to
zero, but the equations can be combined in pairs to give a zero result, with the side
effect that the number of connectivity equations is reduced by one.69 This technical
complication does not invalidate the general properties and uses of the connectivity
theorem.
J εydh
Y
Cxase =
εydh
Y − εxase
Y
J −εxase
Y
Cydh =
εydh
Y − εxase
Y
Thus we have shown that although the flux control coefficients are system
properties of the pathway, they are nevertheless explicable in terms of the
kinetic properties of the constituent enzymes.
This is not a special result for a simple pathway; for a three–enzyme linear
pathway there will be two intermediates, so the one summation theorem and
two connectivity theorems will give the three equations needed to express the
three flux control coefficients in terms of the elasticities. Whatever the length
of a linear pathway, the procedure will work, though this is not the case for
branched pathways or pathways containing certain types of cycle (such as
124 5 Metabolic control analysis
substrate cycles). For these pathways, my student Herbert Sauro and I69
proposed additional equations relating the flux control coefficients and the
relative fluxes through different parts of the system. These equations, known
as the branch–point and substrate cycle theorems, make the flux control coef-
ficients expressible in terms of the elasticities and the flux distribution in the
system. Since then, detailed mathematical analysis by other researchers (in
particular Christine Reder in Bordeaux) has shown that it is always possible
to write a soluble set of equations for any pathway, provided that the path-
way can reach a proper steady state. The branch–point equations will not be
described any further for the moment, but their use will be illustrated later in
an experimental study of oxidative phosphorylation in Chap. 6.3.3.1, p. 186.
state. The strength of the effect of the parameter P on, say, the flux Jydh is
defined in the same way as the control coefficient, except that it is generally
J
referred to as a response coefficient, RPydh , as originally named by Kacser and
119
Burns:
J ∂Jydh P
RPydh = . (5.21)
∂P Jydh
This equation for the response coefficient can also be reexpressed in the same
ways as those for the flux control coefficients, Eqns. [5.3–5.5] and has the same
graphical interpretation as shown in Fig. 5.1.
For the second stage of the argument, Kacser & Burns pointed out that,
if an external, constant metabolite P acts on the flux Jydh through being an
effector of the pathway enzyme xase, the response coefficient for the effect
of P is composed of the flux control coefficient with respect to xase and the
elasticity of xase with respect to P :
J J
ydh xase
RPydh = Cxase εP (5.22)
n
X
J
RP = CiJ εiP (5.23)
i=1
Contributions to the sum only come from the enzymes for which εiP is not
zero. This equation for the multi–site response coefficient does show the po-
tential value of control mechanisms in which an effector acts on more than
one site. In the case where the effector acts only on one enzyme, then it would
be rare for that enzyme to have a flux control coefficient as large as one; the
elasticity of the effector on the enzyme might be around 1, so overall, the re-
sponse coefficient is most likely to be less than 1, i.e. there will be less than a
1% change in flux for a 1% change in the effector concentration. If the effector
acts on several of the enzymes of the pathway that together account for most
of the flux control (i.e. their flux control coefficients add up nearly to 1), and
each of the elasticities is near 1, then the response coefficient can be close to
1. (If the effector acts cooperatively on allosteric enzymes, then the elasticity
terms could be greater than 1 in both cases considered.) Furthermore, there
is likely to be a further, hidden advantage in the multi–site response mode for
the activation of a metabolic pathway. Activating a metabolic pathway via an
effect on a single enzyme usually proves to be rapidly self–limiting, since even
if the enzyme has a relatively large flux control coefficient to start with, it
decreases, with control transferring to other enzymes, as the target enzyme is
activated. This was illustrated previously in connection with large changes in
enzyme amount in Fig. 5.3. At the other extreme, if an effector acted equally
on every enzyme in a pathway, the flux would continue to increase even for
large stimulations of activity without any changes in intermediate metabolite
concentrations or redistribution of control. (See the derivation of the sum-
mation theorem in the Appendix More about flux control coefficients at the
end of this chapter, p. 132, for a justification of this claim, and Chapter 8.1,
p. 263, for further consideration of its implications for metabolic control.) An
effector acting on several enzymes, each with a moderate flux control coeffi-
cient spaced along the pathway could be an adequate approximation to this
extreme case. Pure examples of this effect are difficult to give because there
are usually other control mechanisms that come into play as well. However, to
the extent that the nucleotide AMP can be considered an external effector of
glycogen catabolism in muscle, it activates phosphorylase b and phosphofruc-
tokinase. In the longer of the synthetic pathways for amino acids, there are
cases where the product amino acid (again, if it is appropriate to regard this
as an external metabolite) inhibits at more than one point along the sequence
(see Chap. 7.2.3, p. 210).
This equation for the multi–site response coefficient, Eqn. [5.23], shows
why we do not usually define response coefficients to metabolites that oc-
cur in the metabolic pathway under consideration. If P is in the pathway,
the equation contains the same terms as the connectivity relationship for P ,
128 5 Metabolic control analysis
J
Eqn. [5.18], which proves that the response coefficient in this case, RP , is zero.
The physical interpretation of this result is that if we attempt to change the
rate of the metabolic pathway by adding extra P , a metabolite in the path-
way, this extra material will flow away through the pathway, which will return
eventually to the steady state that existed before the addition, i.e. there are
no long–term consequences of adding a pathway metabolite. (The exception
to this is the case where the metabolite contains a conserved moiety that
cannot be metabolized away, but in this case, the connectivity relationship is
not equal to zero, but has a value that indicates the response to changing the
amount of the conserved material.)
Thus there is a tendency for the enzymes with the largest limiting rate to have
the smallest flux control coefficients and vice versa but, again, this cannot be
relied on to rank the enzymes according to their effects on the flux because
130 5 Metabolic control analysis
Km values and equilibrium constants enter the expressions. The equation has
not been applied in any real cases because these are always more complicated
(involving two substrate enzymes, for example), so it has no practical value
beyond demonstrating the unsuitability of this traditional criterion.
5.6 Summary
1. The qualitative categories of ‘rate–limiting’ and ‘not rate–limiting’are
replaced in Metabolic Control Analysis by a quantitative scale for the
influence of an enzyme on a metabolic flux: the flux control coefficient.
4. The connectivity theorem shows that the flux control coefficients of the
enzymes in a pathway have links to the elasticities, i.e. the system con-
trol properties can be related to the individual kinetic characteristics of
the enzymes.
2. The logarithms in Fig. 5.1c) and Eqn. [5.3] are the natural logarithms,
but logarithms to base 10 can be used without any alteration to the
equations.
3. The flux control coefficient has been shown here as defined relative to
the concentration or amount of an enzyme, as mentioned above. How-
ever there is the potential for complications in cases where the activity
of the enzyme does not relate directly to its concentration. Such con-
ditions are not thought to be particularly common, but the difficulties
can be avoided by a variation of the definition that was proposed by
Reinhart Heinrich and his colleagues, in which the coefficient is defined
with respect to some parameter of the enzyme that acts on the enzyme’s
activity. Suppose this parameter is k; then we define the flux control
coefficient as the ratio of the response of the flux to k relative to the
effect of k on the enzyme activity, i.e. :
Á
Jydh k∂Jydh k∂vxase
Cxase = (5.26)
Jydh ∂k vxase ∂k
in this book and the justification for the simple definition used in this
Chapter. There are possible exceptions, however, such as when the
enzyme concerned forms enzyme–enzyme complexes; then εxase k need
not necessarily equal 1 and the intrinsic flux control coefficient of the
step catalyzed by xase is not equal to the response coefficient of the flux
to the concentration of xase. In such cases, the definition in Eqn. [5.26]
above is preferable.
The summation theorem for flux control coefficients has been proved in
a number of different ways. The simplest to understand is probably the ar-
gument originally used by Kacser & Burns. Again, we have reverted to the
simple definition of the flux control coefficient relative to the amount of en-
zyme, Eqn. [5.2], on the assumption that for any enzyme xase, εxase Exase = 1.
Suppose that the amount of one of the enzymes in the pathway, say the first,
E1 , has its amount increased by a small fraction α = δE1 /E1 . Then according
to Eqn. [5.1], the change in flux is given by:
δJ J δE1 J
= CE = CE α
J 1
E1 1
S ∂S Exase ∂ ln S
Cxase = . = (5.28)
∂Exase S ∂ ln Exase
Being defined in the same way, they can be interpreted in the same way.
5.8 Appendix 2: Concentration control coefficients 133
where Sj represents any one of the variable metabolites of the pathway. This
reflects the result mentioned in the previous Appendix on the flux control
coefficients: simultaneously changing all the enzyme activities by the same
small fractional amount α has no effect on any metabolite concentration.
The connectivity relationship for concentration control coefficients is more
complex than for flux control coefficients272 in that it has one form when the
metabolite whose concentration is the subject of the control coefficients (say
A) is different from the one in the elasticities (say B):
n
X
CiA εiB = 0 (5.30)
i=1
As with the flux control coefficients, the form of the equation changes when
metabolites in conserved cycles are involved.213
Generally, less attention has been given to concentration control coeffi-
cients. However, in Reinhart Heinrich & Tom Rapoport’s original work99 on
control analysis, they were called elements of the control matrix and used to
derive expressions for the flux control coefficients by means of the relation-
ships:
m
X S
CiJi = 1+ εiSj Ci j
j=1
m
X S
CiJk = εkSj Ci j (5.32)
j=1
These equations actually show how the systemic response of the flux to a
modulation of an enzyme can be broken into its components. Thus, the
first one shows that the response of the flux through step i to modulation of
enzyme i is composed of a proportional change from the change in the amount
of enzyme (the ‘1’), on which are superimposed the changes in activity of the
enzyme because of the changes in each of the metabolites Sj , with each of
these effects calculated from the concentration control coefficient of enzyme i
on substrate Sj (to show how much the steady state concentration changes)
134 5 Metabolic control analysis
and the elasticity coefficient for the effect a change in substrate Sj has on the
activity of enzyme i. In the second equation, because the effect of a change
in enzyme i on the flux at k is sought, there is no term for a direct effect of
the change in the amount of enzyme.
On the other hand, the values of the elasticities must be valid for the
conditions in the metabolic system under study.
For these reasons, although enzyme kinetics could in principle supply the
information required as elasticities, the published information is frequently
not appropriate.
Km
εvS = (5.34)
Km + S
The final stage in the calculation of the required elasticity value is to insert
numerical values for the parameters in the equation, in this case only Km =
(0.75 mM), and the value for the concentration of the metabolite (0.5 mM),
to give 0.60 as obtained previously from the graph shown in Fig. 5.5.
The reason the results for most elasticities turn out to be simpler than
might be expected from the complexity of the functions to be differentiated is
because of the cancellations between numerator and denominator terms with
the scaling factor. It is possible to define an ‘elasticity calculus’ that simplifies
the problem of determining elasticities by avoiding generating terms that will
subsequently cancel. Most enzyme rate functions, F , have the form F = N/D,
where both N and D are functions of the metabolite concentration, S, for
which we want the elasticity.
S ∂F
εF
S =
F ∂S
SD ∂( N
D)
=
N µ∂S ¶
SD ∂N/∂S N ∂D/∂S
= −
N D D2
µ ¶
∂N/∂S ∂D/∂S
= S −
N D
136 5 Metabolic control analysis
Further reading
1. Kacser, H. and Burns, J. A. The control of flux. Symp. Soc. Exp. Biol.
27, 65–104 (1973); reprinted with modern notation and terminology as
Kacser, H., Burns, J. A. & Fell, D. A. The control of flux. Biochem.
Soc. Trans. 23, 341–366 (1995)
Problems
1. Suppose an enzyme in a pathway follows Michaelis-Menten kinetics with
V = 100 units and Km = 0.05 mM:
SV
v=
S + Km
SV
v=
S + Km (1 + I/Ki )
fumarate *
) malate
PFK Ald
· · · Fru-6-P −→ Fru-1,6bisP −→ DHAP + GAP · · ·
Measuring control
coefficients
The previous chapter presented the main outlines of the theory behind Metabolic
Control Analysis. However, up to this point I have only shown the potential
of the theory: it has defined possible quantitative measures of the effects of
changes in enzyme amounts or of pathway effectors on the rates of metabolic
pathways, and there is the expectation that such measurements will show that
control is distributed through the pathway (although it is possible for it all
to be located on one enzyme). Now comes the point where it is necessary to
show that:
• when these measurements are made, they support the claims made ear-
lier.
139
140 6 Measuring control coefficients
value for the flux control coefficient. Fortunately there are cases where finer
genetic control of enzymic activity can be obtained.
1 2 3 4
glutamate - ornithine - citrulline - arg–succ - arginine
100
100
a
b
80
80
Flux, %
Flux, %
60 60
40 40
20 20
0 0
0 20 40 60 80 100 0 20 40 60 80 100
with the least activity (4%) of this enzyme. Similarly, the hyperbolic curve
I have drawn through the results for the last step, argininosuccinate lyase,
corresponds to a flux control coefficient of 0.07 in the wild–type, increasing to
0.42 in the heterokaryon with 10% of the wild–type activity. Of the other two
enzymes examined, in step 1 acetyl–ornithine aminotransferase also appeared
to have a low flux control coefficient, around 0.06 in the wild–type organism,
whereas that for argininosuccinate synthase might have been as high as 0.2,
but was not very accurately determined. (Later experiments implied that
this enzyme’s flux control coefficient could not be so large.) Thus little of the
overall control of arginine synthesis can be accounted for by these enzymes
(given that the summation theorem implies that the remaining control to be
found is 1 minus the sum of the four flux control coefficients in the pathway).
However, the main importance of these experiments was the demonstration
that, just as Kacser & Burns had predicted, the influence of an enzyme on a
flux was not all or nothing, but variable depending on the activity level of the
enzyme.
differing activity levels. The various homozygotes and heterozygotes that can
be formed can give a number of different levels of enzymic activity, though
interpretation of the results can become complicated if the allozymes differ
in kinetic properties such as Km as well as their limiting rates. Middleton
& Kacser determined the effect of varying activities of alcohol dehydrogenase
(EC 1.1.1.1) in this way on the catabolism of ethanol in the fruit fly, Drosophila
melanogaster .159 They used three naturally-occurring alleles for the enzyme:
S; F, and Fd . (The letters refer to the rate of migration in electrophoresis:
S for slow and F for fast. Fd is enzymically similar to F but is expressed in
larger amounts.) They bred four homozygous varieties: SS, FF, Fd Fd and
NN, where NN is the null mutant. In addition, they obtained the heterozy-
gotes SN, FN and FS. The measurement of relative enzymic activities was
complicated because the allozymes differ in both the expressed limiting rate,
V , and the Km for ethanol. At low, fixed substrate concentration relative to
the Km values, the activity of a two substrate enzyme such as alcohol dehy-
drogenase will be proportional to V /(Kethanol Ki,NAD ), as can be derived from
Eqn. 3.5, p. 59. This expression can therefore be used to compare the relative
activities of the allozymes whilst taking into account the differences in the
kinetic parameters. Accordingly, Middleton & Kacser made kinetic measure-
ments on extracts of the adult fruit flies and calculated V /(Kethanol KNAD )
100
80
Flux, dpm·h-1
60
40
20
0
0 0.2 0.4 0.6 0.8 1
Fig. 6.3: The effect of alcohol dehydrogenase activity on the flux of ethanol
catabolism in Drosophila.
The genotypes giving the points are, from the left: NN; SN; FN; SS; FF, and F d Fd .
The points on the graph have been recalculated from the original results reported by
Middleton & Kacser,159 and a rectangular hyperbola fitted. The result for FS is not
shown as there is insufficient information to calculate its position on the ordinate.
144 6 Measuring control coefficients
100
60
40
20
0
0.001 0.01 0.1 1 10 100
β–galactosidase, %
for the different genotypes. (Strictly, KNAD is not the correct parameter, but
there was no significant difference in its value between the strains.) However,
in order to show their results in Fig. 6.3, I have recalculated their kinetic
parameters and calculated the term V /Kethanol as the variable reflecting the
relative activity. This does not change the basic appearance of the graph,
nor their conclusion that the enzyme’s flux control coefficient is very close to
zero. This is consistent with their results on the flies’ relative abilities to tol-
erate ethanol (which they would naturally encounter in rotting fruit). All the
enzyme–containing strains had equal tolerance to the lethal effects of ethanol,
i.e. they exhibited an indistinguishable phenotype; only the homozygous re-
cessive, NN, was intolerant. These results indicate that it is unlikely that the
polymorphism at the alcohol dehydrogenase locus is maintained through se-
lective pressure related to alcohol metabolism, at least in the adult fly. They
are also a good demonstration of why mutations in alcohol dehydrogenase are
recessive: although the heterozygotes do contain less enzymic activity, they
have an essentially unchanged flux through the pathway.
The same technique is feasible in a haploid organism if there is a sufficient
number of alleles of differing enzymic activity. Dykhuizen et al.61 studied the
effect of varying activity of the β–galactosidase (EC 3.2.1.23) of Escherichia
coli on its rate of catabolism of lactose, which they assessed by the relative
growth rates of competing strains in a chemostat, where the steady–state
6.1 Manipulation of enzyme activity 145
Mark Stitt and his colleagues, originally at the University of Bayreuth, but
now at Heidelberg, have been using a variety of approaches to estimate the
flux control coefficients for enzymes involved in the photosynthetic production
of sucrose and starch. Their studies using heterozygotes for various enzymes
are summarised in Table 6.1. Because the changes in enzyme activities in
these experiments were large, the flux control coefficients were calculated from
a difference equation that the authors had derived on the assumption of a
hyperbolic relationship between flux and enzyme content:
µ ¶
E2 J2
1−
J1 E1 J1
CE = µ ¶ (6.1)
1
J2 E2
1−
J1 E1
where the subscript 1 signifies the flux (J) and enzyme (E) levels at the first
point, and 2 those at the second. (This equation is equivalent to the ‘large
change’ equation derived later by Small & Kacser, Eqn. [5.8] and mentioned
in the previous chapter.)
change in the amount of the target enzyme. This can be simply because
diverting resources into synthesizing one enzyme in unusually large amounts
competes with the synthesis of other cellular components. Though this is not
likely to be aproblem for the small changes in enzyme amounts needed for
Control Analysis, the effect has been demonstrated in extreme cases. More
specifically though, the mechanisms that normally control the expression of
enzymes in a metabolic pathway may be activated by the increased amount
of one of the enzymes and cause a reduction in the synthesis of the others,
so that any observed change in flux cannot be ascribed solely to the enzyme
that has been added. Even if these problems are overcome, the experiments
can still have the same difficulty as classical genetic approaches: changes in
the gene dosage produce larger changes in enzyme amounts than we would
normally like for determining control coefficients.
In 1986, Heinisch96 overexpressed the allosteric glycolytic enzyme 6–phos-
phofructo–1–kinase in yeast cells about 3.5–fold, but observed no effect on
glycolytic flux to ethanol. This has come as rather a surprise to those who
believed biochemistry textbooks that state that phosphofructokinase is the
rate–limiting step of glycolysis on the basis of evidence such as that cited in
Chapter 4. Similar experiments with 8 of the 11 other glycolytic enzymes
have subsequently failed to show any changes in glycolytic rates either, even
though hexokinase activity, for example, was increased 13.9 fold.220 However,
Davies & Brindle59 showed in a similar experiment that although a 5–fold
excess of phosphofructokinase had no effect on anaerobic glycolysis, it did
stimulate anaerobic ethanol production in aerobic conditions. That is, the
glycolytic rate increased slightly, but the cells nullified the effects by reducing
the amount of pyruvate oxidized by the efficient aerobic route and getting rid
of the excess pyruvate as ethanol. Even so, if the flux control coefficient of
phosphofructokinase on glycolysis is calculated from their results using the
‘large enzyme change’ equation, Eqn. [5.8], it is found to be only about 0.3.
D–galactoside (IPTG). (The reason for using the tac promoter rather than
just the lac promoter is that higher levels of enzyme synthesis can be driven
by the artificial version.)
Their methodology was then adopted for Metabolic Control Analysis by
Ruijter, Postma and van Dam203 in the Netherlands in 1991. They were
studying the first step in the metabolism of glucose by E. coli, which is its
simultaneous transport across the membrane and phosphorylation to glucose–
6–phosphate. As in a number of bacterial transport systems, the energy source
and phosphate donor driving this linked transport and phosphorylation is the
glycolytic intermediate phosphoenol pyruvate, which gets converted to pyru-
vate in the process. This phosphoenol pyruvate:carbohydrate phosphotrans-
ferase system (PTS) consists of several components: two cytoplasmic proteins
that are involved in accepting the phosphate from phosphoenol pyruvate and
that are shared between the different transfer systems, and the carbohydrate–
specific permeases that are located in the membrane and have one or two sub-
units. The glucose–specific PTS has two subunits, and the degree of control
of transport on glucose metabolism was investigated by measuring the flux
control coefficient of one of these, the membrane–bound permease with the
elegant name enzyme IIGlc (EC 2.7.1.69). Ruijter et al. constructed a plas-
mid containing the structural gene for enzyme IIGlc under the control of a tac
promoter and inserted it in an E. coli strain that lacked any chromosomally-
coded glucose transport systems; in this way, only cells containing the plasmid
could grow on glucose, and when the plasmid gene was expressed at low levels,
the content of enzyme IIGlc was below wild type levels. On the other hand,
by varying the concentration of the inducer IPTG, they could vary the level
of IIGlc between 20 and 600% of its wild–type level with little effect on the
expression of other proteins of the system. Slight variations in IIGlc activity
near wild–type levels in the presence of excess glucose had little effect on the
rates of glucose oxidation and growth, so the control coefficients were very
low. At the lowest enzyme levels studied, both rates showed dependence on
the enzyme content in the same sort of quasi–hyperbolic manner we have seen
earlier in this chapter. Of course, one possible explanation of the low flux con-
trol coefficients at wild–type levels could have been that although transport
does have some influence on glucose metabolism and growth, this control is
exerted by one of the other PTS components. This was excluded by studying
the flux in the transport step alone by measuring the rate of accumulation and
phosphorylation of the non–metabolizable sugar methyl–α–glucoside; enzyme
IIGlc had flux control coefficients of about 0.6 on these fluxes at wild–type lev-
els and therefore has a larger effect than the other three components, whose
flux control coefficients together cannot exceed 0.4 by the summation theo-
rem. Of course, it should be mentioned that the measurements of the effect of
the glucose permease on growth and glucose oxidation were made in just the
conditions, the presence of excess glucose, that would be likely to minimize
its influence on the fluxes. Lower glucose concentrations, as in chemostats at
steady state, would almost certainly increase the flux control coefficients of
the transport system.
6.1 Manipulation of enzyme activity 149
1.0 1.0 4 4 4
4
0.8 0.8
0.6 0.6
Crubisco
Crubisco
J
J
0.4 0.4
a) b)
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
Rubisco, % wild-type level Rubisco, % wild-type level
1.0
0.8
0.6
Crubisco
J
0.4
c)
0.2
0.0
0 20 40 60 80 100
Rubisco, % wild-type level
in plants would change the rate of photosynthesis, because the flux control
coefficient had not been measured. In the experiments, tobacco plants were
transformed with an ‘antisense’ gene to rbcS, the gene for the nuclear–encoded
small subunit of rubisco. The transformed plants exhibited varying degrees
of reduced expression of the enzyme, with only minor changes in the contents
of other photosynthetic enzymes. The flux control coefficients of the enzyme
on photosynthesis in the leaves were determined by comparing the rate of
carbon fixation in leaf disks cut from wild–type and transformed plants under
varying conditions of light, CO2 and humidity. In spite of rubisco’s reputa-
tion as the ‘rate–limiting’ enzyme of photosynthesis, the experiments showed
that the maximum flux control coefficient was about 0.8 for the wild–type
range of enzyme contents; the highest values were only observed with strong
illumination, high humidity and low CO2 , whereas they fell to about 0.1 at
high levels of CO2 or low light intensity (Fig. 6.5), and were intermediate at
average levels of these environmental parameters. Thus there must generally
be other steps that are also contributing to the control of the rate of photo-
synthesis, and the influence of environmental factors is very strong. Given
the values of the flux control coefficients, Fig. 5.3, which shows the expected
change in flux for a change in enzyme activity, reveals that very substantial
improvements in rubisco activity would have to be made to produce modest
changes in photosynthetic flux under most conditions, at least for tobacco
plants.
the pathway for the catabolism of the amino acid tryptophan — tryptophan
2,3–dioxygenase (EC 1.13.11.11) — in the livers of rats. Tryptophan that has
been brought into the cells by a transporter in the plasma membrane is con-
verted by this enzyme into kynurenine, which is in turn further catabolized by
kynureninase and other enzymes. Hepatocytes (i.e. liver cells) isolated from
these treated rats retained the induced changes in enzyme amount, and the
differences in their rate of breakdown of tryptophan could be measured under
standardized conditions. The hyperbolic relationship between the amount of
tryptophan 2,3–dioxygenase and the catabolic flux is shown in Fig. 6.6. The
enzyme’s flux control coefficient varied from 0.75 in the rats on a normal
diet (at the lower end of the range of enzyme content) to 0.25 in the max-
imally induced state. We shall see in a later section that they managed to
identify where most of the remaining control could be found by other means
(Section 6.1.4.3, p. 166 & Section 6.3, p. 171).
I have already the mentioned experiments that estimated the control coef-
ficient of β–galactosidase on the growth rate of E. coli. Dykhuizen et al.61 also
carried out further experiments in which they used variable induction of the
lac operon by the gratuitous inducer IPTG to produce parallel changes in the
activities of the lactose permease and the β–galactosidase, since the genes are
in the same operon under the control of a single promoter. As in their experi-
ments mentioned previously (p. 144), in which the activity of β–galactosidase
was varied at a constant level of the permease, the rate of lactose metabol-
10
8
Tryptophan flux
0
0 20 40 60 80 100 120
100
75
Fitness, %
50
25
0
100
P, % 50 100
50
0 0 G, %
Fig. 6.7: Dependence of the growth of E. coli on variation of the permease and
β–galactosidase activities
The axes are expressed as percentages of metabolic flux and enzyme activities (P for
permease and G for galactosidase) obtained with maximal induction of the lac operon.
The experimental values at 100% permease activity are the same results as in Fig. 6.4.
The three–dimensional surface is that fitted by Dykhuizen et al.
ism was measured in terms of the relative growth rate in a chemostat under
conditions of lactose limitation. This time, however, they were attempting to
measure the flux control coefficient of the permease. Although in this case two
of the pathway enzymes changed simultaneously, the contribution from the
changes in β–galactosidase was known from the previous experiments. They
combined the results of the two experiments and fitted them to an equation
that assumed the metabolic flux depended on each enzyme in a rectangular
hyperbolic manner, as shown in Fig. 6.7. When the enzymes are only partly–
induced by the natural effect of the lactose in the chemostat, the two flux
control coefficients can be calculated to be Cperm = 0.53 for the permease
and Cβ−gal = 0.04 for the β–galactosidase. (These figures differ slightly from
those given by Dykhuizen et al. in their paper, but have been calculated from
their fitted function.) The coefficients at the maximally–induced level of ex-
pression are 0.11 and 0.004. One surprising conclusion from this numerical
analysis, which is qualitatively apparent in Fig. 6.7, is that the production
of permease would have to be increased 30–fold to reduce its flux control co-
efficient to the same level as that of the β–galactosidase, but the coordinate
expression of the proteins in the same operon prevents this happening. An-
other interesting feature is that, in these experiments, the lactose permease
had a relatively large flux control coefficient, but in the same organism, the
154 6 Measuring control coefficients
0.10
0.08
0.04
0.02
0.00
0 0.5 1 1.5 2 2.5 3
Enzyme activity
Fig. 6.8: Dependence of glycolytic flux in a rat liver homogenate on added en-
zymes.
The enzymes added were hexokinase ( ) and phosphofructokinase (4). The results are
those of Torres et al.248 with computed best–fit rectangular hyperbolas. The leftmost
point on each curve represents the original activity in the homogenate. The phospho-
fructokinase activity has been multiplied by a factor of 10 for display purposes. Titration
with glucose–6–phosphate isomerase gave no change in flux.
experiments by Ruijter et al. described above showed the glucose PTS system
had a negligible flux control coefficient203 on growth rate. No doubt a relevant
factor is that in the lactose utilization experiments, the bacteria were growing
in a chemostat on limiting lactose, whereas in the others, the bacteria were
growing in batch mode on excess glucose.
10.0
a) 1.0
8.0 b)
0.8
6.0
1/Flux
0.6
Flux
4.0
0.4
2.0 0.2
0.0 0.0
0 1 2 3 4 5 0 1 2 3 4 5
[Inhibitor] [Inhibitor]
1.0
0.8
Flux 0.6
0.4
0.2
0.0
0 1 2 3 4 5
[Inhibitor]
dicted for a rate–limiting enzyme (Fig. 6.12). On the other hand, aspartate
transaminase (which is involved in the shuttle between the mitochondria and
the cytoplasm during the conversion of pyruvate to phosphoenol pyruvate)
gave biphasic responses with the inhibitors aminoxyacetate and cycloserine
and was therefore classed as not rate–limiting.
Readers should not by now be surprised to learn that Metabolic Control
Analysts were not content with Rognstad’s use of inhibitors to divide enzymes
into the two classes of rate–limiting and not rate–limiting. However, perhaps
just as important is the question of whether the shapes of the plots are reliable
qualitative indicators for making this classification anyway.
6.1.4.1 Theory
The application of Metabolic Control Analysis to the problem was first tackled
by the group of researchers at the University of Amsterdam who have given
experimental and theoretical studies of Metabolic Control Analysis such a
great boost from the early 1980s onwards. This group included, amongst
others, Tager, van Dam, Groen, Westerhoff, Wanders and Meijer, some of
whom have already had their work mentioned in this chapter. Rognstad’s
proposal was analysed by Bert Groen et al. in 1982,90 both in general terms,
and with reference to his specific conclusion about the role of PEPCK in
158 6 Measuring control coefficients
gluconeogenesis.
1.0
T(angent)
0.8
0.6
Flux
Imax
0.4
0.2
0.0
0 0.5 1 1.5
[Inhibitor]
RIJ = Cxase
J
εxase
I (6.2)
Therefore, the flux control coefficient of the enzyme can be obtained from
the response of the metabolic flux to an inhibitor relative to the effect of the
inhibitor on the isolated enzyme:
J
Cxase = RIJ /εxase
I (6.3)
response coefficient and elasticity written out in full, for the limit (lim I→0 ) as
I approaches zero:
Áµ ¶
J ∂J I ∂vxase I
lim Cxase = lim (6.4)
I→0 I→0 ∂I J ∂I vxase
1.0
0.8
0.4
0.2
0.0
0 10 20 30 40 50 60 70 80
[Mercaptopicolinate], µM
J Ki
Cxase = −T (6.7)
J0
The conclusion from this is that the flux control coefficient of an enzyme
is related both to the initial slope of the inhibitor titration curve and the
strength of the inhibitor effect on the isolated enzyme. The shape of the titra-
tion curve does not seem to give a reliable indication. This became apparent
when Groen et al.90 used Rognstad’s data200 on the inhibition of gluconeoge-
nesis by the action of mercaptopicolinate on PEPCK (phosphoenol pyruvate
carboxykinase) and applied their equation. Whereas Rognstad had classed
the enzyme as rate–limiting because of the shape of the inhibition curve, they
calculated that the flux control coefficient was only 0.08. We shall see later
that this value was corroborated by an entirely different type of experiment.
Rognstad’s results are shown in Fig. 6.12. Also plotted is the expected inhi-
bition curve for PEPCK based on the reported inhibition constant of 3µM; as
can be seen, the initial slope of the enzyme inhibition curve is about 10 times
steeper than the flux inhibition curve, corresponding to a flux control coeffi-
cient of about 0.1 on the basis of the Kell & Westerhoff slope ratio method.
6.1 Manipulation of enzyme activity 161
µ
¡
¡
1 2 3 ¡4
succinate - - - ∆µ̃H+
@ 5
@
R
@
ATPm
@ 6
@
R
@
ATPc
@7
@
R
Glc6P
added NADH cannot be oxidized by animal mitochondria because it does not cross
the mitochondrial inner membrane and the NADH oxidase is only accessible from
the mitochondrial matrix. Therefore, to study the oxidation of NADH, it is neces-
sary to supply the mitochondria with metabolites that will be transported into the
mitochondrial matrix and be oxidized there by NAD+ –dependent dehydrogenases;
examples of suitable compounds include β–hydroxybutyrate, pyruvate plus malate
and glutamate plus malate.
The oxidation of the substrates by oxygen is carried out by the electron transport
chain, which is composed of a number of complexes of electron transport compo-
nents. Complex I is NADH:ubiquinone oxidoreductase (EC 1.6.5.3); Complex II is
succinate dehydrogenase (ubiquinone), EC 1.3.5.1. (The supposed succinate dehy-
drogenase, described by many textbooks, that reacts with FAD is merely a partial
reaction of the catalytic cycle of this complex.) The ubiquinol (or reduced Coen-
zyme Q) produced by these complexes is oxidized by Complex III, or ubiquinol–
cytochrome-c reductase (EC 1.10.2.2). Reduced cytochrome-c is oxidized with oxy-
gen by the cytochrome-c oxidase complex (EC 1.9.3.1) or Complex IV. In animal
mitochondria, Complexes I, III and IV are coupled to the phosphorylation of ADP,
but this does not imply, as previously assumed, that oxidation of an NADH gener-
ates 3 ATP molecules whereas oxidation of a succinate generates 2.
The linkage between oxidation and phosphorylation is described by Mitchell’s
chemiosmotic theory. The electron transport complexes coupled to phosphorylation
move protons out of the mitochondrial matrix across the inner mitochondrial mem-
brane, thereby creating both a concentration difference (or ∆pH) and a membrane
potential difference (or ∆ψ) that together constitute the protonmotive force (PMF
or ∆µ̃H + ). The PMF is the reservoir of energy that drives phosphorylation of ADP.
This reservoir is tapped by the H+ –transporting ATP synthase (EC 3.6.1.34), which
allows protons back into the mitochondrial matrix and captures their energy for the
formation of ATP from ADP and phosphate. The oxidation of NADH results in 4
protons being pumped out at Complex I, 2 at complex III and 4 at complex IV,
giving a total of 10 protons. It is thought that a total of 4 protons are needed to
take in ADP and phosphate from outside the mitochondria, turn them into ATP
and return the ATP outside the mitochondrion. Therefore, the P/O ratio, that is
the number of ATP molecules formed per oxygen atom used in oxidation of NADH
is 10/4 or 2.5. Since succinate is not oxidized via Complex I, its P/O ratio is 6/4
or 1.5.
Some of the PMF can dissipate without protons going through the ATP synthase.
The principal cause of this is the ‘leak’ reaction, which is most significant when
phosphorylation is going slowly and least significant when phosphorylation is rapid.
The ‘leak’ reaction can be stimulated by uncouplers, which increase the permeability
of the mitochondrial membrane to protons.
The phosphorylation process requires that ADP is transported into the mito-
chondrion and returned back to the outside as ATP. These two transport steps are
coupled as an exchange process catalyzed by the adenine nucleotide translocator.
There are two basic types of experiments performed on mitochondria in the study
of oxidative phosphorylation. In pulse experiments, mitochondria are supplied with
an oxidizable substrate and their rate of respiration measured. A known amount of
ADP is added; this causes a noticeable stimulation of respiration, (known as state 3
respiration) until all the ADP has been converted to ATP, when the respiration rate
returns close to its original rate (state 4). From the oxygen consumption profile,
it is possible to calculate the amount of oxygen used to phosphorylate the known
6.1 Manipulation of enzyme activity 163
amount of ADP. However, in this experiment, the steady states are only transient;
the mitochondria pass from zero phosphorylation to maximal phosphorylation rate
and back with relatively brief periods in the intermediate states. This is unlikely
to be similar to physiological conditions in most cell types. The alternative is to
set up a steady state phosphorylation rate. One way of doing this is to use the
enzyme hexokinase in the presence of glucose as a means of recycling the ATP back
to ADP (along with the formation of glucose–6–phosphate, which can be measured
to find out the phosphorylation rate) as shown in Fig. 6.13. By varying the amount
of hexokinase, the rate of phosphorylation, and hence oxygen consumption, can be
adjusted, and the steady state lasts until the oxygen runs out.
Most of the characteristics usually described in the textbooks relate to mam-
malian mitochondria, or more specifically, rat liver mitochondria. Experiments, in-
cluding Control Analysis studies, have been performed on mitochondria from other
sources, for example from yeast and from various plants, even though it is more
difficult to isolate the organelles succesfully from them. There are, however, some
significant differences between the details of the electron transport components and
coupling sites of these mitochondria from the animal ones.
Fig. 6.14: The effect of carboxyatractyloside on oxygen uptake in rat liver mito-
chondria.
Mitochondria were incubated with succinate, ATP, glucose and varying amounts of hex-
okinase to give different initial respiration rates, as measured by oxygen consumption
with an oxygen electrode. At each respiration rate, the system was titrated with in-
creasing amounts of inhibitor. Carboxyatractyloside does not cause complete inhibition
of respiration, since the leak reaction is responsible for some oxygen consumption when
phosphorylation has been completely inhibited. The I max value for the inhibitor was
estimated from the curves. This graph is reproduced from Groen’s PhD thesis. 89
lowest and highest respiration rates, but goes through a maximum of 0.52 at
intermediate rates. The step that comes closest to being rate limiting is the
leak reaction at the lowest respiration rates, where there is no phosphoryla-
tion and all the respiration is accounted for by the leak. However, the method
of calculation of this flux control coefficient from the activator titration has
been criticized, and so this flux control coefficient has almost certainly been
over–estimated in these experiments. Nevertheless, we shall see later that
a different method of performing this experiment has produced essentially
similar results. The results in Table 6.2 show that the measured flux control
coefficients account for 0.86 relative to the expected summation theorem total
of 1. Two steps have not been measured though: the succinate dehydroge-
nase (ubiquinone) (Complex II), and the proton–translocating ATP synthase.
These might account for the missing control, if indeed there is any missing as
opposed to some experimental error. Although Groen et al. did not measure
the flux control coefficient of the ATP synthase, they believed it to be low
because of the results of previous experiments on the inhibition of oxidative
phosphorylation by the ATP synthase inhibitor oligomycin.
Some very similar experiments carried out by Thierry Letellier and Jean–
Pierre Mazat in Bordeaux148 illustrate the application of the slope ratio
1.0
0.8
Flux control coefficient
0.6
0.4
0.2
0.0
0 20 40 60 80 100
method. Fig. 6.16 shows the inhibition curve of mitochondrial respiration with
pyruvate and malate as substrate, using the Complex IV inhibitor cyanide;
on the same scale is shown the inhibition of the ‘isolated’ Complex IV step
under the same conditions. The ratio of the slopes leads to a value of the flux
control coefficient for Complex IV of 0.2.
6.1.4.4 Problems
The versatility of the inhibitor titration as a method of determining flux con-
trol coefficients is demonstrated by the examples I have cited. Of course,
the method is capitalizing on a long history in biochemistry of discovering
and using inhibitors of enzymes, not least because inhibitors are potential
6.1 Manipulation of enzyme activity 167
100
80
Flux or rate, % 60
40
20
0
0 5 10 15 20
[KCN], µM
Fig. 6.16: Inhibitor titration of mitochondrial respiration flux and Complex IV rate
with cyanide.
The substrate for oxidation was pyruvate + malate. : inhibition of mitochondrial respi-
ration. The solid line is a fitted curve. 4: inhibition of complex IV activity. The dashed
line is calculated from the parameters that fitted the other points for mitochondrial
respiration. Data supplied by Letellier.148
drugs. In spite of these successes, there is a difficulty with the inhibitor titra-
tion, connected with the need to determine the initial slope at zero inhibitor
concentration of the flux against inhibitor curve. Examination of the experi-
mental results in Figs. 6.14 and 6.16 shows why this is not easy. The response
of the flux is often significantly curved, so there is not necessarily an initial
linear region. In any case, the experimental variance in the flux measurements
makes it difficult to tell whether the initial region is a concave curve, linear or
a convex curve, all of which are possible shapes. In addition, one form of the
method requires an estimate of Imax from the graph, and another requires the
initial slope of the enzyme rate against inhibitor graph. My student Rankin
Small228 found that these problems could cause bias and uncertainty in the
estimates of the flux control coefficients. Both he229 and Gellerich and his
colleagues81 proposed that it was better to use a computer fit to the curve.
The difficulties of doing this are made greater because the exact equation
governing the curves is generally not known, and would in most cases be
very complex if it were. Instead, approximate equations have been devised,
matched to the type and properties of the inhibitor, but computer simulation
tests do show that this approach does work better. The titrations of oxidative
phosphorylation shown in Fig. 6.16 were fitted in this way; the curve drawn
through the flux measurements was computer–fitted to Gellerich’s equation.
168 6 Measuring control coefficients
J
Where a range of values has been given for the flux control coefficient (C E ), this
is because the value varies with the experimental conditions.
a tetramethylene sulphoxide
A measure of the success of this approach is that the fitted curve predicts the
degree of inhibition of the isolated enzyme step, and this predicted inhibition
curve is the one drawn through the experimentally measured enzyme rate val-
ues on that same graph. In this example, there was relatively little difference
between the flux control coefficients obtained by the slope ratio method and
the computer fit, but the latter is regarded as more reliable.
molecular details; see Chap. 1.2.1) has to go before it can be regarded as truly
successful.
Many of the contributions from Reinhart Heinrich and his colleagues in
Berlin to the theory of Metabolic Control Analysis have been linked to to the
development of models of the metabolism of the human red blood cell from an
initial model with just the glycolytic pathway 193 through a series of versions to
one that includes the 2,3 bisphosphoglycerate bypass and, via the membrane
ATPase, ion fluxes and volume changes.97, 98, 194 Control coefficients for the
fluxes and metabolite concentrations have been calculated. Other models of
red cell metabolism, also developed in Berlin by Hermann–Georg Holzhütter’s
group, have included theoretical examination of the effects of inherited enzyme
deficiencies108, 109 and the regulation of the hexose monophosphate shunt.222
Of course, it is more reasonable to expect to be able to build a comprehensive
model of human erythrocyte metabolism because the cells contain no nulei or
mitochondria, so there is no nucleic acid and protein synthesis, nor oxidative
phosphorylation.
David Garfinkel, one of the pioneers of computer simulation of metabolism,
developed large models of catabolism in muscle cells. Kohn and colleagues
have applied Metabolic Control Analysis to some of these models.135, 136 How-
ever, it must be said that these models show one of the difficulties of large
computer models: once they reach a certain complexity, even if they are ex-
hibiting similar behaviour to the in vivo system, it can be almost as difficult
to understand why the model behaves as it does as it is for the real system.
A number of models of various aspects of photosynthesis have been pro-
posed to aid understanding of how its rate responds to environmental condi-
tions. Some of these have been subject to one or other types of sensitivity
analysis, including Metabolic Control Analysis.86, 87, 182, 276 Predictions made
by the models differ about the dependence of the control coefficients for the
photosynthetic flux on the conditions, but, like the various experimental stud-
ies reported earlier (e.g. Fig. 6.5), they show that whereas the flux control
coefficient for the CO2 –fixing enzyme, ribulose bisphosphate carboxylase, is
high in some conditions, in others, it declines and other steps have comparable
or higher coefficients.
on the flux to citrulline synthesis (as part of the urea cycle in rat hepa-
tocytes) was 0.96, relative to the control coefficient of the next enzyme
ornithine carbamoyltransferase, which had been determined by inhibitor
titration. (See Q. 1 at the end of this chapter.)
Of these three methods, the most informative is the third, but the control
coefficients obtained in this way are derived entirely indirectly, for the exper-
imental observations neither directly measure changes in flux as a function of
a known change in a single enzyme activity, nor relate the flux control coef-
ficients to another known flux control coefficient (as in the first method). In
addition, it involves certain assumptions, in particular:
• that all the relevant steps in the the metabolic system have been identi-
fied, and that all the significant influences of the metabolites on each of
172 6 Measuring control coefficients
lactate
mito
............................ NAD+ cytoplasm
..
0 ......................
....... ....
. NADH
pyruvate ............................1............... pyruvate
..........
.......
... ........
... ......
... .....
.....
2 .... ....
...
... ...
....
...
3
oxaloacetate ................................. oxaloacetate 5 ..
.
.
..
.... ...
...
. .
. .....
.
......
4 ..
...... . ..........
........ .
PEP
....
6 ....
...
..
6 .....
...
6 .....
..
..
6 .....
.
GAP ........................... ( DHAP )
... .........
..
7 .... .....................
..............
.....
.......
( FBP )
...
.
7 ....
.....
F6P
.....
8 ......
( G6P .. )
.
8 ....
....
glucose
these steps have been recognized. i.e. that all the non–zero elasticities
involved in the system are known.
• that the theorems of control analysis apply to the system. i.e. that
the metabolic system reaches a quasi steady–state (as defined in Chap.
1.4.1) and there are no features of the pathway that would require the
use of modified theorems or invalidate their application altogether. Such
problems are mostly beyond the scope of this book, though a brief ac-
count of some of them is given in the final Chapter 8.3, p. 280.
Provided that these conditions are satisfied, my research group and I showed
that it should be possible to carry out such analyses on any metabolic path-
way.69, 213 Still, in order to show that the assumptions are justified, it is best
if results obtained in this way are confirmed by direct measurement of one
of the control coefficients to check that similar results can be obtained by a
different method.
Before explaining some of the methods available to measure elasticities, I
will first give an example of how flux control coefficients have been determined
from them.
was possible to tell when a steady state had been reached because the glucose
concentration in the perifusate would become constant (after about 20 – 30
min). Some change in the conditions, such as substrate concentration, could
then be made and the cells allowed to reach a new steady state, so that as
many as 6 experiments could be performed on a single sample of cells. The
concentration of the glucose in the perifusate also indicated the gluconeogenic
flux.
Experiments were carried out in the presence of the hormone glucagon,
which stimulates the liver to form glucose, and in its absence. One of the
major differences between these two sets of conditions is that, in the presence
of glucagon, the return loop from phosphoenol pyruvate, catalyzed by the
glycolytic enzyme pyruvate kinase, is suppressed by inactivation of the enzyme
by phosphorylation (Chapter 7.4.3.1). In the absence of the hormone, some
of the phosphoenol pyruvate is recycled to pyruvate.
The complete pathway from lactate to glucose involves about 15 steps
(excluding transport of lactate and glucose into and out of the cell). To mea-
sure the elasticities of all these steps would have required measurement of all
the metabolites that affected them, and this would have been an enormous
amount of work. In addition, not all the metabolic intermediates are easily
measurable. Groen and his colleagues therefore reduced the size of the prob-
lem by grouping some of the reactions together and treating them as a unit.
This means that it is not necessary to know the metabolite concentrations
inside the group. The control coefficient of a group must be the sum of the
control coefficients of the component enzymes (for the summation theorem to
remain valid). It is also possible to imagine that the group has elasticities
to its overall substrate and overall product.69 The relationship of these elas-
ticities to the component elasticities of the individual steps is more complex,
though it can be calculated. However, if the group elasticities are going to be
measured experimentally, this does not matter. The gluconeogenesis pathway
was therefore simplified to a set of 7 or 8 groups of steps between intracel-
lular (cytoplasmic) pyruvate and external glucose (Fig. 6.17). Seven steps
are involved when glucagon is present because it suppresses the flux through
pyruvate kinase. The interconversion of lactate to pyruvate was discounted
since the belief is that this enzymic reaction is always close to equilibrium.
The 8 groups interconvert 6 internal metabolites, which are mitochondrial
pyruvate and oxaloacetate, cytoplasmic oxaloacetate, phosphoenol pyruvate,
glyceraldehyde–3–phosphate and fructose–6–phosphate. These metabolites
were measured either directly or, as in the case of the oxaloacetate concen-
trations, indirectly from the malate concentrations on the assumption the
malate dehydrogenase reaction is near equilibrium. The other metabolites es-
timated indirectly were glyceraldehyde–3–P (GAP) from dihydroxyacetone–
P (DHAP) measurements, assuming the triose phosphate isomerase reaction
was close to equilibrium, and fructose–6–P (F6P) from glucose–6–P (G6P),
assuming the glucose–6–phosphate isomerase reaction was near equilibrium.
By a variety of means, some of which will be described in the following
sections, the elasticities of the groups were determined from these metabolite
6.3 Control coefficients from elasticities 175
Their results for 5 mM lactate are summarized in Table 6.4; the results at 0.5
mM and 1 mM lactate gave a similar pattern but are not shown.
tabsize
The results are those reported by Groen et al.91 for rat hepatocytes incubated with
5 mM lactate and 0.5 mM pyruvate, in the presence and absence of the hormone
glucagon. The enzyme groups correspond to those shown in Fig. 6.17, where the
full enzyme names are given. The flux control coefficients inevitably add up to
1, allowing for rounding error, because the summation theorem was one of the
equations used in the calculation.
Their conclusions were that, in the presence of glucagon, the largest flux
control coefficient was that of pyruvate carboxylase (0.83 – 0.89); no other
enzyme (including phosphoenol pyruvate carboxykinase) had a control coef-
ficient above 0.1. In the absence of glucagon, the largest control coefficient
was still that of pyruvate carboxylase (0.51 in the presence of 5 mM lactate),
but there were a number of other significant values: pyruvate kinase, -0.17;
176 6 Measuring control coefficients
the enolase to phosphoglycerate kinase group, 0.29, and the triose phosphate
isomerase to fructose bisphosphatase group, 0.27. PEPCK showed very small
control coefficients under all conditions. Furthermore, independent validation
of the calculations is provided by the re–analysis of the inhibitor titrations of
PEPCK, quoted earlier in section 6.1.4, that suggested a flux control coeffi-
cient of at most 0.1. Also, the response of the pathway to pyruvate was used
as described in the previous section to calculate that the flux control coeffi-
cient of the block of reactions (groups 1–4) converting pyruvate to PEP was
0.65 in the absence of glucagon. This is in close agreement with the figures
in Table 6.4.
Now that the potential value of determining control coefficients from elas-
ticities has been illustrated, I will describe some of the methods used both by
Groen and his colleagues and other research groups to obtain elasticity values
experimentally.
The rate of enolase depends solely on the two glycolytic metabolites 2PG
(2-phosphoglycerate) and PEP (phosphoenol pyruvate). Mathematically we
state this as veno = f (2P G, P EP ), where f (. . .) means is a function of and
is used to indicate that, at this point, either we do not want to, or we cannot,
be more specific about what this relationship is. At steady state, veno = J,
where J is the glycolytic flux. In a control experiment, the flux Jc and the
concentrations 2P Gc and P EPc have to be measured, whilst in a parallel ex-
periment, the rate of glycolysis is perturbed in some way by a small amount,
for example by altering the input glucose level, and the new levels of flux, J e1
and concentrations, 2P Ge1 and P EPe1 measured at steady state. Mathemat-
ical theory for approximating the change in rate (∆J1 = Je1 − Jc ) in terms of
the changes in the metabolite concentrations gives the result:
∂veno ∂veno
∆J1 ≈ ∆2P G1 + ∆P EP1 (6.8)
∂2P G ∂P EP
Scaling this equation, by dividing throughout by Jc = veno and substituting
the definitions of the elasticities (Eqn. [5.12]) leads to:
1 2 3 4
3PG ................................... 2PG ............................... PEP ..................................... pyruvate ................................. lactate
... . ..... ... ....
+... .... .... ... ...
... .. .. ... .
..
. ...
..
.... .. ... .
. . . NADH NAD+
2,3BPG ADP
.... ATP
.
..... .......
...... 5 ......
........................
.
......... ......
.....
... ....
.
... ..
Glc6P glucose
then the pair of equations can in principle be solved for the two unknown elas-
ticities. It is important for the accuracy of the method that the two imposed
changes result in mathematically independent equations (i.e. ∆2P G1 /∆P EP1 6=
∆2P G2 /∆P EP2 ) otherwise the result will be dominated by the experimental
error. Theoretical analyses by Rankin Small228 and Christoph Giersch84 have
suggested that the only viable experimental strategy is to make one change
upstream of the enzyme under investigation, and one change downstream.
This was the form of the double modulation method Groen et al.91 used in
the experiment described in the previous section on the control of gluconeo-
genesis in hepatocytes. They determined the elasticities of the transport of
oxaloacetate by the malate/aspartate shuttle between mitochondria and cyto-
plasm (see Fig. 6.17) by changing the amount of the pathway substrate lactate
and by inhibiting PEPCK with mercaptopicolinate as the two perturbations.
The substrate and product concentrations, mitochondrial and cytoplasmic
oxaloacetate, had to be estimated and the flux from lactate measured.
Although the double modulation equation (Eqn. [6.9]) has often been pre-
sented as shown above, the approximation formula used is not the best. Since
the elasticity is given directly by the slope of a graph of the logarithm of the
178 6 Measuring control coefficients
∆ ln J1 ≈ εeno eno
2P G ∆ ln 2P G1 + εP EP ∆ ln P EP1 (6.11)
where ∆ ln J1 = ln Je1 −ln Jc . Furthermore, we can turn this into the basis of a
simple graphical analysis of the results of a modulation experiment. Dividing
through by ∆ ln J1 gives:
∆ ln 2P G1 ∆ ln P EP1
1 ≈ εeno
2P G + εeno
P EP
∆ ln J1 ∆ ln J1
In the limit as the change ∆ ln J1 tends to zero, the approximation becomes
the equality:
∂ ln 2P G ∂ ln P EP
1 = εeno
2P G + εeno
P EP (6.12)
∂ ln J ∂ ln J
The reason that this is particularly useful is that ∂ ln 2P G/∂ ln J is the slope
of a graph of ln 2P G against ln J, which can be obtained by modulating the
system by a series of varying perturbations and measuring 2P G and J each
time (e.g. Fig. 6.19). In the same experiments, if P EP is measured, then
the graph of ln P EP against ln J can also be plotted. The advantage of the
graphical analysis is that, by using a number of different sized modulations
and drawing a line or smooth curve through the results, the slopes can be more
precisely determined than by taking differences between a control and a single
modulated point. Thus, with an experiment of this type, the two elasticities
become the only unknown terms in Eqn. [6.12] above. Repeating the exper-
iment by applying a second type of modulation to varying degrees will give
two equations in the two unknowns. This analysis was originally proposed by
Rankin Small228, 230 and can be illustrated by applying it to some modulation
experiments performed by Christoph Giersch in Darmstadt.85 By perturbing
the in vitro model of a short section of glycolysis shown in Fig. 6.18, he ob-
tained the results plotted in Fig. 6.19. From the four slopes, the elasticities
of enolase to 2–phosphoglycerate and phosphoenol pyruvate can be calculated
as εeno eno
2P G = 2.58 and εP EP = -2.32.
2.1 2.0
a) b)
1.9
Log [PEP] or Log [2PG]
1.8
1.9
1.7
1.8
1.6
1.7 1.5
1.9 2.0 2.1 2.2 2.00 2.05 2.10 2.15
Log J Log J
Giersch’s analysis of the modulation results is more flexible and extensive than the
simple double modulation method, and I am grateful to him for providing me with
his results and for giving me permission to analyse them in a different way. The
first enzyme, phosphoglycerate mutase, was modulated by varying the concentration
of the obligatory cofactor 2,3 bisphosphoglycerate (Fig. 6.19a). For the second
modulation, the third enzyme, pyruvate kinase, was subjected to variations in the
concentration of its second substrate ADP (Fig. 6.19b). The four lines on the figure
give the coefficients in the following two equations:
1 = −0.182εeno eno
2P G − 0.632εP EP
1 = 2.59εeno eno
2P G + 2.44εP EP
volving the sink metabolite should be zero. Both of these cases apply in the three
enzyme pathway studied by Giersch. Taking the experiment shown in Fig. 6.19a),
the enzyme modulated to alter the flux was pyruvate kinase. However, flux changed
at all three enzymes. In the case of enolase, as we have seen, the change in flux was
related to the changes in the concentrations of its substrate and product. In the
case of phosphoglycerate mutase, however, only the concentration of its product,
2PG, changed, so the double modulation equation reduces to:
∂ ln 2P G
1 = εP GM
2P G (6.13)
∂ ln J
where ∂ ln 2P G/∂ ln J is the slope of the 2PG line in Fig. 6.19a), which is -0.182. It
follows that the elasticity, εP GM
2P G = -5.49. Similarly, in Fig. 6.19b), where the first
enzyme PGM is modulated by varying the concentration of its cofactor, the flux
through pyruvate kinase varies solely because of the change in the concentration
of phosphenol pyruvate. The slope of the PEP line in this figure, 2.44, is 1/εP K
P EP ,
PK
giving εP EP = 0.411.
The analysis of these two modulation experiments has, in this case, given all
the information that is needed to determine the flux control coefficients of the three
enzyme system. There are three equations involving the flux control coefficients: a
summation theorem and two connectivity theorem equations with respect to 2PG
and PEP respectively:
CPJ GM + Ceno
J
+ CPJ K = 1
CPJ GM εP GM J eno
2P G + Ceno ε2P G = 0
J eno J PK
Ceno εP EP + CP K εP EP = 0
Subsituting the values obtained above for the four elasticities allows calculation of
the three flux control coefficients as CPJ GM = 0.07, Ceno
J
= 0.14 and CPJ K = 0.79.
Table 6.5: Additional examples of the use of modulation methods for elasticity
measurements in Control Analysis.
These examples do not include the similar ‘top–down’ type of experiments described
in the next section. The examples above fall into two groups: those carried out on
hepatocyte metabolism by members of the Amsterdam group, and the photosyn-
thesis experiments by Mark Stitt and his coworkers.
L
µ
¡
¡4
1 2 3 ¡
succinate - - - ∆µ̃H+
C @ 5
@
R
@
ATPm
@ 6
@
R
@
ATPc
@7
P @
R
Glc6P
these problems was to look at them from the other end. Instead of starting
with the components of a pathway and seeking how to group these into a
small number of components (a bottom–up approach), they proposed taking
the pathway and choosing a point at which it could be broken into two or three
components. They called this method the top–down approach to emphasize
that it worked in the opposite direction. For example, a linear pathway might
be divided into two about a single metabolite that is the product of one block
and the substrate of another, i.e. :
Scheme 6.2
Block 1 Block 2
X0 −→ S1 −→ X1
As far as the theory of Metabolic Control Analysis is concerned, this does not
raise any difficulties. We have already seen that it is legitimate to have flux
control coefficients of groups of reactions that must, by reason of the summa-
tion theorem, be formed of the sums of the flux control coefficients of their
component steps. There is also no difficulty about applying the definition
of the elasticity to a group of reactions rather than to a single one,69, 116, 213
though we would almost certainly have to measure its value experimentally,
since we cannot state general relationships between the kinetics of the com-
ponents and the elasticity in the same way that we can relate the kinetics of
an individal enzyme to its elasticity. In addition, it is necessary that the two
blocks are independent of one another, that is, no metabolite from one block
is a substrate or effector of any of the steps in the other block apart from the
single intermediate metabolite.
6.3 Control coefficients from elasticities 183
1.0 1.0
a) b)
Flux control coefficient, CCJ
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
1.0
c)
Flux control coefficient, CLJ
0.8
0.6
0.4
0.2
0.0
0 20 40 60 80 100
Respiration rate, %
Fig. 6.21: Flux control coefficients for mitochondrial respiration rate as a function
of respiration rate.
The results of Hafner et al.94 are compared with the corresponding results of Groen
et al.92 a) The flux control coefficient of the respiratory chain (block C, Fig. 6.20).
Points from Hafner et al. are shown with the curve of Groen et al. for the dicarboxylate
carrier only (step 1 in block C, Fig. 6.20) taken from Fig. 6.15. b) The flux control
coefficient of the phosphorylation block (block P, Fig. 6.20). Again, the experimental
points are shown with the sum of the two curves through the the coefficients for the
adenine nucleotide translocator and hexokinase (steps 6 and 7, block P) from Fig. 6.15.
c) The flux control coefficient of the leak (block L). The experimental points are shown
with the corresponding curve from Fig. 6.15.
6.3 Control coefficients from elasticities 185
measured, i.e. the relative contributions of the leak and the phosphorylating
system to the consumption of the protonmotive force had to be determined. A
more detailed explanation of how this was done is given in the supplementary
material.
The elasticity of any one block with respect to the protonmotive force
was determined by inhibiting or stimulating one of the other blocks in order
to change the value of the protonmotive force. Thus if the elasticity of the
phosphorylating system was required, it was obtained from the slope of the
phosphorylation-coupled component of the respiration rate as the respiratory
chain block was titrated with the inhibitor malonate. Again this is covered in
more detail in the supplementary section.
and a branch point equation worked out according to a set of rules devised
by Herbert Sauro and me69
CPJC C JC
− L = 0. (6.16)
JP JL
6.3 Control coefficients from elasticities 187
1.0 1.0
a) b)
Flux control coefficient on JC
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
0 20 40 60 80 100 0 20 40 60 80 100
1.0
c)
Flux control coefficient on JL
0.5
0.0
-0.5
-1.0
0 20 40 60 80 100
Respiration rate, %
Fig. 6.22: Comparison of the distribution of control over the respiration, phos-
phorylation and leak fluxes in mitochondrial oxidative phosphorylation.
The results were derived by top–down control analysis of rat liver mitochondria. 94 a)
Control of the respiration flux, JC . Points from a single experiment from Fig. 6.21
summarized on one graph. b) Control of the phosphorylation flux, J P . One of the sets
of results from Fig. 6.21 recalculated. c) Control of the leak flux, J L , again recalculated
from results in Fig. 6.21. : flux control coefficients for the respiratory chain block. :
flux control coefficients for the phosphorylation block. 4: flux control coefficients for
the leak block. This figure is based on one by Brand. 20
188 6 Measuring control coefficients
The results obtained in this way by Hafner, Brown & Brand94 are shown
in Fig. 6.21. The range of the control coefficients on the respiration flux
between the minimum and maximum respiration rates was approximately:
CLJC = 0.9 – 0.0; CPJC = 0.0 – 0.5, and CC JC
= 0.1 – 0.5. The profiles between
these points resembles the results Groen et al.92 had obtained by inhibitor
titrations, although to see this, it is necessary to group these latter results
into the appropriate blocks. In order to do this, I used the smooth curves
drawn on the graph of their results (Fig. 6.15) and superimposed them on
Fig. 6.21. Under the circumstances, the degree of agreement between the
results of the Cambridge and Amsterdam groups is remarkably good, and it
can be taken as providing some independent validation of both experimental
methods: the inhibitor titration method, and the top–down control analysis
via block elasticities.
Once again, the results demonstrate that the distribution of the control
of flux varies according to the metabolic steady state. There is also another
feature of control that can be illustrated by further analysis of this system.
Because there are three fluxes in this branched system, we can ask whether
the distribution of control is the same for all three; we shall see that it is not.
Suppose that we wish to calculate the distribution of control over the flux in
the phosphorylation branch, JP . The flux control coefficients will be related
by a summation theorem, exactly the same as Eqn. [6.14] above, except with
JP instead of JC . The connectivity theorem equation will likewise be the
same as Eqn. [6.15] above. The third equation, the branch point equation, is,
however, different. It is now:
JP
CC C JP
+ L = 0. (6.17)
JC JL
Just as Eqn. [6.16], which applied to the respiratory chain flux JC , did not
JC
contain the control coefficient CC , the new equation does not contain the
control coefficient with respect to the phosphorylation block, CPJP . The sign
has also changed, because the two fluxes now involved have opposite orienta-
tions, in that JC produces the protonmotive force and JL uses it, whereas in
the previous case, both the fluxes related to utilization of protonmotive force.
The solutions of the set of three equations are consequently different, as can
be seen in Fig. 6.22b). Over most of the range of respiratory rates, the major-
ity of control of the phosphorylation flux lies in its own block. This top–down
experiment cannot tell the distribution of this control between the different
steps within the block. However, the share of the phosphorylation control ex-
erted by the adenine nucleotide translocator and the ATP utilization system
should be the same as their share of respiratory control as observed by Groen
et al. in Fig. 6.15. (This ratio of control is not affected by the change in the
reference flux.)
When the flux control coefficients are calculated for the control over the
leak flux, JL , a different pattern is seen again. Once again, the same summa-
tion and connectivity theorem equations apply, but the branch point equation
6.3 Control coefficients from elasticities 189
1 glucose −→ NADH −→ −→ −→
0.22
0.29
%
2 glucose −→ −→ −→ ∆µ̃H+
&
0.49
0.22
0.15 - 0.3 0 - 0.14
%
1&2 glucose −→ NADH −→ ∆µ̃H+
&
0.49
Top–down analysis is also being used by Patti Quant and her colleagues in
Cambridge to study the control of fatty acid utilization and ketone body for-
mation.153, 187, 188 In the case of ketogenesis, the traditional criteria for iden-
tifying rate–limiting enzymes had led some groups to favour carnitine palmi-
toyltransferase I and others to support hydroxymethylglutaryl CoA synthase.
The measurements of control coefficients so far show that their values vary
with metabolic state, but neither of these enzymes are fully rate–limiting in
any of the experiments. (Some of their results are presented in Problem 5 at
the end of this Chapter.)
6.3 Control coefficients from elasticities 191
factors. Furthermore, we know that the elasticity alone is not enough, its
influence on the flux control coefficient or response coefficient of the enzyme
has to be considered.
In spite of the admitted problems in getting reliable information for elas-
ticity calculations, there is another reason why it may still be worthwhile
attempting. This is that the results of Metabolic Control Analysis are not
necessarily particularly sensitive to inaccuracies in the information used in
the calculations either of the elasticities or, from them, of the flux control
coefficients. One reason can be seen by looking at the diagrams of elastic-
ity values against metabolite concentration (Figs. 5.6–5.9): for much of the
range, the elasticities are not changing very rapidly relative to substrate con-
centration, so errors in the substrate concentration (or equivalently, in the
corresponding Km value) do not cause errors of corresponding size in the
elasticity. Two cases where this is not so turn out to be rather different from
one another. One case is for enzymes near to equilibrium, where the elasticity
grows very rapidly as the metabolite concentrations tend to their equilibrium
values. Inaccuracy here is of little significance because, as summarized in
Chapter 5.5.1, Kacser & Burns showed119 that the flux control coefficients of
such enzymes tend to zero. The other concerns small inhibition elasticities;
studies with my coworkers Rankin Small and Simon Thomas have shown that
these are much more important.
In our work,228, 232, 246 they devised procedures for assessing the influence
that individual elasticity values have on the calculated values of the flux con-
trol coefficients. Some of our conclusions have been that:
• the values of the control coefficients are not equally sensitive to the
values of all the elasticities;
• the dependence of the value of any control coefficient on the value of
an elasticity follows a hyperbolic or inverse hyperbolic response, i.e.
the value of the control coefficient changes relatively rapidly with the
elasticity when the elasticity is small (much less than 1), but tends to a
constant value as the elasticity increases;
• this applies particularly to elasticities of near equilibrium reactions, be-
cause the elasticities are not only large in magnitude, but appear in
substrate–product pairs whose contributions tend to cancel, leading to
the tendency to small control coefficients mentioned above, and
• small feedback and product inhibition elasticities will tend to be par-
ticularly important, in that small changes in their values will have the
largest effects on the control coefficients. (See also Section 6.2, p. 169
for evidence from computer modelling that even weak product inhibition
effects are important in metabolism.)
The general conclusion is that using an approximate value for many of the
elasticities will not make a significant difference to the calculated flux control
coefficients, and the validity of this assumption can always be tested by the
6.3 Control coefficients from elasticities 193
6.3.4.2 Applications
Keith Snell drew my attention to the evidence that the control of the path-
way for synthesis of the amino acid serine in mammalian liver might be
rather unusual. The pathway branches off from the glycolytic intermedi-
ate 3–phosphoglycerate and involves three enzymes, phosphoglycerate dehy-
drogenase, phosphoserine aminotransferase, and phosphoserine phosphatase
catalyzing the sequence:
Scheme 6.3
3–phosphoglycerate −→ phosphohydroxypyruvate
−→ phosphoserine −→ serine
The first step is an NAD–dependent oxidation, the second a transamination
with glutamate as donor, and the third the hydrolysis of a phosphate group.
The pathway can be regarded as starting at 3–phosphoglycerate because the
flux to serine is so small relative to the glycolytic flux that the serine path-
way will have little effect on the 3–phosphoglycerate levels. What appeared
unusual about this pathway was firstly that published measurements of the
metabolite levels in rabbit liver suggested that the first two steps were nearer
to equilibrium than the final step, and secondly that the most obvious regu-
latory interaction known from previous enzyme kinetic studies was inhibition
of the third step, phosphoserine phosphatase, by the end–product serine.
At first sight, there seemed to be enough information to calculate the
elasticities of the enzymes and thus the flux control coefficients, but as so
often, not all the information was quite complete. For example, phosphohy-
droxypyruvate concentrations in liver are too low to be readily measurable;
however, since 3–phosphoglycerate and phosphoserine were nearly in equilib-
rium (taking into account the NAD+ , NADH, glutamate and 2–oxoglutarate
levels), the first two reactions could be combined as one near–equilibrium re-
action and its elasticities to 3–phosphoglycerate and phosphoserine calculated
from the displacement from equilibrium. The kinetic information on the other
step was also not ideal, in that the measurements had been made on chicken
liver and rat liver, not rabbit liver. However, we reanalysed the measurements
194 6 Measuring control coefficients
and showed that both sets were explained best by the serine inhibition’s being
of the uncompetitive type (Fig. 6.24). Although there were differences in the
Km values and inhibition parameters between the rat and chicken enzymes,
these were not so great as to give grounds for believing that the rabbit en-
zyme would be very different. With the equation for uncompetitive inhibition
and the parameters of the rat liver enzyme, we calculated the elasticities of
phosphoserine phosphatase to its substrate phosphoserine and the pathway
end–product, serine.
Thus the problem we analysed had become a two step pathway with a sin-
gle intermediate, phosphoserine, for which we had the elasticities. Examples
of the calculation method have already been given, so it will not be shown
here; Q. 2 in the Problems section asks you to calculate the solution. We
found71, 235 that, in normal rabbit liver where there is a relatively high biosyn-
thetic flux, phosphoserine phosphatase has the largest flux control coefficient
(0.97), whereas the first two enzymes are close to equilibrium and have negli-
gible flux control coefficients. Phosphoserine phosphatase is inhibited by the
pathway product serine, giving a calculated response coefficient of −0.63 for
the effect of serine on the pathway flux. In contrast, the response coefficient
of the flux to the pathway source, 3–phosphoglycerate, is about a tenth of this
value, so the flux is largely determined by the serine concentration through
100
80
Velocity, % maximum
60
40
20
0
0.0 0.1 0.2 0.3 0.4 0.5
[phosphoserine], mM
inhibition of the final step. This is an exception to the common pattern where
the first enzyme after a pathway branches off is usually the enzyme subject to
end–product inhibition. The strong control exerted by serine on the pathway
flux is dependent on the strange properties of uncompetitive inhibition. In
uncompetitive inhibition, the inhibitor has contradictory actions: it inhibits
by reducing the apparent limiting rate (like a non–competitive inhibitor), but
at the same time it reduces the Km of the enzyme for its substrate, which
would normally cause an activation. This has the result that the stimulatory
and inhibitory effects are almost in balance at low substrate concentrations,
but inhibition predominates as the substrate concentration increases. As can
be seen from the graph of phosphoserine phosphatase kinetics (Fig. 6.24), the
degree of inhibition is higher at high substrate concentrations than at low
ones. This is the exact opposite of competitive inhibition, but is well adapted
to feedback inhibition, since the general effect of inhibiting a pathway enzyme
is to increase the concentration of its substrate, which could potentially nullify
the action of a competitive inhibitor. In fact, at intracellular levels of serine
in rabbit liver, phosphoserine phosphatase is effectively almost saturated with
its substrate.
fluxes to the external levels of the amino acids to calculate the flux control
coefficients of the transport steps. These varied from about 0.25 for transport
of tyrosine and tryptophan at basal catabolic levels to 0.93 for phenylalanine
transport when catabolism was induced.
6.4 Summary
1. There is now a large and growing body of experimental measurements of
flux control coefficients in a range of different pathways and organisms.
Although the experiments require a good degree of accuracy in order
to obtain reasonably reliable values for flux control coefficients, they
are based on familiar experimental techniques from genetics, molecular
biology and biochemistry.
4. There are only a few cases where a flux control coefficient is very close to
1.0, and in some of these, in extreme rather than physiologically normal
conditions. Thus the control of a pathway by a single, rate–limiting
enzyme has been experimentally proved not to be the norm.
Further reading
1. Fell, D. A. Metabolic Control Analysis: a survey of theoretical and ex-
perimental developments. Biochem.J 286, 313–330 (1992)
Computer programs
Several programs are available to carry out simulation and metabolic con-
trol analysis of metabolic pathways. At the time of writing, the following
programs for PC compatible microcomputers are available over the Internet
by ‘anonymous FTP’ from my laboratory at bmsdarwin.brookes.ac.uk in the
directory pub/software/ibmpc:
Problems
1. Citrulline is synthesised in mitochondria by the reactions:
(a) NH+ −
4 + HCO3 + 2ATP −→ carbamoyl phosphate + 2ADP + Pi
1 2
3-phosphoglycerate −→ phosphoserine → serine
the elasticity of the first step, ε1pser , is -1.43 in the liver of rabbits on
a normal low protein diet. (The first step is actually catalyzed by two
enzymes, but the elasticity is the ’combined’ elasticity for them both,
so they can be treated as a single step.) The elasticity of the second
step, ε2pser , is 0.041. What are the flux control coefficients, C1J and C2J ,
of the two steps?
198 6 Measuring control coefficients
Make two log–log plots of the fluxes against [acetyl CoA]:[CoA] and
measure the slopes to determine the elasticities. Using these elasticity
measurements, determine the flux control coefficients of the acetyl CoA
producing block and the consuming block. (This calculation method
will not give exactly the same answer as that published by Quant and
colleagues using a different analysis method.188 )
200 6 Measuring control coefficients
7
Control structures in
metabolism
So far, we have considered the potential that an enzyme might have for con-
trolling a pathway without having given particular consideration to its place
in the metabolic network. However, there is further organization in metabol-
ism beyond the serial sequences of reactions that we call metabolic pathways.
In this chapter, we shall look at some of the features of these higher levels
of organization and their potential roles in control, both as conventionally
interpreted and in the light of Metabolic Control Analysis. Once again we
will find that there are aspects of the conventional views that do not stand
up to close analysis.
This is just the same simplification that is used in the ‘top–down’ method
of Metabolic Control Analysis considered in Chapter 6.3.3, p. 181, and as in
that method, the supply and demand steps can each be a composite of many
individual steps without changing the essential nature of the problem. For
example M might be an amino acid, in which case the supply step would
be its synthetic pathway and the demand step could be protein synthesis.
201
202 7 Control structures in metabolism
4
1
a)
3 0
or Csupp
or Cdem
-1
M
M
2
-2
Csupp
Cdem
J
J
-3
1
-4 b)
0 -5
0.1 0.7 1.3 1.9 2.5 -0.1 -0.7 -1.3 -1.9 -2.5
supp
εdem
M
εM
J εdemand
M 1
Csupply = = (7.1)
εdemand
M − εsupply
M
1+Q
J −εsupply
M Q
Cdemand = = (7.2)
εdemand
M − εsupply
M
1+Q
flux control distribution rather than their absolute values. (The minus sign
gives Q a positive value because εMsupply , the product inhibition elasticity, is
typically negative.) The control coefficients on the concentration M can be
derived in a similar fashion (see Chapter 5.8.1, p. 133) to give:
M 1 1 1
Csupply = = (7.3)
εdemand
M − εsupply
M
εdemand
M 1+Q
M −1 −1 1
Cdemand = = (7.4)
εdemand
M − εsupply
M
εdemand
M 1 + Q
M M
Obviously Csupply = -Cdemand , but it is not possible to express the concen-
tration control coefficients solely in terms of the elasticity ratio. In fact, the
magnitude of the concentration control coefficients are inversely proortional to
(εdemand
M − εsupply
M ), the sum of the magnitudes of the two elasticities (because
supply
εM is itself negative).
Why is this important? It is because there are many instances where
pathways exhibit large changes in flux accompanied by relatively very much
smaller changes in metabolite concentrations. The extreme example is the
changes in glycolytic flux in animal muscle on going from the resting to the
working state, which can be between 100 and 1000–fold, yet the changes in
the concentrations of glycolytic intermediates are trivial.24, 101, 107, 204 There
are significant advantages to metabolite homoeostasis during flux changes:
1. if pathway intermediates are also participants in other metabolic path-
ways that are not required to change in rate at the same time as the
pathways being controlled, it is evidently better to minimize distur-
bances in the common intermediate concentrations;
2. having to make large changes in intermediate levels slows down the rate
at which a pathway can respond to a control signal with a change in
flux,62 so faster responses can be made if metabolite concentrations are
kept as near–constant as possible, and
3. avoiding large concentration changes during large flux changes min-
imizes the possibility of sudden adverse changes in cellular osmotic
strength.6
If control is exercised on a metabolic pathway, effective control of the flux com-
bined with good homoeostasis of the intermediates requires that the pathway
element acted on by the control mechanisms must have a high flux control
coefficient (so it can exhibit a strong response to the effector signal) but low
concentration control coefficients, so that the response of the concentrations
to the same signal is small. The conditions that have to be met to achieve
this are different depending on whether control is exerted in the supply block
or the demand block.
J
In control by the supply block, a high flux control coefficient, Csupply , is
demand supply
obtained if εM is significantly greater than | εM |, corresponding to
204 7 Control structures in metabolism
a low value of Q (see Fig. 7.1a and Eqn. [7.1]). The concentration control
coefficient only becomes smaller than the flux control coefficient if εdemand
M is
greater than 1 (Fig. 7.1a and Eqn. [7.3]). The elasticities we are considering
are composite elasticities for a block of enzymes rather than those of single
enzymes. For single enzymes, elasticities lie mainly in the range -1 to 1 for
those showing simple kinetics, or from around -3 to 3 for allosteric enzymes,
except where the reaction is near to equilibrium (Chapter 5.3). The way
in which the elasticities of the component enzymes combine to give block
elasticities generally makes a block elasticity smaller in magnitude than any of
its components, and since the blocks we are considering are pathway segments,
they would not be near equilibrium. A small (negative) value for εsupply M
would not be difficult to obtain since εsupply
M is a product inhibition elasticity,
and these are usually smaller in magnitude than substrate elasticities such
as εdemand
M in the case of single enzymes. When we are dealing with a block
of enzymes, this difficulty tends to be even worse, since the overall product
inhibition elasticity of the block is expected to be much weaker than those
of the component enzymes (reflecting a rather poor backward transmission of
the signal from the metabolite M ). However, too small a value of | εsupply M |
(e.g. giving Q < 0.1) would not be an advantage, since the larger it is, the
smaller the concentration control coefficient (Eqn. [7.3]).
If the elasticity of the demand block could be dominated by the elastic-
ity of an enzyme with cooperative substrate kinetics, it might be as large
as 3. My investigations of this problem with my colleague Simon Thomas
have shown that this may indeed be possible, for example if the cooperative
enzyme is the first enzyme in the demand block. The cost of this, though,
is poorer metabolite homoeostasis within the demand block. Improving on
this probably requires unusual pathway structures not often seen in practice
(e.g. sequences of several cooperative enzymes). Thus in spite of the tra-
ditional view (Chapter 4.1) that control should be exerted at the beginning
of a pathway (i.e. in the supply block), it is not easy to see how this can be
implemented with the available components to give a well–regulated pathway,
i.e. one that shows good metabolite honoeostasis during flux changes.
What about control by demand? An experimental example where control
has been shown to be shifted to the demand end of the pathway is hepatocyte
energy metabolism (Chapter 6.3.3.2). The requirement for control by the
J
demand block is that Cdemand J
exceeds Csupply , which requires that | εsupply
M |
demand
should be significantly greater than εM , giving a high value for Q (see
M
Fig. 7.1b and Eqn. [7.2]). The concentration control coefficent, Cdemand , only
supply
becomes smaller than the flux control coefficient if εM is more negative
than -1 (Fig. 7.1b). A low value of εdemand
M will be obtained if the demand
block is approaching saturation with M . This will also make the flux in the
demand block relatively insensitive to changes in the level of M , which may
well be an advantage if there are other processes that can affect it. The large
supply
negative value required for εM might seem to be a problem, especially as
I have just explained that the product elasticity of a block of enzymes tends
7.2 Feedback inhibition 205
to be weaker than those of any of its component enzymes. This would create
an obstacle to efficient homoeostasis in the supply–demand sequence, were it
not that metabolism has a solution to this difficulty, as we shall see in the
next section.
If bacteria such as E. coli are growing on [14 C]glucose together with a nitro-
gen source, the amino acids incorporated into their proteins are radioactively
labelled showing that they are being synthesized from the glucose carbon. If
one of these amino acids is added to the medium, it was established in the mid
1950s that the synthesis of this amino acid is selectively discontinued. Part of
the mechanism of this effect has been found to involve specific repression of
the expression of the synthetic enzymes in the presence of end product, but
this cannot provide a complete explanation, since growing bacteria degrade
very little of their proteins and therefore repression cannot account for the
existing synthetic enzymes not making the amino acid after it has been added
to the medium. The other part of the mechanism was discovered in 1956
when Umbarger250 found that the first enzyme in the pathway to isoleucine,
threonine deaminase, is specifically and strongly inhibited by isoleucine. This
feedback inhibition by the end product of a pathway is highly specific since the
structurally similar amino acids leucine and valine are not effective inhibitors
of the enzyme.
In the same year, Yates and Pardee281 also reported feedback inhibition of
aspartate transcarbamylase in E. coli: this first step in the synthesis of pyrim-
idine nucleotides from aspartate, is inhibited by the end product CTP. Once
these two examples had been discovered, further ones followed and it became
clear that feedback loops onto the first irreversible enzyme of a pathway are
a common structural feature in metabolism. In Chapter 3.5, p. 71, we saw
that the unusual inhibition kinetics of these enzymes led to the recognition
that they form a special class of allosteric enzymes. Although it was clearly
established for some of the early examples, particularly in amino acid synthe-
sis, that the feedback does operate to slow the rate of the synthetic pathway,
in many other cases it is a presumption that the purpose of the feedback is
206 7 Control structures in metabolism
1.0
0.8
Flux control coefficient
b
0.6
0.4
0.2 a
0.0
-5 -4 -3 -2 -1 0
full pathway.
For the reasons given in the earlier section on supply and demand (section 7.1),
the first two effects are in fact inextricably linked. Thus enzymes subject
to inhibition are regulatory enzymes in the sense in which regulation was
defined in Chapter 1.1, p. 1; they are not, contrary to common belief, effective
control sites for changing the pathway flux because they have small flux control
coefficients.
Of course, part of the reluctance to accept this conclusion arises because
a regulatory enzyme subject to feedback inhibition is demonstrably inhibited,
and the degree of inhibition does change with metabolic circumstances (as
shown by cross-over plots that identify phosphofructokinase as a site of inhi-
bition in glycolysis, Fig. 4.5). In fact, this confers no special status on the
feedback–inhibited enzyme; every enzyme in the pathway has also changed in
rate under the same circumstances. The only difference is that the feedback–
inhibited enzyme had the potential to have a more dominant role in control,
but feedback inhibition has suppressed that and made it subordinate to the
steps that utilize the feedback metabolite.
........................................... ........
.............. ........X2
a)
. ........................ ª .................. .........X2 b)
ª.........
...... ... . . ..
....ª...... .. . .. 6 ... . .
. 6
. .
.. .... .. ....
.. . S .... . S
... ..... ......... 4 ..
.. ... ...
.. ª ..............
.
.... ......... 4
.
..... ...
4 ..... ......
...... ...... 4
... .
.. .... .
... .
. .
X0 .
1 . S1..
... ......
2 . S2... X0 .
1 . S1.
.... 2 ..S2...
.
.....
...... ..... .....
.. 3 3
.. ..... .......... .... .......
..
... .
.. S3... ... ...S3.
...ª .. ..... .. .....
...... .. .. ..
....... .... 5 .. ... 5 ..
.......... ..
................ ª ................................. ª ........
............................
........................................X1 X1
J1 ε12
<1 (7.5)
J2 ε22
Since the flux before the branch point (J1 ) is greater than that in the branch
(J2 ), this requires the product inhibition elasticity of S2 on its own branch
to be significantly stronger (as indicated by the elasticity) than its inhibition
of the common pathway. In fact, the equation can be interpreted as a re-
quirement that a given change in the end product, ∆S2 , should potentially
212 7 Control structures in metabolism
cause a bigger change in the rate of the branch than of the common pathway
(∆v2 > ∆v1 ) if an increase in S2 is not to cause a fall in S3 . (I say potentially
because these are the predictions for the isolated steps, as they are derived
from the elasticities, not for the whole system.) Of course, there is an equiv-
alent need for the other end product to inhibit its own branch more strongly
than the common pathway. Because of the velocity ratio in Eqn. [7.5] and the
relatively limited range of inhibition elasticity values, it seems that satisfying
both requirements simultaneously will be unlikely unless the two branch fluxes
have approximately similar shares of the common feed flux. However, several
of the observed molecular mechanisms to be described below may exist to en-
sure the correct relative relationships of the inhibition strengths. On the other
hand, incorrect operation of these systems is been observed experimentally in
cases where provision of an excess of one end–product (typically an amino
acid) results in growth inhibition by causing a deficiency in the synthesis of
another.
This requirement on the inhibition strengths has the incipient danger that
one of the end products might not inhibit the common pathway sufficiently
so that when its concentration is increased, first S1 and then the other end–
product start to accumulate excessively. This in fact is Savageau’s other
undesirable form of behaviour of the nested feedback inhibition pattern: there
is the possibility of the breakdown of the steady state if
J1 ε12
¿1 (7.6)
J2 ε22
Thus the more effective homoeostasis possible with nested inhibition can only
be reliably exploited where the relative inhibition strengths, represented by
the inhibition elasticities scaled by the relative fluxes, fall within a restricted
range. However, in spite of this apparent unreliability, nested feedback inhi-
bition structures are found in metabolism.
2 - S2 4 - X4
X0 1 - S1
3 - S3 5 - X5
pep + ery-4-P
.......
... ....ª .....
.....
DAHP ....
...... ...
..... ...
.... ..
..
shikimate ..
... ... ..
... ... ............. ..
... ... ª ...... ...
... ... .. .
.. ..
shikimate-3-P ... .
..... ... ...
. .
...... ......
.
...................... chorismate
. .
ª ...................... .... .... ........ ......
. ....
...
.. .
... ......... . .. ....
..... ...... ...... ........
.... ..anthranilate
. .... ......................
.. ........... ..................ª................................. .............................ª
prephenate
......
tryptophan........ ...
...
. .. .......... ....
. .
.. .
.. ..
..... phenylpyruvate .
. hydroxyphenylpyruvate ..
..
.
. ....
... ..
............
.. . ................ ......... ....
.
phenylalanine tyrosine
aspartate
................. 1 .........ª
ª ....................
...............
. ... .......
........ ... ... ......
.........
. aspartyl–P .....
. .
. ...
.. ..
.. . ..
..
.. . ... ..
..
.. aspartate semialdehyde
. ..
.. ..... ª.....
.... ....... .
...
.... ...
...
.. .........ª
....... .. ..
.... ....... ...
.. .
... .
..
........ . .... ........... ..
... ....... ..
.
..........
.. ...
...
. . . ..
.. .. . .
.....
. 2 ..
. .
... ... ...
. . dihydropicolinate .. ..
.. homoserine ...... . .
... ... .............. ........ .... ...
.. .. .......... ....... .. ..
. . ... . .
lysine threonine
Fig. 7.7: Nested inhibition with multiple enzymes in threonine and lysine synthesis.
Three aspartate kinase isoenzymes (step 1) occur in this pathway in E. coli., though the
one repressed by methionine (which is formed from homoserine) has not been shown
here. Step 2 is homoserine dehygrogenase.
.................................................................
.......................... ......................X2 ...
........
. .
. ...... ......
. ..
.
...... ª .
... .. ..
..
... .. 6 ..
. .... ª .. ...... ..
.. .. .. .
.. .
.. ....S4
... ..... ...
. .
.. ..
..... .... 4
.
. ... ...
X0 .
1 ..........S1 2 ..........S2... .............
.....
3
..... ..........
... S3.....
.. ...
... 5.
.
..
ª ......... ..
............................
X1
Fig. 7.8: Schematic diagram of the concerted form of nested feedback inhibition.
..................................................................
....................................................................................X2 ....
............................. .. .. .
....... ........ ..
..
................ ª
.... . . ..
..
.
........ . 6 ..
.. .. .. .
... ..
... S ...
ª
.. .... . ª .
..
...... ..........
4
...... ...
. .
.......... ....
4
....
.
.. . ....
X0 1 .........S1 2 .........S2... ..........
...... .....
.. 3
..
.. ..... ..........
.. .
... ... S3.....
...ª
...... .. ...
....... ... 5.
.......... ... .
................ ª ....................... ........
...............................................X1
tion, with the additonal property that the inhibition of the common enzyme
in the presence of both end–products is greater than that caused by either
separately. However, this greater inhibition is exactly that expected from the
combination of the effects of each when they act independently. For example,
if one caused one–half inhibition, and the other caused one–third, the expected
outcome for independent effects is the product of the fractional activities, i.e.
1/2 × 2/3 = 1/3, or two–thirds inhibition.
This was first suggested to be the case for the glutamine synthase (glut-
amate–ammonia ligase) of E. coli:
This allosteric enzyme has separate inhibitory sites for AMP and GMP, but
both of them together interact to cause greater inhibition than would be
expected. Similar synergestic inhibition is seen with the glutamine synthase
of Bacillus lichiniformis, which is only slightly inhibited by low concentrations
of glutamine, histidine or AMP, but almost completely inhibited by AMP plus
histidine or glutamine plus histidine.
A1 A2
..... .........
...... .. .
.....
........
........... ..
....... .......
...................................................
.............. .................
................ v2a
.........
.......
.
.... ...... .
v1 .... ......... v3
X0 - S1 S2 - X1
.......... ...
. ...... .
.
....... ....
.......... v .......
.................... 2b .............................
.... ....
....... ..............
............... ..........
.......... ........
...
........ ......
........ ....
...
A4 A3
favourable (i.e. exergonic) even though they link the same two metabolites in
opposite directions, and for this to be the case, at least one of the sets must
be coupled to some other exergonic process (for example, phosphorylation by
ATP). Hence the cyclic flow leads to no net change other than the dissipation
of energy by a net flux through the coupled process that drives the cycle. This
characteristic - the occurrence of two oppositely directed sets of reactions that
would operate to achieve no change other than dissipation of energy - is often
given as the definition of a substrate cycle126, 196, 237 and is the particular
justification for the term ‘futile cycle’.
The early evidence that cycling was not just a possibility but did actually
occur in cells under certain circumstances was provided by:
a) b)
............. .
.................v..............................
............... v ..................... .......
. .. ....... 2a ......
.. .... 1a ......
..
v1 v3 v2
X0 - S1 S2 - X1 X0 S1 - X1
........ ... ........ ...
........ v .
...
.. ........ v .....
.
.. ..
....................
....................2b .....................
...................1b
c) d)
X0
@ v1
@
.................v..............................
.
.................v..............................
.
....
....... 1a ...... @R
@ ....
....... 3a ......
.. ..
S1 S2 S1 S2
........ ........
........ v ..........
. ¡ ................ v3b ...................
.
.
....................
....................1b
¡ ........................
¡
¡
ª v2
X1
f)
e)
X0 X0 X2
@ v1 @ v1 v4 ¡
@ @ ¡
..............v.............................. ..............v..............................
@
R
@ ......
........ 3a ....... @R
@ ......
........ 3a ....... ¡
¡
ª
.. ..
S1 X2 S1 S2
........ . ........ .
¡ ................ v3b ................... ¡ ................ v3b ................... @
........................ ........................
¡ ¡ @
¡
¡
ª v2 ¡
¡
ª v2 v5 @
R
@
X1 X1 X3
1. The flux pattern in the network cannot be fully described as the com-
bination of the minimum number of linear paths needed to account for
the mass flows connecting the inputs and outputs. (This can apply to
larger, more complex networks like Fig. 7.11f.)
3. There is one step of the cycle that can be deleted in principle and still
leave a network capable of connecting the observed input fluxes to the
observed output fluxes. (This is the criterion that shows the cycle is
intrinsically unnecessary; it eliminates conserved cycles like Fig. 7.11c
and hypothetical oddities like Fig. 7.11d.)
c)
6
?
pep
ADP
Á @
@
GDP Y 5@ - ATP
CO2 @ ¡
k @
R
@ ¡
ª
7 pyruvate
¡ CO@
2
GTP 6 ¡ ATP @
R
¡
¡
¡ W
¡
ª ADP
oxaloacetate
1
³ PP
³³ q
P
Fig. 7.12: The three substrate cycles of glycolysis and gluconeogenesis.
a) The glucose:glucose–6–phosphate cycle catalyzed by hexokinase IV (1) and glucose–
6–phosphatase (2). b) the fructose–6–phosphate:fructose–1,6–bisphosphate cycle cat-
alyzed by phosphofructokinase (3) and fructose–1,6–bisphosphatase (4). c) the phos-
phoenolpyruvate:pyruvate cycle catalyzed by pyruvate kinase (5), pyruvate carboxylase
(6) and PEPCK (7).
ing the fate of the radioactive hydrogen isotope, tritium (3 H). This is because
certain of the reactions of the pathway specifically cause the loss of 3 H at-
tached to a particular carbon atom of the metabolites. For example, glucose–
6–phosphate isomerase causes the loss of 3 H from [2–3 H]glucose–6–phosphate
when it is converted to fructose–6–phosphate; since the 3 H enters water and
is greatly diluted, it is not re–incorporated when the reverse reaction occurs.
Therefore, if [2-3 H]glucose is supplied to mammalian liver or kidney cortex,
the loss of 3 H from the 2 position of the glucose can be used as a mea-
sure of cycling because it shows that the glucose has been phosphorylated to
glucose–6–phosphate (which rapidly equilibrates with fructose–6–phosphate,
losing the 3 H, because of the high rate of the glucose–6–phosphate isomerase
reaction), and then dephosphorylated again back to glucose. Because of the
other metabolic reactions occurring simultaneously, the relative loss of 3 H
compared with the 14 C content of a doubly–labelled glucose molecule gives a
better measure of the cycling. Interpretation of such experiments is difficult
and often controversial; the topic is discussed at length in some of the arti-
cles listed under Further reading at the end of the Chapter. Randomization
of 14 C label between different positions of the glucose molecule can also be
used. For example, the fructose–6–phosphate:fructose–1,6–bisphosphate cycle
can be monitored in this way since, when aldolase splits the latter into two
trioses, the fragments can be reassembled at opposite ends of the hexose from
their original position.
Measuring metabolic fluxes by monitoring the spread of isotope through
a pathway (Chapter 2.5) can give the rates of the component reactions of
the cycle. This is because the cycle can give rise to faster migration of the
label between intermediates than would be expected from the net pathway
flux, just as near–equilibrium reactions do (Fig. 4.2, Chapter 4.3). This ap-
proach has been used extensively by Jacob Blum’s group at Duke University,
particularly in connection with the measurement of substrate cycling in rat
hepatocytes.10, 58, 190 The experiments are much more informative than mon-
itoring isotope loss or rearrangement in a single compound, but involve much
more work, both experimentally and computationally.
The results from these measurements differ between the three cycles of gly-
colysis and gluconeogenesis in rat liver and hepatocytes and between different
metabolic states. In general, though, the highest relative rates of cycling are
observed with the glucose:glucose–6–phosphate cycle, for which the cycling
rate can be many times the net flux of glucose uptake or release.261 This is
consistent with the lack of any obvious mechanism that can suppress the cycle
by ensuring that the enzymes are not active simultaneously. In the liver of
some mammalian species (including rats, mice and humans) the hexokinase
is present as an isoenzyme (hexokinase IV, also commonly but somewhat
erroneously known as glucokinase51 ) that is competitively inhibited by the
reversible binding of a complex between a specific regulatory protein256, 257
and fructose–6–phosphate, the concentration of which varies in parallel with
glucose–6–phosphate. There is, however, no known regulatory mechanism
acting on the glucose–6–phosphatase (although it is spatially separated by its
7.3 Substrate or ‘futile’ cycles 223
7.3.4 Thermogenesis
It seems surprising that the generation of body heat by metabolism should
need any special explanation, since catabolism is characteristically exother-
mic. However, the coupling mechanisms of metabolism create a difficulty,
because part of the energy released during catabolic reactions is conserved
by the phosphorylation of ADP to ATP and the regulatory mechanisms of
catabolism ensure it slows down unless this ATP is hydrolyzed. The anal-
yses of the control of supply and demand pathways and feedback inhibition
presented earlier (Sections 7.1–7.2) imply that better homoeostasis of ATP
may be obtained if control by demand predominates over control by supply
(catabolism), and this expectation is supported by the results of the top–
down analysis of hepatocyte energy metabolism (Chapter 6.3.3.2). Therefore,
if there is a requirement for extra heat generation, either the demand for ATP
must be increased or the coupling of catabolism to ATP production must be
weakened. There is evidence that both approaches are adopted, and one of
the ways that the demand is increased is by substrate cycling.
Eric Newsholme and Bernard Crabtree suggested that the fructose-6–phos-
phate:fructose–1,6–bisphosphate cycle is used by bumble bees to raise the tem-
perature of their flight muscles in cold weather.168 They had made measure-
ments that showed that flight muscles of several species of bumble bees contain
unusually high levels of fructose–1,6–bisphosphatase, which is required for the
gluconeogenic pathway and is not usually present in significant amounts in a
highly glycolytic tissue like muscle. Other species of bee that have relatively
little of this enzyme are unable to fly at such low ambient temperatures (10 ◦ C)
as bumble bees. Even at these temperatures, bumble bees can maintain their
thoracic temperature at the 30◦ C necessary for them (and many other insects)
to fly. The hypothesis was supported by measurements made by a different
laboratory39, 40 that showed that cycling could be measured in bumble bee
muscles at low temperatures, but the cycling decreased with temperature and
had disappeared at 27◦ C. Furthermore, the cycling was prevented in flight,
probably by inhibition of the fructose bisphosphatase by the rise in Ca 2+ that
occurs on stimulation of the muscle. However, Newsholme and Crabtree also
calculated that the energy yield of this substrate cycle was not enough by itself
to account for the total heat generation, so there must be other mechanisms
as well.
This also seems to be true of another tissue for which substrate cycling
could make a contribution to thermogenesis: brown adipose tissue. Brown
adipose tissue is a form of adipose tissue in which the cells have more cyto-
plasm and mitochondria, and contain several small lipid droplets rather than
7.3 Substrate or ‘futile’ cycles 225
the single large one found in the more abundant white adipose tissue. Brown
adipose tissue is present in neonatal mammals, which generally are less ef-
fective than adults at temperature regulation, and in hibernating mammals.
In both cases, brown adipose tissue has a significant role in heat generation.
The triacylglycerol:fatty acid cycle is active in brown adipose tissue, though
calculations imply that it could only account for a small proportion of the
heat production by this tissue.168 The mitochondria of brown adipose tissue
contain specific mechanisms that uncouple respiration from ATP synthesis
and make a much larger contribution.
.........................
.............v =100....................
.
.........
. 2a ......
v1 =10 .. ..... v3 =10
X0 - S1 S2 - X1
........
....... ....
.......... v2b =90 ...............
......................................
Around 1970, Eric Newsholme and his coworkers in Oxford168, 169, 171 de-
veloped the proposal that substrate cycles could be devices for increasing the
sensitivity of the regulation of pathway flux by an effector acting on one or
both of the pathway enzymes. The initial form of their argument169, 171 is
shown in Fig. 7.13; generalized algebraically, it led to the proposal that an
effector that makes a given percentage change in the activity of enzyme 2a
will cause a percentage change in the net flux (= v1 = v3 ) that will be larger
by a factor (1 + v2b /v1 ) than would occur if the pathway were linear with
v2b = 0.168 Similarly, the response of the net flux to a change in the activ-
ity of enzyme 2b would be amplified by the factor −v2b /v1 (with the minus
sign showing that an increase in the enzyme activity produces a decrease in
the net flux). If the same effector acted to stimulate one cycle enzyme and
to inhibit the other, these effects would be additive. This could be particu-
larly relevant to the fructose–6–phosphate:fructose–1,6–bisphosphate cycle in
which AMP and fructose–2,6–bisphosphate activate phosphofructokinase but
inhibit fructose–1,6–bisphosphatase.
Stein and Blum240 pointed out that the basis of the theory was paradox-
ical since it was derived on the assumption that the concentrations of the
226 7 Control structures in metabolism
(The bars around the terms indicate that only their magnitude is considered;
any negative signs, such as in the product elasticity ε11 are ignored.) Here
7.3 Substrate or ‘futile’ cycles 227
the first term on the left shows that the product inhibition of 2a must be
much weaker than the elasticity of 3 with respect to its substrate, and the
second that the substrate elasticity of 2a must be much smaller in magnitude
than the product inhibition of 1 (a condition most likely to be met when 2a
is saturated, as noted by Crabtree54 ). These same conditions also maximize
the magnitude of the negative flux control coefficient of the reverse limb of
the cycle, 2b.
In conclusion, the conditions under which substrate cycles can generate
greatly increased flux control coefficients (which are necessary to give in-
creased responsiveness to effectors of the cycle enzymes) are quite restric-
tive. Also, an effector that acts to increase the net flux through the cycle
also reduces the flux control coefficients of the cycle enzymes, and therefore
its own response coefficient. At present, there is no indisputable example
of the phenomenon; the best candidate is the fructose–6–phosphate:fructose–
1,6–bisphosphate cycle in mammalian liver glycolysis/gluconeogenesis, but it
is not certain that the cycling rates are high enough and the relevant elasticity
values are unknown.
the protonmotive force and export from the matrix by a Ca2+ /2H+ antiport
(i.e. exchanger) driven by the pH gradient component of the protonmotive
force, so both components are exergonic and effectively unidirectional.
The problem with the concept that buffering is a specific property of sub-
strate cycles is revealed by considering the Ca2+ cycle, even though this is
strictly a conserved cycle (type Fig. 7.11c) if Ca2+ exchange with the en-
doplasmic reticulum12 is ignored. If we take the cytoplasmic Ca2+ , then
its steady–state value is that which makes the rate of the supply process
(mitochondrial export) equal the rate of the demand process (mitochondrial
import). This is true even though the supply and demand processes meet
in the mitochondrial matrix to form a cycle, and is in fact a general state-
ment about any metabolite at steady state. Any buffering therefore reflects
the dynamic equilibrium, or non–equilibrium steady state, determined by the
specific kinetics of the process. Several conclusions follow:
• Any buffering effect that does exist cannot be a specific function of a
substrate cycle.
• There is no specific effect of cycling rate, since the rates of the demand
and supply processes can be multiplied by any arbitary factor and the
same steady state concentration is obtained.
• The question of homoeostasis of the concentration can be addressed
by the supply–demand analysis of Hofmeyr and Cornish–Bowden that I
described at the beginning of this Chapter in section 7.1. The sensitivity
of the control is represented by the concentration control coefficient; the
kinetic features of the pathway, but not its flux, will determine whether
the sensitivity is high or low.
• If buffering is meant to imply the tendency of the system to return
to the same steady state after an arbitary perturbation of the concen-
trations, then this is a common property of metabolic systems with
a dynamically–stable steady–state, and again the cycling is irrelevant.
However, the rate of return to the steady state after a perturbation is
related to the turnover rate of the intermediates, and will be shorter if
the flux is high.
Another demonstration that substrate cycling does not affect the ho-
moeostasis of metabolites follows from the Metabolic Control Analysis ap-
plied to the substrate cycle of Fig. 7.13 in Section 7.3.5 and Appendix 1 of
this Chapter. I have used the same technique that shows that the flux control
coefficients depend on the cycling rate to show that the concentration control
coefficients do not.
Is there no truth in the concept that the glucose:glucose–6–phosphate cy-
cle helps to regulate blood glucose concentration? I believe there is, and the
computer simulations carried out by Stein and Blum240 suggested it worked.
The explanation cannot be given just in terms of the effect arising from the
cycle. Firstly, if blood glucose was determined solely by the balance of supply
7.4 Regulation by covalent modification of enzymes 229
by absorption from the intestine and demand by the tissues, it would exhibit
poor homoeostasis because the inhibition of absorption by blood glucose is
too weak (Section 7.1). Through the glucose:glucose–6–phosphate cycle, blood
glucose is connected to other supply and demand processes, and the home-
ostatic characteristics are now those of the total system. The homoeostasis
is therefore related to the switching properties of the cycle, whereby it can
bidirectionally connect the input and output processes for blood glucose to
liver carbohydrate metabolism. As noted by many authors,111, 171 the steady–
state level of blood glucose is close to the calculated value at which the rates
of liver hexokinase IV and glucose–6–phosphatase would be equal and there
would be no net flux through the cycle. This is because the half–saturation
value of the sigmoidal hexokinase IV rate curve is in the 1–10 mM range
characteristic of mammalian blood glucose levels. When blood glucose level
changes, the substrate cycle automatically operates as a switch with no need
for any other mechanism. Thus a rise in blood glucose level, which rapidly
equilibrates with liver cytosol glucose causes an increases in the rate of phos-
phorylation but has no significant effect on the glucose–6–phosphatase rate
(since the demand and supply system for glucose–6–phosphate in the liver
is fairly effective at ensuring homoeostasis of its concentration), so the cycle
switches to cause net phosphorylation and glucose uptake by the liver. The
reverse happens when blood glucose level falls and the rate of phosphorylation
drops below that of the phosphatase. Thus the homoeostasis results from the
kinetic balance of supply and demand at a complex metabolic crossroads, in
which the switching property of the cycle is a relevant factor. There is no
need for (and no point in) invoking a specific buffering action of the cycle in
itself.
that precede any effects these signals may have on synthesis and degradation
of enzymes, and without direct interactions between the signal molecules and
specific pathway enzymes. This is clearly one of the most important aspects of
metabolic control. In this section I will present examples that show that many
of these controls are exerted via mechanisms that involve post–translational,
covalent modifications of proteins that affect their enzymic activity. Since
the processes involve changes in enzymic activities, they can be interpreted
in Metabolic Control Analysis through flux control coefficients.
7.4.3 Phosphorylation
It is impossible to know where to start to produce a simple, coherent descrip-
tion of such a complex, interacting and interlocking network as the eukaryotic
phosphorylation systems. The difficulty is illustrated by the general history
of discovery of a system. This usually stems from the observation that a par-
ticular target enzyme undergoes phosphorylation. In turn, this leads to the
identification of a kinase that is responsible, and which is known initially by
the name of its target (even when this leads to comic names like phosphory-
lase kinase kinase for the kinase that phosphorylates phosphorylase kinase).
Generally, it later emerges that this kinase is similar, or even identical, to
one known by a different name because it has been shown to phosphorylate
7.4 Regulation by covalent modification of enzymes 233
a different target, and the kinase is named after the signal that stimulates it
to act rather than by its target.
The position is even more complicated with the protein phosphatases, since
these do not show unique associations with target enzymes, specific protein
kinases or signal systems, though they are affected by all of these.
Furthermore, although phosphorylation of an enzyme can lead to activity
changes, many enzymes are subject to multiple phosphorylations at differ-
ent sites, and some of the sites appear to affect ease of phosphorylation or
dephosphorylation at other sites rather than to directly affect activity.
Signal Enzyme
NH2
C
N
N C
CH
HC C
N
N
5’ CH2 O
C C
O 1’
H H
O P C C
3’
O- O O
OH
brane. There are specific receptors at the extracellular surface that can bind
stimulatory or inhibitory signal molecules (Rs and Ri respectively). The in-
teraction of these receptors with the catalytic subunit is mediated by guanyl–
binding regulatory proteins, again of stimulatory and inhibitory types (G s
and Gi ). The G–proteins bind GTP when they interact with the receptor–
hormone complex, and the G–protein complex with GTP in turn interacts
with adenylate cyclase (or other target protein in the general case). The
action of the G–protein terminates when it hydrolyzes its bound GTP. Hor-
mones that act on adenylate cyclase through a stimulatory G–protein in-
clude adrenaline (epinephrin), adrenocorticotropin, thyroid–stimulating hor-
mone (TSH), luteinizing hormone (LH), secretin and glucagon, not necessarily
all in the same tissue though those mentioned do all stimulate adenylate cy-
clase in adipose tissue, eventually causing lipolysis. The action of adrenaline
is complex because it acts on adenylate cyclase via a class of receptors called
β–adrenergic receptors, found particularly in muscle, but via α–adrenergic
receptors (for example, in liver) it acts to raise intracellular Ca2+ concentra-
tions. Adenosine inhibits adenylate cyclase in adipose tissue via an inhibitory
G–protein.
Stimulation of adenylate cyclase leads to a rise in the intracellular concen-
tration of cyclic AMP, the principal, perhaps only, effect of which is stimula-
tion of cyclic AMP–dependent protein kinase (protein kinase A). The amount
by which the cyclic AMP concentration increases must depend on the activity
and kinetics of the phosphodiesterase enzymes present in the cells, but since
there are phosphodiesterases with Km values above the intracellular levels of
cyclic AMP, it is likely that the concentration changes are at best propor-
tional to the stimulation of the cyclase. Protein kinase A occurs as a set
of isoenzymes, but there are two basic classes: Types I and II.13 Both are
tetrameric proteins formed of two regulatory subunits (R) and two catalytic
subunits (C) in the inactive state. Activation involves binding of four cyclic
AMP molecules by the regulatory subunits and release of free, active catalytic
subunits:
R2 C2 + 4 cyclic AMP *
) R2 (cyclic AMP)4 + 2C
The two sites for cyclic AMP per regulatory subunit interact cooperatively
so that the Hill coefficient for kinase activation with respect to cyclic AMP is
about 1.6.13 The difference between the Type I and Type II kinases is that
the latter undergo autophosphorylation of the regulatory subunit, though the
physiological significance of this, and the respective physiological roles of the
isoenzymes are not very clear.
There is a large number of targets for phosphorylation by protein kinase
A. Not all protein serine or threonine residues can be phosphorylated, nor all
proteins, so there must be specific sites recognised by the kinase. A common
sequence at phosphorylation sites involves two basic amino acids (lysine or
arginine) to the N–terminal side, for example arg-arg-X-ser and arg-lys-X-ser.
The effects on the target enzyme can be inhibitory or activatory; in general
where antagonistic enzymes in potential substrate cycles are phosphorylated,
236 7 Control structures in metabolism
cyclic AMP.44
10
1
b
0.1
ANP, mM
0.01
0.001
0.0001
0 1 2 3 4
ATP, mM
Relative sensitivity
-5
a
-10
-15
-20
-25
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
• The total quantity of the adenine nucleotides, ATP, ADP and AMP, is
fixed, at least in the short term, by the availability of their common
adenosine moiety.
2 ADP *
) ATP + AMP
The effect of this is illustrated in Fig. 7.15, which shows the equilibrium
relationship between the three nucleotides as a function of ATP for an assumed
total of 5 mM adenine nucleotides. Measurement of total adenine nucleotide
levels in cells commonly shows that ATP can be maintained at over 80%
of the total, even in muscle after a 100–fold increase in energy expenditure.
Nevertheless, an increase in ATP utilization in the cell does cause a drop in
its concentration, however slight measurements show this to be. What the
Figure shows is that, in these regions of high relative ATP content, a very
slight fall in ATP is accompanied by a rapid rise in AMP from its very low
level. It is true that ADP concentrations also show an almost as rapid rise,
but ADP has a disadvantage as a signal: any particular ADP concentration
could correspond to two different levels of ATP. The qualitative impression
of the greater sensitivity of AMP levels to an ATP change is confirmed by
my quantitative calculations of the sensitivities69 shown in Fig. 7.16. These
show that sensitivity to AMP change is always the greatest provided ATP
concentration exceeds that of AMP (i.e. for fractional ATP contents greater
than about one third), and that the factor becomes very large indeed when
most of the nucleotide exists as ATP. When corrections are made for bound
ADP and AMP contents in cells, the ATP increases to as much as 98% of
the total free adenine nucleotides.164, 258 This would suggest the possibility
of extremely high sensitivities to ADP and AMP, though the exact values are
hard to predict as the binding of ADP and AMP affects the graphs (without
changing the qualitative picture).
The AMP–dependent protein kinase was discovered as a kinase responsible
for the phosphorylation, and inactivation, of acetyl CoA carboxylase, the first
step in the pathways of fatty acid and cholesterol synthesis.28 It then became
7.4 Regulation by covalent modification of enzymes 241
apparent that this kinase was the same as one that had been identified as re-
sponsible for the phosphorylation (and inactivation) of hydroxymethylglutaryl
CoA reductase, the first step in the branch to cholesterol and other steroids.
It has also been found that the AMP–dependent kinase phosphorylates the
hormone–sensitive triacylglycerol lipase of adipose tissue; in this case, the
phosphorylation does not affect the enzyme’s activity directly, but it prevents
its phosphorylation (and activation) by protein kinase A, thus preventing re-
lease of fatty acids and cholesterol from the fat stores. The AMP–dependent
kinase is activated in two ways by AMP: firstly, AMP is an allosteric activa-
tor of the enzyme, and secondly, AMP makes the enzyme more susceptible
to phosphorylation by a kinase (called kinase kinase!) that remains attached
to the the AMP–dependent kinase throughout nearly all purification steps. 270
Together, the effects cause up to a 50–fold increase in activity of the kinase in
response to AMP, providing further amplification of the effect of a change in
ATP. The metabolic consequence is that fat metabolism is inhibited in condi-
tions where demand for ATP is causing its level to drop, as signalled by the
rise in AMP.
The activity of the AMP–dependent protein kinase is also stimulated by
very low concentrations of fatty acyl CoA esters, so it probably also limits the
synthesis of fatty acids and cholesterol and the hydrolysis of triacylglycerols
under conditions where fatty acyl CoA is already accumulating. On this basis,
it could also be regarded as an example of the next class of protein kinases.
7.4.3.2 Phosphatases
hormone + receptor *
) hormone–receptor complex
...
.. ⊕
..... PDE.
ATP ...................... cAMP ........................AMP
...
.
.....⊕
protein kinase A
...
...
... ... ...
.. .. .
... ... ATP..
........... .....⊕ ..........ADP ...
... ... ..... .
..
..
.
..
...
...
.. ..
................. .
.. .. ...... ..........
.. ... ........ ........
... ... ..... ....
.. .. phos kinase phos kinase–P
... ... ⊕
... ....... ... ...
ATP .
... ADP .......... .........
..
... ... ...................................... .....
.. ....................................... .
... ...
.
. ..
......
. . ..
....
........... .. .
.... .. . . ... ... .
...
.......
........ ........ ....... ... ...... ...
..
... ..... ...... Pi .. ⊕ H2 O ....
...... .. .
...
... Inhib 1
ª
Inhib 1–P ............................................ ... ...⊕
....
....... . . .
..
.. ..... prot Pase 1 ...
... .........
............. .
....
....
... .......... .
... ... ........... ...
.. ............ .... ............ .
.. ADP
.
... Pi . ⊕ H O ATP
.......... ... .......... .
... ... 2
....................................
. ............
..
... prot Pase 2B
...... ........... ........
... ⊕ .. . ..
. .......
.. phos b phos a
..
ATP
......... ...... ........ADP ... ........
........ ...
. .. . . .
.
.....
..................................
.... .
...
...
...
............
. . . .. .. .
.... ....
.. . . . . ................ .
...
... .........
........ ....... Pi ....
. ........... ... ............... .
..
..... .... ª
............................................. . . .. . ⊕ H O ..
ª
GBS–P ...................................................................................................................
...... ... 2 .
GBS . ..⊕
......
........ ... ...... prot Pase 1
. ...
............ .....
................................. ..
............... ..... ............... ....
Pi ... ⊕ H2 O .....
. glycogen + Pi ........................... G1P
prot Pase 2B
When cyclic AMP levels fall, phosphorylase will tend to return towards
the dephosphorylated b state. This will in part be caused by the reduced
rate of phosphorylation of phosphorylase kinase, Inhibitor 1 and glycogen
binding subunit. Inhibitor 1 and glycogen binding subunit will return more
towards their dephosphorylated state because they are acted on by protein
phosphatase 2B, which has not been inhibited. This relieves the inhibition on
protein phosphatase 1, so that the rates of dephosphorylation of phosphorylase
kinase and phosphorylase are enhanced.
The same mechanism that activates phosphorylase inactivates glycogen
synthase, which in its phosphorylated form is only active when allosterically
stimulated by raised concentrations of glucose–6–phosphate. The dephospho-
rylated form is active without glucose–6–phosphate. Protein kinase A phos-
phorylates glycogen synthase, but the details of the system are much more
involved than for phosphorylase because glycogen synthase undergoes multiple
phosphorylations and a range of protein kinases are responsible. So although
it is certainly the case that cyclic AMP–linked hormones inactivate glycogen
synthase, the relative importance of the different components in physiological
conditions has been difficult to establish. In liver there is an additional link
to the reciprocal relationship between phosphorylase and glycogen synthase:
protein phosphatase 1 is inhibited by phosphorylase a, so that glycogen syn-
thase cannot be dephosphorylated to its active form until virtually all the
phosphorylase has been returned to the relatively inactive b form.
There are a number of points in the phosphorylase cascade where Ca2+ can
act. It directly activates phosphorylase kinase, as mentioned above, regard-
less of phosphorylation state. However, it also activates protein phosphatase
2B, which will lower the amounts of the inhibitory forms of Inhibitor 1 and
glycogen binding subunit so that protein phosphatase 1 becomes more active.
In addition, by activating one of the cyclic AMP phosphodiesterases, Ca 2+
will tend to lower cyclic AMP levels and reduce the degree of activation of
protein kinase A.
7.4.4 Nucleotidylation
Whereas phosphorylase was the mammalian enzyme that led to the discov-
ery of cyclic covalent modification systems, in bacteria it was the glutamine
synthase (properly glutamate–ammonia ligase) of E. coli that furnished the
first example. Glutamine synthase has a central role in nitrogen metabolism,
since approximately 12% of cellular nitrogen comes from the amide nitrogen of
glutamine, and most of the remainder largely from the α–amino group of glu-
tamate rather than directly from ammonia. In bacteria and higher plants, the
net synthesis of glutamate from 2–oxoglutarate and ammonia in turn depends
on a cycle involving both glutamine synthase and glutamate synthase:
...
..
UTP ......... ......⊕ .........PPi
.................................
. . ..
..
.. .......... ..... ................
.
..... ................. ......
PII .... UT .... PII (UMP)4 ...
.......
...... ............
.....UR
....
...
.......... ...... ............... ...
..
. .
................................ ...
......... ...... ....... ...
UMP ⊕..
.. ....⊕
H2 O
.... .. ...
.. ..
glutamine .. .. 2–oxoglutarate
.. ...
.. .. ..
.. .. ..
Pi. ª.. .. ...⊕ .ADP
......... ..... ..... .... .........
.. .
................................
.
............. ..... ...............
.... ..... ...... ......
...ATd.....
GS(AMP)12 ... .
. GS
....... ....ATa .............
....... . .
......
........... ..... ..............
................................
......... ...... .......... .......
⊕ PPi . .... ..ª ATP
⊕ .. ..
NADP+ + ADP + Pi
The problem that organisms have to resolve is the need to maintain appropri-
ate levels of glutamate and glutamine to meet synthetic requirements without
withdrawing all the 2–oxoglutarate from the tricarboxylic acid cycle, and to
manage to do this across the range from conditions where usable nitrogen is
in short supply to those where it is in excess. The enzyme has already been
mentioned in Chapter 1.2.1 and earlier in this chapter (Section 7.2.3.3.3) be-
cause it is an allosteric enzyme affected by 8 different feedback inhibitors.
(The ninth effector is Mg2+ .)
It was Earl Stadtman’s group at Bethesda that was largely responsible
for establishing that glutamine synthase in E. coli was subject to inactivation
by adenylylation (addition of AMP groups) of tyrosine residues38 catalyzed
by an adenylyl transferase. Glutamine synthase is composed of 12 identical
subunits, and each of these has one modifiable tyrosine. The enzymic activity
of a glutamine synthase molecule is inversely proportional to the number of
subunits that have been modified. Reactivation of the enzyme is by phospho-
rolysis of the adenylyl groups to give free enzyme and ADP:
7.4 Regulation by covalent modification of enzymes 247
The relative levels of the two activities of the adenylyl transferase are recip-
rocally modulated by two types of factor (Fig. 7.18). Firstly, there are metabo-
lites that affect the activity. Of these, glutamine itself and 2–oxoglutarate
(product and substrate of the synthase) have oppositely directed effects with
glutamine stimulating the inactivation and inhibiting the activation, whereas
2–oxoglutarate does the opposite. The second factor is the tetrameric regu-
latory protein PII discovered by Shapiro. This also exists in two forms. The
unmodified form is an activator of the adenylylation (inactivation) of glu-
tamine synthase. It is modified by uridylylation (Fig. 7.18), and the uridy-
lated form is an activator of the de–adenylylation (activation) of glutamine
synthase. The interconversion reactions affecting PII mirror the modification
of glutamine synthase itself. The protein–PII uridylyl transferase and the
uridylyl–removing enzyme (which this time is a hydrolysis reaction to form
UMP) are separate activities on the same polypeptide chain (EC 2.7.7.59).
Glutamine stimulates the uridylyl–removing activity, thereby leading to ad-
ditional activation of the adenylylation of glutamine synthase (in addition
to its direct affect on this latter modifying reaction). Again, 2–oxoglutarate
has the opposite activity, since it stimulates the uridylyl transferase, forming
the modified PII that stimulates the de–adenylylation of glutamine synthase.
Thus the control of glutamine synthase involves a cascade system of two cyclic
modification reactions, but these cycles are doubly coupled by the reciprocal
activities of the two forms of PII . The four modifying activities between them
also respond to over 20 other assorted metabolites, including nucleotides and
intermediates of the glycolytic pathway and the tricarboxylic acid cycle. (A
further layer of complexity is added by the role of PII in modulating the tran-
scription of glutamine synthase. This will not be discussed further here, but
is reviewed in reference [197].)
The detailed properties of a system like this with such a network of inter-
acting effects are not immediately obvious from the structure of the cascade.
Stadtman and Chock, and other members of their group, therefore undertook
an extensive theoretical analysis of the possible behaviours of cyclic modifica-
tion cascades. Their conclusions are discussed in the next section, as they are
applicable to cyclic modification schemes in general (including phosphorylase)
and not just glutamine synthase.
248 7 Control structures in metabolism
Ki + N *
) Ka
. ..
..
ATP ... ADP
. ....... ...
.......................................
........ .. .
.. .
........
.... ..........
........ ......
........ ..... .
... .....
E E–P
....... ...
.....
....... ..
........
........... .
...
.... ............
................................................
..... ... ..
Pi .. H2 O
..
Pi + F *
) Pa
the difference in the amounts of the two enzymes. This is undoubtedly an im-
portant aspect of many modification cycles, such as cases where small amounts
of hormones (at perhaps 10−9 M) combine with relatively few receptors to act
eventually on relatively abundant enzymes (perhaps at concentrations around
10−5 M). Another example is in the irreversible modification cascade of blood
clotting, where the initial signal is tiny but a macroscopic blood clot is formed.
Important as it may be, there is little mystery about catalytic amplifica-
tion. It should be noted that a high degree of catalytic amplification is only
obtained at a price that is not evident in steady state studies: the larger the
amount of target enzyme relative to modifying enzyme, the longer the system
will take to respond to a change in the amount of an effector.
1.0
0.8 c
b
Fractional modification
0.6 a
0.4
0.2
0.0
-3 -2 -1 0 1 2
log N
7.4.5.3 Amplitude
Fig 7.20 also shows that the modification cycle cannot necessarily produce
complete conversion of the target to its phosphorylated form. The symmetry
of the cycle implies that there will also be circumstances where there cannot
be complete conversion to the dephosphorylated form. The amplitude is the
range of fractional modification that is covered in going from zero level of an
effector to saturating levels. For curve b of Fig. 7.20 it is nearly 1.0, whereas
for curve a it is only 0.67. Variation of the parameters of the model can
produce amplitudes anywhere from 0 to 1.
7.4.5.4 Sensitivity
Stadtman & Chock were concerned to know whether their model of a covalent
modification cycle produced a curve that was steeper or less steep than the
familiar rectangular hyperbola for enzyme kinetics and ligand binding. To
measure this they introduced a measure of sensitivity. They took as their ref-
erence the change in ligand concentration required to go from 10% to 90% sat-
uration for a normal rectangular hyperbolic binding curve. From Eqn. [3.10],
Chapter 3.4, it is possible to calculate that this requires an increase in the
ligand concentration equal to 8.89 times the dissociation constant (or ligand
concentration that gives 50% binding). They therefore defined sensitivity, S,
as:
8.89 N0.5
S= (7.11)
N0.9 − N0.1
where N0.5 is the concentration of the effector N that gives 50% of the maximal
amplitude of modification. This sensitivity function differs from those defined
in Metabolic Control Analysis (i.e. control coefficients and elasticities) in that
it is measured over a finite interval rather than from the gradient of a curve,
but it does share the properties of being dimensionless and of measuring the
steepness of response. It is actually more like the Hill coefficient (Chapter 3.4,
7.4 Regulation by covalent modification of enzymes 251
p. 69) in that values of S greater than 1 indicate that the curve relating the
fractional modification to effector concentration is steeper than a rectangular
hyperbola, whereas values less than 1 show it is less steep. As it happens,
Stadtman & Chock’s relatively simple model for the modification cycle gives
sensitivity values equal to 1 unless an effector acts at more than one site.
For example, if N activates the kinase and inhibits the phosphatase, then the
sensitivity is greater than 1. (The result is similar to that seen in curves a
and c of Fig. 7.21, though these were calculated for a different model.) This
is equivalent to positive cooperativity in its outcome, though it is obtained
without any cooperative effects in the binding of the effector. An example of
an effector that acts in opposite directions on the two converting enzymes of
the cycle is glutamine in the modification of glutamine synthase (Fig. 7.18).
There are other ways that the sensitivity can be altered that will be discussed
further below.
7.4.5.7 Ultrasensitivity
The next development in the theoretical understanding of covalent modifica-
tion systems came from Albert Goldbeter of the Free University of Brussels
and Daniel Koshland of the University of California. Whereas the analy-
sis described above had used simple chemical kinetics for the modification
reactions, Goldbeter and Koshland represented them as true enzymic reac-
tions with the dependence of rate on the concentration of their substrates
(the unmodified and modified forms of the target enzyme) described by the
Michaelis–Menten equation.88 For their sensitivity measure, they used the
measure earlier developed by Koshland and his colleagues for describing the
steepness of cooperative binding curves in allosteric enzymes.137 For a factor
N that affects the degree of modification, their sensitivity RN is defined as:
N0.9
RN = (7.12)
N0.1
N could be either an effector, or the ratio of limiting rates of the two convert-
ing enzymes. Unlike Stadtman and Chock’s sensitivity measure (Eqn. [7.11]),
or the Hill coefficient, this value decreases with increasing steepness of the
curve, from a value of 81 for a rectangular hyperbolic curve, to a minimum of
1 for an infinitely steep transition. The use of the different definitions means
that neither the equations derived by the two groups nor the numerical mea-
sures of sensitivity are exactly comparable, though in qualitative terms they
are referring to the same concept. (Such differences in the operational defini-
tion of sensitivity, of which these are merely two instances, was the main rea-
7.4 Regulation by covalent modification of enzymes 253
1.0
c a
Fractional modification
0.8 b
0.6
0.4
0.2
0.0
0.01 0.1 1 10 100
[Effector], F
son that Kacser and Burns abandonned their original term sensitivity for the
flux control coefficient.) The analysis carried out by Goldbeter and Koshland
showed that the sensitivity was much more variable than in Stadtman and
Chock’s studies, where it only changed in response to the number of steps at
which an effector acted.
Their novel finding was that, if one or both of the converting enzymes
is operating close to saturation, the fractional modification of the target en-
zyme can become an extremely sensitive function of the ratio of the limiting
rates of the converting enzymes. They termed this zero order ultrasensitivity,
since a saturated enzyme shows near zero order kinetics in the terminology
of chemical kinetics, and it was possible to show theoretical curves with a
steepness equivalent to a Hill coefficient, nH , of 13 or more. The effect is
illustrated in Fig. 7.21, where curve a has been calculated with their model
for concentrations of the target enzyme well below the Km values of the mod-
ifying kinase and phosphatase. Curve b is far steeper, but the only change is
that the total amount of target enzyme has been increased to 10 times the
level of both Km values. For comparison, the smaller increase in steepness
obtained with an effector that acts on both kinase and phosphatase is shown
in curve c. Goldbeter and Koshland also showed that the ultrasensitivity can
be amplified in a multicycle cascade if each component cascade exhibits ultra-
254 7 Control structures in metabolism
unsaturated with the target enzyme, and hence under these circumstances,
the covalent modification cycle actually acts to reduce the response of the
flux to the effector. We were also able to confirm and extend the conclusions
reached by Cárdenas and Cornish–Bowden about the conditions that would
lead to ultrasensitivity, with the advantage that many of our arguments de-
pend on relative values of elasticities, and not on the molecular mechanisms
underlying them.
Our overall conclusions were that only a quite restricted set of conditions
generate ultrasensitivity and ensure it is translated into a more sensitive re-
sponse of pathway flux to an effector. This is even more so in multicyclic
cascades, where additional terms appear in the response coefficient, but each
of these with a value less than 1 contributes to an unavoidable attenuation of
the sensitivity of the flux to the signal. Nevertheless, the actual examples of
covalent modification cycles presented above show that the theoretical mod-
els studied so far are relatively simple and that there is plenty of potential
for complex properties to emerge from covalent modification cascades. Also,
most of the analysis has been concerned with steady state behaviour, but the
transient behaviour in the change between metabolic states is almost certainly
very important as well. However, the weakest link in all the studies, whether
practical or theoretical, has been that little attention has been paid to the
link between the change in modification state and the effect on the metabolic
pathway of which the target enzyme is part. (Modular Control Analysis is a
recent development of Control Analysis that deals more systematically with
this problem.123 ) Metabolic Control Analysis has shown that enzymes that
are truly rate–limiting, with flux control coefficients of 1, are not generally
found. It has also shown that for an enzyme with a flux control coefficient
less than 1, the scope for increasing flux by activating the enzyme is relatively
limited (see Chapter 5.2.2). How then, can we explain covalent modifica-
tion mechanisms that only appear to exist to cause a significant change in
flux by making large changes in the amount of active enzyme in response
to small changes in signal molecules, when the evidence suggests this cannot
be effective? The traditional paradigm of covalent modification systems was
that they control metabolism by acting on ¡rate–limiting enzymes. This is no
longer tenable; a new view of how control of metabolism is actually exerted
in the light of the analyses presented in this Chapter will be developed in the
next and final chapter.
7.5 Summary
1. Traditional theories of metabolic control assumed there were obvious
advantages of controlling the rate of a pathway near its start. Analysis
of the control of supply of and demand for a metabolite shows that
better homoeostasis of metabolite concentrations is achieved when the
greater flux control is exerted by the demand reactions, i.e. towards the
end.
7.6 Appendix 1: Control Analysis of substrate cycles 257
Further reading
1. Stadtman, E. R. Mechanisms of enzyme regulation in metabolism in The
Enzymes, Boyer, P. D. (ed.) 3rd ed. Vol. 1, pp. 397–459, Academic
Press, New York (1970)
2. Katz, J. & Rognstad, R. Futile cycles in the metabolism of glucose,
Current Topics in Cell Regulation, 10, 237–289, (1976)
3. Hue, L. The role of futile cycles in the regulation of carbohydrate meta-
bolism in the liver, Adv. Enzymol. 52, 247–331, (1981)
4. Hue, L. Futile cycles and regulation of metabolism in Metabolic Com-
partmentation, Sies, H. (ed.) pp. 71–97, Academic Press, London
(1982)
5. Boyer, P. D. & Krebs, E. G. (eds.) Control by Phosphorylation, Parts A
& B in The Enzymes, Vols. XVII & XVIII, Academic Press, New York
(1986)
6. Shaltiel, S. & Chock, P. B. (eds) Modulation by Covalent Modification.
Current Topics in Cell Regulation, 27, (1985)
258 7 Control structures in metabolism
Algebraic solution of these equations gives each of the flux control coef-
ficients in terms of the elasticities and the rates of the cycle enzymes. 69 The
flux control coefficient of enzyme 2a is:
Conclusion
261
262 8 Conclusion
Thus there is no obvious ‘functional gap’ in flux control that allosteric effects
have to fill so long as covalent modification mechanisms are operating.
Even the teleological principle, that it is advantageous to control a pathway
near its beginning, is open to doubt. The study of supply and demand path-
ways (Chapter 7.1), combined with the effects of feedback inhibition, shows
that metabolic control structures exist to give good regulation (i.e. strong flux
control combined with relatively weaker effects on metabolite concentrations)
if flux control is concentrated at the demand end of the pathway. It is not not
known at present whether there are equivalent control structures to ensure
such good regulation if control is exerted near the beginning of the path-
way, but the evidence is against it. In this context, it is most remarkable that
control of flux in cells is often achieved with so little change in metabolite con-
centrations. For example, when a 1000–fold increase in the glycolytic rate of
rat muscle was obtained by stimulating the muscle, the measured metabolites
increased less than 10–fold in concentration.107 These and similar experi-
ments led Bücher and Rüssmann24 to speculate that metabolic pathways are
controlled at more than a single enzyme in order to ensure homeostasis of
metabolite concentrations, as mentioned in Chapter 1.4.3. Similarly, Helmre-
ich and Cori101 noted the constancy of metabolite levels with large glycolytic
flux changes in vertebrate muscle and also concluded that glycolysis must be
controlled as a unit, particularly as they could not account for the changes in
rates of several of the enzymes by the small changes in concentrations of their
substrates and effectors.
Metabolic Control Analysis has confirmed these authors’ suspicions that
there was a problem still to be solved about the control of glycolysis. Even
if the fallacy of the rate–limiting step is made apparently more respectable
by supposing that control is exerted by action at the enzyme with the largest
flux control coefficient, control at a single site has a clear problem: Small and
Kacser’s theory of large changes (Chapter 5.2.2) shows that it is impossible
to achieve anything other than small total changes in the flux in response to
massive changes in an enzyme’s activity if its flux control coefficient is only
slightly less than 1. Therefore, if large changes in flux occur in cells (such as
in glycolytic flux in muscle upon changes in mechanical activity, or in pho-
tosynthesis in plants upon changes in illumination) they cannot be achieved
by controls acting at a single site; there must be widespread sites of action
of any control signal. Furthermore, even though the covalent modification of
phosphorylase apparently provides an obvious mechanism for flux control of
muscle glycolysis, the analysis of homoeostasis and regulation makes it un-
likely that control of flux solely from the beginning of the pathway can explain
such good homeostasis of metabolite concentrations as is observed.
Thus although this book has covered a systems theory of the control of
metabolism and has examined the range of control and regulatory mecha-
nisms available in metabolism, we have arrived at its final chapter without a
satisfactory answer to the question: how is the flux of a metabolic pathway
controlled in vivo? In the next section, I will suggest an answer, developed
with my colleague Simon Thomas,72 that is consistent with the theory and
8.1 A new view of the physiological control of flux 263
' $
3 Fig. 8.1: Schematic cell
? metabolism for the Universal
2
S - G Method.
The metabolic inputs are grouped
as N, the desired product is P,
& %
and all the other products and
1 cell biomass are G. S is the
metabolite at which synthesis of
P branches from the rest of met-
? abolism. Each arrow represents a
P complex metabolic network.
directly to this output from the input(s), with the largest changes in activity
being made in all the enzymes in the final linear sequence to the output, and
with progressively smaller changes being made in each preceding branch so
that the fluxes in colateral branches remain unchanged. In Fig. 8.1, suppose
that the rate of production of P (J1 ) is to be increased 5–fold. All the en-
zymes along that branch would have to be increased by a factor of 5. This
creates an additional consumption of S equal to 4J1 . To ensure that the con-
centration of S, and hence the rate of synthesis of G, remain constant, the
rate of production of S has to be increased by the same amount. That is
J3 has to be increased to J3 + 4J1 . The fractional increase required in the
enzymes synthesising S is therefore 4J1 /J3 . Since J1 is much less than J3 ,
the change required in J3 is quantitatively less significant, and probably the
need to change enzyme levels dies away as the pathway is traced back to the
inputs. In effect, the method exploits the different behaviour of fluxes and
concentrations when the activities of a sequence of enzymes are all increased
in the same proportion. According to the summation theorem for flux control
coefficients (Chapter 5.2.3), the flux will increase in proportion to the enzyme
activities, whereas the metabolite concentrations will remain constant in ac-
cordance with the summation theorem for concentration control coefficients
(Chapter 5.8.1). It might not be necessary to increase every enzyme along
each branch affected provided that any enzymes omitted have flux control
coefficients very near zero.
The potential for success of this strategy of multiple enzyme manipula-
tions, compared with the failure of manipulating single enzymes in a pathway,
has been illustrated by Peter Niederberger and his colleagues for the trypto-
phan biosynthesis pathway of yeast.117, 175 There are 5 enzymes from cho-
rismate, the common precursor of the amino acids tryptophan, tyrosine and
phenylalanine, to tryptophan itself, coded by the genes TRP1 to TRP5. The
approximate control coefficients of these 5 enzymes were estimated from mea-
surements of tryptophan synthesis flux in mutant/wild–type heterozygotes in
tetraploid yeast. None of the enzymes had a large control coefficient, and
the sum of the five was 0.26. The five enzymes were overexpressed in yeast
by incorporating their genes, either singly or in combination, into a yeast
plasmid. As might have been expected from the low flux control coefficients,
overexpression of any of the enzymes singly, even up to 50–fold, had barely
any effect on the flux. The more interesting result was the demonstration
that the flux increase achieved when all the enzymes were changed together
was far greater than the product of the flux changes obtained by increasing
each enzyme separately. The results are summarized in Table 8.1, where it
is clear that the best increase in flux is obtained when all five enzymes are
overexpressed. Even so, an average 23–fold excess of the enzymes still only
produces a 9–fold increase in flux because no changes have been made to in-
crease the supply of chorismate. Incidentally, the experiments also show that
abolishing the allosteric feedback on anthranilate synthase (TRP2) does not
significantly stimulate tryptophan synthesis.
The Universal Method is a theoretical proposal for engineering a desired
8.1 A new view of the physiological control of flux 265
- - - - - 1 1.0
- - + + - 58 2.0
+ + - + - 35 2.4
+∗ - + + - 34 1.2
+∗ + + + - 30 2.1
+ + - + + 19 8.2
+∗ + + + + 23 8.8
8.1.1.1 Lipogenesis
Lipogenesis, or triacylglycerol synthesis, (Fig. 8.2) exemplifies all three. One
type of genetic obesity in mice maps to a single locus, the obese gene. This
codes for a secreted protein, specifically made by adipose tissue, that ap-
pears to have signalling functions and the absence of which causes profound
obesity.282 The obese phenotype is linked to the alteration of enzyme levels
in adipose tissue and liver relative to wild–type levels25 as summarized in
Table 8.2. The mutant mouse appears to be capable of a rate of lipid syn-
thesis per unit lean tissue some three fold higher than normal mice.25 This
is consistent with the 2–4 fold activation of eight of the major enzymes of
lipid and carbohydrate metabolism as shown in Table 8.2 (point 2). (Effec-
tively it is more enzymes than this, because fatty acid synthase is itself a
multi–enzyme complex of the seven enzymes needed to add malonyl–CoA to
a growing acyl–CoA chain.) In connected parts of carbohydrate metabolism,
Table 8.2 shows that four enzymes are elevated by about 50%, and two adi-
pose tissue lipases are significantly depressed25 (point 3). Of the enzymes
in this table, acetyl CoA carboxylase, fructose–1,6–bisphosphatase, hormone–
sensitive lipase, phosphofructokinase and pyruvate kinase are known to be
subject to control, directly or indirectly, by protein phosphorylation and de-
phosphorylation reactions that respond to dietary status105 (point 4).
Two lines of evidence suggest that this is not an odd result from a par-
ticular type of mutant mouse. Firstly, Bulfield and his colleagues obtained
essentially similar results in a later study of the differences in enzyme contents
between lean and fat lines of mice that had been selected over 26 generations
from the same base population.4 For most of the enzymes in Table 8.2, there
is good evidence that the controls on their levels of expression respond to
starvation, refeeding or high carbohydrate diets105 in rats. On the other
hand, there is a lack of experimental evidence supporting any system–wide
role for allosteric effectors in physiologically significant control of lipogenesis
by dietary factors.105
8.1 A new view of the physiological control of flux 267
cytosol
glucose
NADP+ NADPH
8
? ¸
glucose-6-P - pentose-P pathway
7
? 6
- glycerol-P
?
P-enolpyruvate
5 NADPH NADP+
? ]
pyruvate ¾ malate ¾
$
² 4
CO2
oxac
? 6
pyruvate
- triacylglycerol
J 3
citrate
J
^
J
acetyl–CoA 3́
´
´ ?
´
´ acetyl–CoA
? q ´
oxac citrate 2
¸
TCA ?
cycle malonyl–CoA
® 1
}
?
fatty acid
mitochondrion
Group 1: elevated
adipose glycerol kinase 371
liver hexokinase IV 357
liver acetyl–CoA carboxylase 335
liver fatty acid synthase 267
liver malic enzyme 254
liver ATP citrate lyase 231
liver fructose–1,6–bisphosphatase 213
liver glycerol–3–phosphate dehydrogenase 210
Group 2: slightly elevated
liver glucose–6–phosphatase 163
liver phosphofructokinase 160
adipose glucose–6–phosphate dehydrogenase 151
liver pyruvate kinase 150
liver glucose–6–phosphate dehydrogenase 139
Group 3: depressed
adipose hormone–sensitive lipase 73
adipose monoacylglycerol lipase 20
measured increased significantly. All the enzymes of the urea cycle increased
2–3 fold in activity, which is sufficient to account for the flux change if there is
also a slight stimulation by increased substrate supply (point 2 above). There
were also large increases in alanine transaminase and glucose–6–phosphate
dehydrogenase.
George Weber and his colleagues carried out extensive investigations in the
1960s into the effects of dietary changes, starvation and hormone treatments
on carbohydrate metabolism in rat liver. The hormones whose effects were
covered included glucocorticoid steroids, which stimulate gluconeogenesis, and
insulin, which suppresses it.268, 269 They concluded that the liver enzymes fall
into three groups: the exclusively gluconeogenic enzymes (in the left hand
8.1 A new view of the physiological control of flux 269
See Fig. 6.17 for the positions of most of these enzymes in gluconeo-
genesis. The effects shown are ones for which the molecular mech-
anisms have recently been established.147 But the existence of these
two groups of oppositely–regulated enzymes has been known for 30
years.268
column of Table 8.3), the exclusively glycolytic enzymes (those in the right
hand column of Table 8.3 except aldolase), and the dual–function enzymes
(the rest). The amounts of the members of a group change coordinately,
but the directions and magnitudes of the changes differ between the groups,
exemplifying points 1 and 3 above. For several of these enzymes, the molecular
mechanisms of control of enzyme amount have been shown to include the
actions of insulin, cyclic AMP and glucocorticoids on transcription.147, 184
Some support for a similar pattern of control by other mechanisms (point 4)
comes from the effects of protein phosphorylation. The actions of glucagon
and α–adrenergic agents in stimulating gluconeogenesis in rat liver involve
the inhibition of pyruvate kinase by phosphorylation, and the activation of
fructose–1,6–bisphosphatase and the inhibition of phosphofructo–1–kinase by
changes in the level of fructose–2,6–bisphosphate brought about by protein
phosphorylation. As it is possible that as many as 1 in 4 mammalian cell
proteins can be reversibly phosphorylated,37 it is even conceivable that there
are more undiscovered sites of phosphorylation involved in the stimulation of
gluconeogenesis.
acetyl–CoA cytoplasm
†
† uptake synthesis µ
?ª secretion
cholesterol ¾
1
cholesterol esters
∗
> µ
'
‡
† transport
$
?
cholesterol cortisol CS
6 6
NADPH
1P45011β ∗ DC
?
adrenodoxin
¸
ª R DC) 3
∗ P450SCC P45011β ∗ 11–DOC
6
11-DOC9 2 NADPH
© H
¼
© j
H
?
pregnenolone - pregnenolone
mitochondrion endoplasmic reticulum
8.1.1.5 Photosynthesis
Finally, light–dependent activation of plant photosynthesis in the chloroplasts
and associated metabolism in the plant cytoplasm illustrates rapid control
mechanisms acting at multiple sites throughout a metabolic network (point
4). As well as the diurnal cycle, plants are subject to continual fluctuations in
illumination from light flecks (shifting patches of sunlight filtering through the
leaf canopy) and changing orientations of their leaves as they bend and shake
in the wind. These transient increases in illumination can be responsible for a
significant fraction of daily carbon fixation and can cause sudden increases in
assimilation of CO2 of the order of 10–fold on a time scale of a few seconds.180
There are several components to plant photosynthesis. For the C3 plants
typical of temperate regions these are:
• the light reactions in the chloroplast, where photosystems I and II har-
ness the light energy on the thylakoid membranes and generate ATP
and NADPH;
• the so–called dark reactions in the chloroplast that use the ATP and
272 8 Conclusion
Short–term variations in the rate of the light reactions largely follow the
availability of light; however, the balance between the production of ATP
and NADPH is kept by regulating the ratio between non–cyclic photophos-
phorylation (which uses both photosystems and produces both products) and
cyclic photophosphorylation, which uses only photosystem I and produces
only ATP. This regulation involves reversible phosphorylation of the light–
harvesting component of photosystem II by a protein kinase that responds to
the redox state of the intermediate electron transport chain between the two
photosystems,15 but is not the aspect I wish to focus on here. Rather, I want
to look at how the assimilation of carbon in the chloroplast and cytoplasm
(Fig. 8.4) is controlled in accordance with the fluctuations in the supply of
$
cytosol CO2
' $
pep 4
⊕
- oxaloacetate
chloroplast
NH+ -NH+ - amino -
4 4 acids
6
CO2
1
RuBP
NO− ¾ NO− ¾ 3
NO− ¾
2 2 3
7
⊕ ⊕ 6
6
⊕
N
Ru5P PGA protein dephosphorylation
O
⊕ ⊕ ⊕
1 5
) - triose P -hexose P ?
- sucrose P
triose P
2
? - Pi
?
hexose P
sucrose -
6
ª
?
& %
starch
ATP and NADPH from the light reactions. The levels of ATP and NADPH
change relatively little during the day, so there is obviously control on the de-
mand reactions (i.e. the dark reactions) in accordance with the requirement
for effective homeostasis in supply–demand pathways.
The activation state of four of the enzymes of the Calvin cycle was discov-
ered by Buchanan to depend on light–mediated reduction carried out by small
proteins called thioredoxins.3 This is a covalent modification reaction involv-
ing the oxidation state of cysteine sulphydryl groups in the enzymes. When
they are oxidized to form disulphide linkages, the enzymes are inactive; when
they are reduced to the sulphydryl form, the enzymes are active. Thioredox-
ins are abundant in chloroplasts and themselves contain adjacent sulphydryl
groups that can be oxidized to disulphides. They are oxidized by atmospheric
oxygen, but are reduced by ferredoxin, the electron transport component
that is itself reduced by photosystem I and that in turn reduces NADP+ to
NADPH. Thus the redox state of thioredoxin senses the activation of the light
reactions and transmits this to activate phosphoribulokinase, glyceraldehyde–
3–phosphate dehydrogenase, fructose–1,6–bisphosphatase and sedoheptulose
bisphosphatase. The system can also respond to the state of the Calvin cycle
because the susceptibility of the enzymes to reduction is also modulated by
levels of the Calvin cycle intermediates.80 Ribulose–bisphosphate carboxylase
itself, the enzyme that actually fixes the CO2 , is controlled by multiple effects
including a covalent modification: an activation by carbamylation of a lysine
group promoted by rubisco activase, which is itself activated by light. 80 In
some plants, rubisco is inhibited by carboxyarabinitol–1–phosphate, which is
synthesized in the dark and degraded in the light;180 again, rubisco activase is
involved in releasing rubisco from this inhibition. Calvin cycle enzymes also
respond to the light–induced changes in stromal pH and Mg2+ concentra-
tion277 that occur as the light reactions pump protons out of the stroma into
the thylakoid space. In C4 plants, such as tropical grasses, the activity of one
of the key enzymes in their alternative mode of CO2 fixation, pyruvate phosph-
ate dikinase is controlled by phosphorylation in a light–dependent manner. 110
In addition, light–dependent changes in protein phosphorylation activate en-
zymes of the assimilatory pathways in the cytoplasm (Fig. 8.4), including
sucrose phosphate synthase, nitrate reductase and phosphoenol pyruvate car-
boxykinase110 presumably so that the relative amounts of starch, sucrose and
amino acids being synthesized are kept in balance both with one another and
with the amount of CO2 being fixed.
Much remains to be learnt about the molecular mechanisms underlying the
several different light–activated control systems operating in photosynthesis.
There are also differences between different types of plant, such as C3 and C4
plants. However, even though the account given above is not exhaustive in
terms of the enzymes whose activities are controlled by light, it is sufficient to
demonstrate that the rate of carbon assimilation is very far from being con-
trolled by a single enzyme such as rubisco. Indeed, the measurements of the
flux control coefficient of rubisco, made by Mark Stitt’s group and described
in Chapter 6.1.1.5, p. 149, reveal how unlikely this would be: under most
274 8 Conclusion
The first of these criticisms has been made particularly forcefully by both
Eric Newsholme and Dan Atkinson. Newsholme, with Bernard Crabtree, has
developed an alternative theoretical approach to metabolic control which in-
volves characterizing the net sensitivity of a pathway to an effector of an
enzyme.55 If necessary, a hypothetical effector is assumed to exist in order
to allow definition of the net sensitivity. In Metabolic Control Analysis, the
equivalent measure is the response coefficient (Chapter 5.4), so the key dif-
ference between the approaches is that in Control Analysis the flux control
coefficient is separated out as an individual entity that can be combined with
different elasticities to indicate the responses of the pathway flux to a range
of effectors whereas in Newsholme’s theory the net sensitivity is the primary
entity. Atkinson8 objected that the key feature in control of metabolism is the
action of effectors of allosteric enzymes which frequently change the affinity of
an enzyme for its substrates, not the catalytic activity. Again, the response
coefficient of the flux to the effector, formed from the flux control coefficient
of the enzyme and the elasticity of the enzyme to the effector, provides a mea-
sure of the strength of the response. Thus it is the belief of the advocates of
Metabolic Control Analysis, admittedly not shared by the objectors, that the
measures they seek are provided in Control Analysis. A more fundamental
rebuttal of their objection would be that control by allosteric effectors is not
276 8 Conclusion
the only control mechanism; the equally important effects of genes, DNA ma-
nipulation, environment, diet and hormones on metabolism can be mediated
through the change in the concentration of the active form of an enzyme, and
the flux control coefficient indicates what the effect of such a change will be.
The second objection returns to a theme that has run through this book.
There is no doubting that regulatory enzymes are often predicted by Control
Analysis to have small flux control coefficients, and the argument is there-
fore that there is little use in a measure that does not identify the important
characteristics of an enzyme. The reasons why regulatory enzymes subject to
feedback regulation have small flux control coefficients has been considered at
length in the Chapter on feedback inhibition (Chapter 7.2). I have sympathy
for the concept that there ought to be a measure of the strength of the regula-
tory role of an enzyme; in fact, some measures have been suggested but have
not been widely adopted. However, as long as molecular biologists are over-
expressing regulatory enzymes in the expectation of altering metabolic flux
and are being disappointed by the results, I believe the priority is to empha-
size the distinction between regulation and control and to persuade molecular
biologists to pay more attention to the flux control coefficient, which could
have been used to predict their disappointment.
There is truth in the third objection, made by Newsholme & Crabtree and
Savageau & Voit. The flux control coefficient is only valid under the condi-
tions of its measurement, and therefore it can make only limited predictions
about what will happen if circumstances change greatly. Everybody would
be happier if it was possible to make accurate predictions about the effects
of large perturbations induced by factors such as hormonal stimulation, but
this is extremely difficult. If we had complete models of metabolism, it would
be possible, but the information needed to build completely accurate models
is not, and may never be, available. If it were, we would not need Metabolic
Control Analysis. Henrik Kacser was sensitive to the force of this criticism
and his response to it was expressed in his characteristically entertaining way
in the last article he completed117 before his death in March 1995. In re-
sponse, he cited his work on the relationship between flux control coefficients
and large changes in enzyme activity (summarized in Chapter 5.2.2), and also
the Universal Method, referred to earlier in this chapter (section 8.1) for ob-
taining large changes in flux. As he pointed out, these go beyond the realm
where Metabolic Control Analysis is strictly valid, but have depended on the
insights from Control Analysis. Savageau’s criticism reflects the different aims
he had in developing his Biochemical Systems Theory, which is intended as
a complete tool for the modelling and analysis of stability and sensitivity. 217
Furthermore, he claims that Metabolic Control Analysis is only a limited sub-
set of his theory;218, 219 I accept that it is possible to view the two theories in
this light, but I and other advocates of Control Analysis feel that our theory
avoids some of the aspects of Biochemical Systems Theory that we regard as
undesirable.48
There have also been criticisms of the roles of the theorems, such as the
summation and connectivity theorems, in Metabolic Control Analysis. Kacser
8.2 Limitations of Control Analysis 277
& Burns119 gave them a central role in the development of the theory, and I
have used them to provide the explanations in this book. Savageau and col-
leagues219 have claimed that they are unnecessary because they are implicit
in the mathematical description of a metabolic system at steady state. It is
also true that Christine Reder in Bordeaux, Marta Cascante in Barcelona and
Christoph Giersch in Darmstadt have shown that pathways can be analyzed
in terms of Metabolic Control Analysis without assuming the existence of the
theorems,31, 32, 83, 195 and that the theorems then emerge from the analysis.
This is extremely useful work because it provides deeper mathematical foun-
dations for Control Analysis. It is not the same as saying the theorems have
no use; they summarize important properties that assist in explanations of
how control is distributed and make a link to a biochemical interpretation of
the mathematics. But at this stage, I am content to abide by the reader’s
judgement: take a look at some of the papers cited in this section, and see if
you prefer their approach.
one influenced the rates of the other, even though they did not appear as
substrates or products of its reactions? It is not difficult to find examples:
covalent modification cascades consist of a hierarchy of pathways where the
substrates and products of one pathway are the catalysts of the next path-
way down, but do not participate in the flux of that pathway. Transcription
and translation of genes to form enzymes, which then catalyze pathways is
another example where the mass flows at the protein synthesis level might
be weakly coupled to the mass flows at pathway level, but again there is a
catalytic relationship between them, and even the possibility of influence in
the reverse direction through effects of metabolites from the pathway on the
transcription and translation of the enzymes.
One possible solution in simple cases is to define the response of the path-
way to events occurring at a higher level in an ad hoc way. An example is
the treatment that Rankin Small and I gave to covalent modification reac-
tions, described in Chapter 7.4.5. However, Daniel Kahn in Toulouse and
Hans Westerhoff in Amsterdam have developed a Modular (or hierarchical)
Control Analysis that uses the same mathematical formulation as Metabolic
Control Analysis, but which can handle these problems systematically. 123, 224
In essence, the solutions contain terms that represent the standard Control
Analysis of the separate pathway components, but these are modified by in-
teraction terms between the pathways. The approach has much potential,
but is not for the mathematically faint–hearted; experimental applications
are under way, but are not yet concluded.
continue.
Because this is a biochemistry book, it would be easy to overlook the
important implications of Metabolic Control Analysis for genetics, which
were mentioned in Chapter 5.2.2 in connection with its relevance to the phe-
nomenon of dominance. Putting biochemistry and genetics together links
Metabolic Control Analysis with the evolution of metabolic pathways and even
with population ecology. These aspects are again more theoretical than prac-
tical, but Jean–Pierre Mazat’s group in Bordeaux has been using Metabolic
Control Analysis to diagnose and understand the genetic disorders known as
mitochondrial cytopathies.149, 150 These are a diverse group of inborn errors
of metabolism with variable symptoms. They are caused by mutations in the
mitochondrial genome, but the variability arises because there are multiple
copies of the mitochondrial DNA in every human cell, so that the proportion
of mutant to wild–type genes can be almost continuously variable, in con-
trast to nuclear genes where there are generally just two copies per cell. The
variability in the symptoms reflects the non–linear relationship between the
amount of an enzyme and the pathway flux. For many of the enzymes, the
content can be reduced significantly with limited effects on flux, and therefore
with little in the way of symptoms, but beyond a certain point (an apparent
threshold), the flux decreases more rapidly with enzyme content, and the res-
piration rate of the mitochondria becomes progressively more limited by the
step with the mutation.
Another potential area of application of Control Analysis in medicine is the
identification of particularly suitable sites for the manipulation of metabolism
with drugs. This is not a new idea; Hans Krebs141 advocated it in 1957,
but he considered that the answer would be to find the pacemaker enzymes.
The concept has been reformulated in Metabolic Control Analysis terms by
Chris Pogson and his group at the Wellcome pharmaceutical company.208
They reasoned that even though it was unlikely that true pacemaker, or rate–
limiting, enzymes existed, it would nevertheless still be worthwhile in the
drug–discovery process to target the enzymes with the highest flux control
coefficients. Very often, the aim with a drug is to inhibit a metabolic process,
such as an essential pathway in a parasite or cancer cell, and in this case, any
enzyme in the process is a potential target because every enzyme in a sequence
is essential for the sequence to work. Even so, the effects on metabolism are
likely to be obtained with lower concentrations of drug if an enzyme with a
high flux control coefficient is being inhibited, rather than one with a low
coefficient.
The area where Metabolic Control Analysis could make the biggest impact
in the near future is the area of biotechnology known as metabolic engineering.
Modern genetic engineering techniques have the potential to sidestep the tra-
ditional genetic approaches of selective breeding for organisms with a desired
characteristic by permitting directed, rational changes that could modify the
metabolism of organisms in a desired direction. Generally, increasing the flow
in a specific pathway is the more usual goal, and as discussed earlier in this
Chapter, increasing the flux is a more difficult problem than decreasing it. In
282 8 Conclusion
this area, Metabolic Control Analysis is already winning the argument over
traditional concepts of metabolic control, because so–called rate–limiting en-
zymes have already been over–expressed in cells with little effect. Identifying
enzymes with large flux control coefficients would allow a more rationally tar-
getted approach, but the two limitations of this approach have already been
discussed:
8.4 Summary
1. Control of flux cannot be exerted by one or a few enzymes at the be-
ginning of a pathway when large changes in pathway flux are seen to be
associated with much smaller relative changes in metabolite concentra-
tions.
4. Control Analysis alone may not be able to tackle all conceivable prob-
lems in regulation and control. Nevertheless, it is a better starting point
for aquiring a deeper understanding than the qualitative principles of
conventional biochemistry.
8.4 Summary 283
Further reading
1. Fell, D. A. (ed.) The control of flux: 21 years on. Biochem. Soc. Trans.
23, 341–391 (1995).
284
References
1. Albe, K. R., Butler, M. H. and Wright, B. E. (1990) J. Theor. Biol. 143, 163–195
2. Albe, K. R. and Wright, B. E. (1994) J. Theor. Biol. 169, 243–251
3. Anderson, L. E. (1986) Adv. Bot. Res. 12, 1–46
4. Asante, E. A., Hill, W. G. and Bulfield, G. (1991) Genet. Res. Camb. 58, 123–127
5. Ashworth, J. M. and Kornberg, H. L. (1963) Biochim. Biophys. Acta 73, 519–522
6. Atkinson, D. E. (1969) Curr. Top. Cell. Regul. 1, 29–43
7. Atkinson, D. E. (1977) Cellular Energy Metabolism and its Regulation., Academic
Press, New York
8. Atkinson, D. E. (1990) in Cornish-Bowden, A. and Cárdenas, M. L., eds., Control
of Metabolic Processes, pp. 3–11, New York, Plenum Press
9. Babul, J., Clifton, D., Kretschmer, M. and Fraenkel, D. G. (1993) Biochemistry 32,
4685–4692
10. Baranyi, J. M. and Blum, J. J. (1989) Biochem. J. 258, 121–140
11. Bartrons, R., Van Schaftingen, E., Vissers, S. and Hers, H.-G. (1982) FEBS Lett.
143, 137–140
12. Becker, G. L., Fiskum, G. and Lehninger, A. L. (1980) J. Biol. Chem. 255, 9009–9012
13. Beebe, S. J. and Corbin, J. D. (1986) in Boyer, P. D. and Krebs, E. G., eds., The
Enzymes, vol. XVII, pp. 43–111, Academic Press, New York, 3rd edn.
14. Benevolensky, S. V., Clifton, D. and Fraenkel, D. G. (1994) J. Biol. Chem. 269,
4876–4882
15. Bennett, J. (1991) Annu. Rev. Plant. Physiol. Plant Mol. Biol. 42, 281–311
16. Berry, M. N. and Friend, D. S. (1969) J. Cell Biol. 43, 506–520
17. Berthon, H. A., Bubb, W. A. and Kuchel, P. W. (1993) Biochem. J. 296, 379–389
18. Blackman, F. F. (1905) Ann. Bot. 19, 281
19. Blum, J. B. and Stein, R. B. (1982) in Goldberger, R., ed., Biological Regulation
and Development, vol. 3A, chap. 3, pp. 99–125, Plenum Press, New York
20. Brand, M. D. (1993) in Hochachka, P. W., Lutz, P. L., Sick, T., Rosenthal, M.
and van den Thillart, G., eds., Surviving Hypoxia: Mechanisms of Control and
Adaptation, chap. 20, pp. 295–309, CRC Press, Boca Raton
21. Brand, M. D., Hafner, R. P. and Brown, G. C. (1988) Biochem. J. 255, 535–539
22. Brindle, K. M. (1988) Biochemistry 27, 6187–6196
23. Brown, G. C., Lakin-Thomas, P. L. and Brand, M. D. (1990) Eur. J. Biochem. 192,
355–362
24. Bücher, T. and Rüssmann, W. (1964) Angew. Chem., Internat. Ed. 3, 426–439,
Originally published as Angew. Chem., 73, 881 in 1963.
25. Bulfield, G. (1972) Genet. Res. Camb. 20, 51–64
26. Cahill, G. F., Ashmore, J., Renold, A. E. and Hastings, A. B. (1959) Am. J. Med.
26, 264–282
27. Cárdenas, M. L. and Cornish-Bowden, A. (1989) Biochem. J. 257, 339–345
28. Carling, D. J., Zammit, V. A. and Hardie, D. G. (1987) FEBS Lett. 223, 217–222
29. Carling, D. J., Clarke, P. R., Zammit, V. A. and Hardie, D. G. (1989) Eur. J.
Biochem. 186, 129–136
30. Cascante, M., Canela, E. I. and Franco, R. (1990) Eur. J. Biochem. 192, 369–371
31. Cascante, M., Franco, R. and Canela, E. I. (1989) Math. Biosci. 94, 271–288
285
286 References
32. Cascante, M., Franco, R. and Canela, E. I. (1989) Math. Biosci. 94, 289–309
33. Chance, B., Holmes, W., Higgins, J. J. and Connelly, C. M. (1958) Nature (London)
182, 1190–1193
34. Chance, B. and Williams, G. R. (1955) J. Biol. Chem. 217, 409–427
35. Chance, B., Williams, G. R., Holmes, W. F. and Higgins, J. (1955) J. Biol. Chem.
217, 439–451
36. Chance, E. M., Seeholzer, S. H., Kobayashi, K. and Williamson, J. R. (1983) J. Biol.
Chem. 258, 13785–13794
37. Chelsky, D., Ruskin, B. and Koshland Jr., D. E. (1985) Biochemistry 24, 6651–6658
38. Chock, P. B., Shacter, E., Jurgensen, S. R. and Rhee, S. G. (1985) Curr. Top. Cell.
Regul. 27, 3–12
39. Clark, M. G., Bloxham, D. P., Holland, P. C. and Lardy, H. A. (1973) Biochem. J.
134, 589–597
40. Clark, M. G., Kneer, N. M., Bosch, A. L. and Lardy, H. A. (1974) J. Biol. Chem
249, 5695–5703
41. Cohen, G. N. (1969) Curr. Top. Cell. Regul. 1, 183–231
42. Cohen, P. (1985) Curr. Top. Cell. Regul. 27, 23–37
43. Cohen, P. (1989) Annu. Rev. Biochem. 58, 453–508
44. Cohen, P. and Chen, P. T. W. (1989) J. Biol. Chem. 264, 21435–21438
45. Cohen, P. and Klee, C. B., eds. (1988) Calmodulin. Molecular Aspects of Cellular
Regulation, Vol. 5., Amsterdam, Elsevier
46. Cohen, S. M. (1987) Biochemistry 26, 573–580
47. Cornish, A., Greenwood, J. A. and Jones, C. W. (1988) J. Gen. Microb. 88, 3111–
3122
48. Cornish-Bowden, A. (1989) J. Theor. Biol. 136, 365–377
49. Cornish-Bowden, A. (1995) Fundamentals of Enzyme Kinetics, Portland Press, Lon-
don, 2nd edn.
50. Cornish-Bowden, A. and Cárdenas, M. L. (1987) J. Theor. Biol. 124, 1–23
51. Cornish-Bowden, A. and Cárdenas, M. L. (1991) Trend Biochem. Sci. 16, 218–282
52. Cornish-Bowden, A. and Hofmeyr, J.-H. S. (1991) Comp. Appl. Biosci. 7, 89–93
53. Cornish-Bowden, A. and Hofmeyr, J.-H. S. (1994) Biochem. J. 298, 367–375
54. Crabtree, B. (1976) Biochem. Soc. Trans. 4, 999–1002
55. Crabtree, B. and Newsholme, E. A. (1985) Curr. Top. Cell. Regul. 25, 21–76
56. Crabtree, B. and Newsholme, E. A. (1987) Biochem. J. 247, 113–120
57. Crabtree, B. and Newsholme, E. A. (1987) Trends Biochem. Sci. 12, 4–12
58. Crawford, J. M. and Blum, J. J. (1983) Biochem. J. 212, 585–598
59. Davies, S. E. C. and Brindle, K. M. (1992) Biochemistry 31, 4729–4735
60. Denton, R. M. and Pogson, C. I. (1976) Metabolic Regulation., Chapman Hall,
London
61. Dykhuizen, D. E., Dean, A. M. and Hartl, D. L. (1987) Genetics 115, 25–31
62. Easterby, J. S. (1981) Biochem. J. 199, 155–161
63. Easterby, J. S. (1986) Biochem. J. 233, 871–875
64. Easterby, J. S. (1990) Biochem. J. 269, 255–259
65. Eisenthal, R. and Cornish-Bowden, A. (1974) Biochem. J. 139, 715–720
66. El-Yassin, D. I. and Fell, D. A. (1982) J. Mol. Biol. 156, 863–889
67. Endrenyi, L. and Chan, F.-Y. (1981) J. Theor. Biol. 90, 241–263
68. Fell, D. A. (1980) J. Theor. Biol. 84, 361–385
69. Fell, D. A. and Sauro, H. M. (1985) Eur. J. Biochem. 148, 555–561
70. Fell, D. A. and Sauro, H. M. (1990) Eur. J. Biochem. 192, 183–187
71. Fell, D. A. and Snell, K. (1988) Biochem. J. 256, 97–101
72. Fell, D. A. and Thomas, S. (1995) Biochem. J. 311, 35–39
73. Fiegelson, M. and Fiegelson, P. (1965) Adv. Enzyme Regul. 3, 11–27
74. Flint, H. J., Tateson, R. W., Bartelmess, I. B., Porteous, D. J., Donachie, W. D.
and Kacser, H. (1981) Biochem. J. 200, 231–246
75. Forsberg, H., Zetterqvist, O. and Engström, L. (1969) Biochim. Biophys. Acta 181,
171–175
76. Fraenkel, D. G. (1992) Annu. Rev. Genet. 26, 159–177
References 287
77. Galazzo, J. L. and Bailey, J. E. (1990) Enzyme Microb. Technol. 12, 162–172
78. Garfinkel, D. (1971) Comput. Biomed. Res. 4, 18–42
79. Garfinkel, D. (1981) Trends Biochem. Sci. 6, 69–71
80. Geiger, D. R. and Servaites, J. C. (1994) Annu. Rev. Plant Physiol. Plant Mol. Biol.
45, 235–256
81. Gellerich, F. N., Kunz, W. S. and Bohnensack, R. (1990) FEBS Lett. 274, 167–170
82. Gerhart, J. C. (1970) Curr. Top. Cell. Regul. 2, 275–325
83. Giersch, C. (1988) J. Theor. Biol. 134, 451–462
84. Giersch, C. (1994) J. Theor. Biol. 169, 89–99
85. Giersch, C. (1995) Eur. J. Biochem. 227, 194–201
86. Giersch, C., Lämmel, D. and Farquhar, G. (1990) Photosynth. Res. 24, 151–165
87. Giersch, C., Lämmel, D. and Steffen, K. (1990) in Cornish-Bowden, A. and Cárdenas,
M. L., eds., Control of Metabolic Processes, pp. 351–361, New York, Plenum Press
88. Goldbeter, A. and Koshland, D. E. (1981) Proc. Natl. Acad. Sci. U.S.A. 78, 6840–
6844
89. Groen, A. K. (1984) Quantification of control in studies on intermediary metabolism.,
Ph.D. thesis, University of Amsterdam
90. Groen, A. K., van der Meer, R., Westerhoff, H. V., Wanders, R. J. A., Akerboom,
T. P. M. and Tager, J. M. (1982) in Sies, H., ed., Metabolic Compartmentation, pp.
9–37, Academic Press, London
91. Groen, A. K., van Roermund, C. W. T. Vervoorn, R. C. and Tager, J. M. (1986)
Biochem. J. 237, 379–389
92. Groen, A. K., Wanders, R. J. A., Westerhoff, H. V., van der Meer, R. and Tager,
J. M. (1982) J. Biol. Chem. 257, 2754–2757
93. Haber, J. E. and Koshland, D. E. (1967) Proc. Natl. Acad. Sci. U.S.A. 58, 2087–2093
94. Hafner, R. P., Brown, G. C. and Brand, M. D. (1990) Eur. J. Biochem. 188, 313–319
95. Hales, C. N. (1967) in Campbell, P. N. and Greville, G. D., eds., Essays in Biochem-
istry, vol. 3, pp. 73–104, The Biochemical Society / Academic Press, London
96. Heinisch, J. (1986) Mol. Gen. Genet. 202, 75–82
97. Heinrich, R. (1985) Biomed. Biochim. Acta 44, 913–927
98. Heinrich, R. (1990) in Cornish-Bowden, A. and Cárdenas, M. L., eds., Control of
Metabolic Processes, pp. 329–342, New York, Plenum Press
99. Heinrich, R. and Rapoport, T. A. (1974) Eur. J. Biochem. 42, 89–95
100. Heinrich, R. and Rapoport, T. A. (1974) Eur. J. Biochem. 42, 97–105
101. Helmreich, E. and Cori, C. F. (1965) Adv. Enzyme Regul. 3, 91–107
102. Hershko, A. and Ciechanover, A. (1992) Annu. Rev. Biochem. 61, 761–807
103. Hess, B. (1977) Trends Biochem. Sci. 2, 193–195
104. Higgins, J. (1963) Ann. N. Y. Acad. Sci. 108, 305–321
105. Hillgartner, F. B., Salati, L. M. and Goodridge, A. G. (1995) Physiol. Rev. 75, 47–76
106. Hofmeyr, J.-H. S. and Cornish-Bowden, A. (1991) Eur. J. Biochem. 200, 223–236
107. Hohorst, H. J., Reim, M. and Bartels, H. (1962) Biochem. Biophys. Res. Commun.
7, 137–141
108. Holzhütter, H.-G., Jacobasch, G. and Bisdorff, A. (1985) Eur. J. Biochem. 149,
101–111
109. Holzhütter, H.-G., Schuster, R., Buckwitz, D. and Jacobasch, G. (1990) Biomed.
Biochim. Acta 49, 791–800
110. Huber, S. C., Huber, J. L. and McMichael, R. W. (1994) Int. Rev. Cytol. 149, 47–98
111. Hue, L. (1981) Adv. Enzymol. 52, 247–331
112. Hue, L. (1982) in Sies, H., ed., Metabolic Compartmentation, pp. 71–97, Academic
Press, London
113. Jensen, P. R., Michelsen, O. and Westerhoff, H. V. (1993) Proc. Nat. Acad. Sci.
U.S.A. 90, 8068–8072
114. Jensen, P. R., Westerhoff, H. V. and Michelsen, O. (1993) EMBO J. 12, 1277–1282
115. Jensen, P. R., Westerhoff, H. V. and Michelsen, O. (1993) Eur. J. Biochem. 211,
181–191
116. Kacser, H. (1983) Biochem. Soc. Trans. 11, 35–40
117. Kacser, H. (1995) Biochem. Soc. Trans. 23, 387–391
288 References
161. Monod, J., Wyman, J. and Changeux, J.-P. (1965) J. Mol. Biol. 12, 88–118
162. Morgan, H. E., Randle, P. J. and Regen, D. M. (1959) Biochem. J. 73, 573–579
163. Murphy, M. P. and Brand, M. D. (1987) Biochem. J. 243, 499–505
164. Narabayashi, H., Lawson, J. W. R. and Uyeda, K. (1985) J. Biol. Chem. 260, 9750–
9758
165. Neuhaus, H. E., Kruckeberg, A. L., Feil, R. and Stitt, M. (1989) Planta 178, 110–12
166. Neuhaus, H. E., Quick, W. P., Siegl, G. and Stitt, M. (1990) Planta 181, 583–592
167. Neuhaus, H. E. and Stitt, M. (1990) Planta 182, 445–454
168. Newsholme, E. A. and Crabtree, B. (1976) Biochem. Soc. Symp. 41, 61–109
169. Newsholme, E. A. and Gevers, W. (1967) Vitamins and Hormones 25, 1–87
170. Newsholme, E. A. and Randle, P. J. (1964) Biochem. J. 93, 641–651
171. Newsholme, E. A. and Start, C. (1973) Regulation in Metabolism., Wiley and Sons,
London
172. Newsholme, E. A. and Underwood, A. H. (1966) Biochem. J. 99, 24C–26C
173. Newsholme, P. and Walsh, D. A. (1992) Biochem. J. 283, 845–848
174. Nicholls, D. G. and Crompton, M. (1980) FEBS Lett. 111, 261–268
175. Niederberger, P., Prasad, R., Miozzari, G. and Kacser, H. (1992) Biochem. J. 287,
473–479
176. Norby, J. G., Ottolenghi, P. and Jensen, J. (1980) Analyt. Biochem. 102, 318–320
177. Orr, H. A. (1991) Proc. Natl. Acad. Sci. U.S.A. 88, 11413–11415
178. Ovadi, J. (1991) J. Theor. Biol. 152, 1–22, See also the following discussion by
various authors on pp23–141.
179. Page, R. A., Kitson, K. E. and Hardman, M. J. (1991) Biochem. J. 278, 659–665
180. Pearcy, R. W. (1990) Annu. Rev. Plant Physiol. Plant Mol. Biol. 41, 421–453
181. Pedersen, R. C. and Brownie, A. C. (1986) Biochem. Action Horm. 13, 129–166
182. Pettersson, G. and Ryde-Pettersson, U. (1988) Eur. J. Biochem. 175, 661–672
183. Pietrobon, D., Di Virgilio, F. and Pozzan, T. (1990) Eur. J. Biochem. 193, 599–622
184. Pilkis, S. J. and Claus, T. H. (1991) Annu. Rev. Nutr. 11, 465–515
185. Pilkis, S. J., Claus, T. H. and El-Maghrabi, M. R. (1988) Adv. Second Mess. Phos-
phoprotein Res. 22, 175–191
186. Przybylski, F., Otto, A., Nissler, K., Schellenberger, W. and Hofmann, E. (1985)
Biochim. Biophys. Acta 831, 350–352
187. Quant, P. A. and Makins, R. A. (1994) Biochem. Soc. Trans. 22, 441–446
188. Quant, P. A., Robin, D., Robin, P., Girard, J. and Brand, M. D. (1993) Biochim.
Biophys. Acta 1156, 135–143
189. Quick, W. P., Schurr, U., Scheibe, R., Schulze, E.-D., Rodermel, S. R., Bogorad, L.
and Stitt, M. (1991) Planta 183, 542–554
190. Rabkin, M. and Blum, J. J. (1985) Biochem. J. 255, 761–786
191. Rainey, W. E., Shay, J. W. and Mason, J. I. (1986) J. Biol. Chem. 261, 7322–7326
192. Randle, P. J., England, P. J. and Denton, R. M. (1970) Biochem. J. 117, 677–695
193. Rapoport, T. A., Heinrich, R., Jacobasch, G. and Rapoport, S. (1974) Eur. J.
Biochem. 42, 107–120
194. Rapoport, T. A., Heinrich, R. and Rapoport, S. M. (1976) Biochem. J. 154, 449–469
195. Reder, C. (1988) J. Theor. Biol. 135, 175–201
196. Reich, J. G. and Sel’kov, E. E. (1981) Energy Metabolism of the Cell, Academic
Press, London
197. Rhee, S. G., Chock, P. B. and Stadtman, E. R. (1989) Adv. Enzymol. 62, 37–92
198. Rigoulet, M., Leverve, X. M., Plomp, P. J. A. M. and Meijer, A. J. (1987) Biochem.
J. 245, 661–668
199. Rodbell, M. (1992) Curr. Top. Cell. Regul. 32, 1–47
200. Rognstad, R. (1979) J. Biol. Chem. 254, 1875–1878
201. Rolleston, F. S. (1972) Curr. Topics Cell Regul. 5, 47–75
202. Rudolph, F. B. and Fromm, H. J. (1971) J. Biol. Chem. 246, 6611–6619
203. Ruijter, G. J. G., Postma, P. W. and van Dam, K. (1991) J. Bacteriol. 173, 6184–
6191
204. Sacktor, B. and Wormser-Shavit, E. (1966) J. Biol. Chem. 241, 624–631
205. Saier, M. H., Wu, L.-F. and Reizer, J. (1990) Trends Biochem. Sci. 15, 391–395
290 References
253. Van Schaftingen, E., Hue, L. and Hers, H. G. (1980) Biochem. J. 192, 887–895
254. Van Schaftingen, E., Hue, L. and Hers, H. G. (1980) Biochem. J. 192, 897–901
255. Van Schaftingen, E., Jett, M.-F., Hue, L. and Hers, H.-G. (1981) Proc. Natl. Acad.
Sci. U.S.A. 78, 3483–3486
256. VanderCammen, A. and Van Schaftingen, E. (1990) Eur. J. Biochem. 191, 483–489
257. Vandercammen, A. and Van Schaftingen, E. (1993) Biochem. J. 294, 551–556
258. Veech, R. L., Lawson, J. W. R., Cornell, N. W. and Krebs, H. A. (1979) J. Biol.
Chem. 254, 6538–6547
259. Voit, E. O. and Savageau, M. A. (1987) Biochemistry 26, 6869–6880
260. Waley, S. G. (1964) Biochem. J. 91, 514–517
261. Wals, P. A. and Katz, J. (1994) J. Biol. Chem. 269, 18343–18352
262. Walsh, K. and Koshland Jr., D. E. (1985) Proc. Natl. Acad. Sci. U.S.A. 82, 3577–
3581
263. Walsh, K., Schena, M., Flint, A. J. and Koshland Jr., D. E. (1987) Biochem. Soc.
Symp. 54, 183–195
264. Walter, R. P., Morris, J. G. and Kell, D. B. (1987) J. Gen. Microbiol. 133, 259–266
265. Wanders, R. J. A., van Roermund, C. W. T. and Meijer, A. J. (1984) Eur. J.
Biochem. 142, 247–254
266. Wang, J. H., Pallen, C. J., Sharma, R. K., Adachi, A. M. and Adachi, K. (1985)
Curr. Top. Cell. Regul. 27, 419–436
267. Waterman, M. R. and Simpson, E. R. (1989) Recent Prog. Horm. Res. 45, 533–566
268. Weber, G., Singhal, R. L. and Srivastava, S. K. (1965) Adv. Enzyme Regul. 3, 43–75
269. Weber, G., Singhal, R. L., Stamm, N. B., Lea, M. A. and Fisher, E. A. (1966) Adv.
Enzyme Regul. 4, 59–81
270. Weekes, J., Hawley, S. A., Corton, J., Shugar, D. and Hardie, D. G. (19994) Eur. J.
Biochem. 219, 751–757
271. Welch, G. R. (1985) J. Theor. Biol. 114, 433–446
272. Westerhoff, H. V. and Chen, Y.-D. (1984) Eur. J. Biochem. 142, 425–430
273. Westerhoff, H. V., Groen, A. K. and Wanders, R. J. A. (1984) Biosci. Rep. 4, 1–22
274. Westerhoff, H. V. and Kell, D. B. (1987) Biotechnol. Bioeng. 30, 101–107
275. Wilkinson, G. N. (1961) Biochem. J. 80, 324–332
276. Woodrow, I. E. (1986) Biochim. Biophys. Acta 851, 181–192
277. Woodrow, I. E., Murphy, D. J. and Latzko, E. (1984) J. Biol. Chem. 259, 3791–3795
278. Wright, B. E. and Albe, K. R. (1994) J. Theor. Biol. 169, 231–241
279. Wright, B. E., Butler, M. H. and Albe, K. R. (1992) J. Biol. Chem. 267, 3101–3105
280. Wright, B. E. and Reimers, J. M. (1988) J. Biol. Chem. 263, 14906–14912
281. Yates, R. A. and Pardee, A. B. (1956) J. Biol. Chem. 221, 757–770
282. Zhang, Y., Proenca, R., Maffei, M., Barone, M., Leopold, L. and Friedman, J. M.
(1994) Nature (London) 372, 425–432
283. Zuurendonk, P. F. and Tager, J. M. (1974) Biochim. Biophys. Acta 333, 393–399
284. Zuurendonk, P. F., Tischler, M. E., Akerboom, T. P. M., van der Meer, R.,
Williamson, J. R. and Tager, J. M. (1979) Methods Enzymol. 56, 207–233
292
Index
293
294 Index
response coefficient and, 125, 171 substrate binding, 15, 58, 62–64, see
substrate also binding sites
irreversible reaction, 116–117, 135 substrate complex, 48
reversible reaction, 117 subunit structure, 66, 71, 73, 76,
electron transport chain, 31, 95, 162 77, 81
intermediate, 272 titration, see flux control coefficient
electrophoresis enzyme kinetics, 47–85
cell proteins, 232 induced fit, 58, 76
isoenzymes, 143 inhibition, see inhibitor
endoplasmic reticulum, 30, 90, 223, 228, initial rate, 8, 48, 55
270, 271 limitations of, 5–6, 169, 191
energy maximal activity, 129
dissipation, 88, 218, 227 mechanism, 57–59
metabolism, 189–190, 204, 240 compulsory–order, 58, 62–63
transduction, 149 double–displacement, 58–59, 64–
enolase, 93, 177 65
disequilibrium, 89, 90 random–order, 58, 63–64
elasticity, 122, 176, 178, 179 Michaelis Menten, 47–56
computer fitting, 53–54, 83
flux control coefficient, 122, 175, 176,
178–180 equation, 48
graphical analysis, 50–54
low activity, 93
reversible, 55
enzyme
multi–enzyme sequence, 103, 106
activity, 5, 24, 92–94
optimal design, 54, 60
changing, 2, 8, 9, 18, 94, 99, 124,
partial reactions, 65
269
permeabilized cells, 29
large changes, see flux, large changes
product inhibition, 17, 39, 55–56,
measurement, 37–39
64
metabolic flux and, see flux con-
importance, 169, 192
trol coefficient
rate equation, 134, 135
allosteric, see allosteric enzymes
two–substrate, 56–65, 97
assay, 32–33
computer fitting, 61, 83
coupled, 32, 33 experiment, 60
association, 132, 169, 278, 280 graphical analysis, 60–61
catalysing several steps, 278 rate equation, 59–60
covalent modification, see covalent enzyme multiplicity, 215
modification feedback inhibition, 214–215
degradation, 8, 230 enzyme IIGlc
elasticity, see elasticity flux control coefficient, 148
expression epididymal fat pad, 26
controllable, 147–149, 151, 152 epinephrin, see adrenaline
high in vivo concentration, 280 equilibrium, 13
isolated, 30 chemical, 10, 13
mutations, 140, 141, 170, 281, see constant, 14, 55, 66, 73, 77, 128
dominance in vivo correction, 89–90
near–equilibrium, see equilibrium, displacement from, see disequilib-
near–equilibrium reaction rium ratio
non–equilibrium, see equilibrium, non– dynamic, 10, 13, 228
equilibrium reaction near–equilibrium reaction, 15–17, 37,
phosphorylation, see covalent mod- 48, 64, 88, 90–92, 117, 122,
ification 129, 173, 174, 192, 193, 204,
production, 2 222
regulatory, 19–20, 87–100, 210, 275, non–equilibrium reaction, 15–17, 19,
276 88–92, 98, 129
mutant, 99 erythrocyte, see blood, cell, red
properties, 98, 99, 261 Escherichia coli, 8, 94, 99, 144, 147–149,
reversibility, 55 152, 205, 206, 214–216, 245,
298 Index
constitutive, 145 32 P, 41
parameter, 50, 59, 75, 124, 124, 125, 126, phosphohydroxypyruvate, 193
131 phosphoprotein phosphatase, 233, 241–
Pardee, A. B., 205 244, 254
partial reactions, see enzyme kinetics 1, 242, 243, 245
Passonneau, J. V., 90, 93 2A, 242
Pasteur, L., 97 2B, 242, 243, 245
PEPCK, 94, 160, 172, 269 calmodulin, 237, 243
flux control coefficient, 160, 168, 175– 2C, 242
177 phosphoribulokinase, 273
light activation, 272, 273 phosphorylase, see glycogen phosphory-
rate–limiting?, 156, 173 lase
perchloric acid, 31, 36 phosphorylase kinase, 231–233, 237, 243–
perfusion, see heart, organ 245, 254
permeabilized cells, see cell, permeabi- calmodulin, 237
lized phosphorylation, protein, see covalent mod-
Perutz, M., 80 ification
phenotype, 108, 144, 266 phosphoserine
phenylalanine, 168 aminotransferase, 193
catabolism phosphatase, 193, 195
control analysis, 171 flux control coefficient, 194
synthesis, 213 phosphotransferase system, see phospho-
transport enolpyruvate
flux control coefficient, 196 photorespiration, 149
phenylalanine–4–monooxygenase photosynthesis, 266, 271–274
phosphorylation, 236 control analysis, 145, 149–151, 181
phenylsuccinate, 163 computer models, 170
phosphodiesterase, see cyclic–nucleotide dark reactions, 272
phosphodiesterase large flux changes, 271
phosphoenolpyruvate light activation, 273–274
assay, 32 light reactions, 272
phosphotransferase system, 148 oscillations, 208
phosphoenolpyruvate carboxykinase, see photosystem II, 272
PEPCK phosphorylation, 272
phosphoenolpyruvate:pyruvate cycle, 42, ping–pong, see enzyme kinetics, mecha-
174, 221, 223 nism, double–displacement
phosphofructo–2–kinase, see bifunctional plasmid, 94, 146–149, 264, 265
enzyme pleiotropic effects, 146
phosphofructokinase, 38, 93, 221, 268, 269 Pogson, C. I., 19, 281
activation, see fructose 2,6–bisphosphate polymorphism, 144
AMP activation, 127, 208, 223, 225 Postma, P. W., 148
ATP inhibition, 208, 236 power law, 109
cross–over, 97, 210 pregnenolone, 270, 271
disequilibrium, 89, 90 Prigogine, I., 13
flux control coefficient, 154, 155, 195 product inhibition, see enzyme kinetics
non–inhibited, 206 promoter, see lac, tac
over–expression, 104, 147 protein kinase A, 36, 221, 223, 233–236,
phosphorylation, 98, 236, 266, 270 241–243, 245
product activation, apparent, 38, 223 cooperativity, 235
rate–limiting?, 99, 147 specificity, 235
phosphoglucomutase, 89, 90, 93, 145 Types I & II, 235
phosphoglucoseisomerase, see glucose–6– protein kinase C, 238
phosphate isomerase protein kinase, redox–dependent, 272
phosphoglycerate dehydrogenase, 193 protein phosphatase, see phosphoprotein
phosphoglycerate kinase, 93, 172, 176 phosphatase
phosphoglycerate mutase, 80, 89, 90, 172, proteolysis, 230, see enzyme degradation
177, 179 prothrombin, 230
flux control coefficient, 178–180 proton leak, 161, 162, 182, 183, 185
304 Index