You are on page 1of 28

PHYSIOLOGICAL REVIEWS

Vol. 77, No. 3, July 1997


Printed in U.S.A.

Cellular Energy Utilization and Molecular Origin


of Standard Metabolic Rate in Mammals
DAVID F. S. ROLFE AND GUY C. BROWN

Department of Biochemistry, University of Cambridge, Cambridge, United Kingdom

I. Introduction 732
II. Quantitative Assessment of the Origin of Standard Metabolic Rate in Mammals 733
A. Contribution of different tissues to standard metabolic rate 734
B. Contribution of non-ATP-consuming reactions in the major oxygen-consuming tissues to standard
metabolic rate 735
C. Contribution of ATP-consuming reactions in the major oxygen-consuming tissues to standard
metabolic rate 737
D. Service functions 741
E. Relative ATP use by cellular ATPases 742
F. Reactions that uncouple standard metabolism 742
G. Substrate utilization and glycolytic ATP production in standard metabolic rate 743
H. Summary 743
III. Contribution of Feeding, Cold, and Muscle Use to Field and Maximal Metabolic Rate 744
A. Cold-induced thermogenesis 744
B. Feeding-induced thermogenesis 745
C. Muscle use-induced thermogenesis 745
IV. Heat Production and Free Energy Dissipation 745
A. Thermogenesis and enthalpy dissipation 745
B. Free energy dissipation and efficiency 747
C. Distribution of energy dissipation 748
V. Changes and Differences in Metabolic Rate 749
A. Control of cellular respiration 749
B. Causes of species differences in standard metabolic rate related to body size and phylogeny 750
C. Summary 751
VI. Functions of Standard Metabolic Rate 751
A. Maintenance of cellular composition against undesigned uncoupling reactions 751
B. Standard metabolic rate as a means of heat production 752
C. High standard metabolic rate enables high maximal and field metabolic rate 752
D. Standard metabolic rate increases the potential for regulation 753
E. Standard metabolic rate as a means for regulating harmful free radical production 753
F. Summary 753
VII. Conclusions 753

Rolfe, David F. S., and Guy C. Brown. Cellular Energy Utilization and Molecular Origin of Standard Metabolic
Rate in Mammals. Physiol. Rev. 77: 731- 758,1997. - The molecular origin of standard metabolic rate and thermogen-
esis in mammals is examined. It is pointed out that there are important differences and distinctions between the
cellular reactions that 1) couple to oxygen consumption, 2) uncouple metabolism, 3) hydrolyze ATP, 4) control
metabolic rate, 5) regulate metabolic rate, 6) produce heat, and ?‘) dissipate free energy. The quantitative contribution
of different cellular reactions to these processes is assessed in mammals. We estimate that -90% of mammalian
oxygen consumption in the standard state is mitochondrial, of which -20% is uncoupled by the mitochondrial
proton leak and 80% is coupled to ATP synthesis. The consequences of the significant contribution of proton
leak to standard metabolic rate for tissue P-to-O ratio, heat production, and free energy dissipation by oxidative
phosphorylation and the estimated contribution of ATP-consuming processes to tissue oxygen consumption rate
are discussed. Of the 80% of oxygen consumption coupled to ATP synthesis, -25-30% is used by protein synthesis,
19-28% by the Na’-K’-ATPase, 4-8% by the Ca2+-ATPase, Z-8% by the actinomyosin ATPase, 7- 10% by gluconeogen-
esis, and 3% by ureagenesis, with mRNA synthesis and substrate cycling also making significant contributions. The
main cellular reactions that uncouple standard energy metabolism are the Na+, K’, H+, and Ca2’ channels and
leaks of cell membranes and protein breakdown. Cellular metabolic rate is controlled by a number of processes

0031-9333/97 $15.00 Copyright 0 1997 the American Physiological Society 731

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
732 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

including metabolic demand and substrate supply. The differences in standard metabolic rate between animals of
different body mass and phylogeny appear to be due to proportionate changes in the whole of energy metabolism.
Heat is produced by some reactions and taken up by others but is mainly produced by the reactions of mitochondrial
respiration, oxidative phosphorylation, and proton leak on the inner mitochondrial membrane. Free energy is
dissipated by all cellular reactions, but the major contributions are by the ATP-utilizing reactions and the uncoupling
reactions. The functions and evolutionary significance of standard metabolic rate are discussed.

I. INTRODUCTION imal to SMR are seen; for example, a trained athlete’s


maximal metabolic rate may be 20 x SMR, whereas in an
untrained human a value of 12 x SMR would be more
Mammalian cells and organisms use a considerable
usual (15).
amount of energy in the basal or standard state, when no
net work is done and all the free energy is dissipated. The energy use in the standard state is sometimes
This review examines where this energy goes, what con- regarded as a fixed requirement for maintenance of the
trols the energy usage, and what possible functions this organism, to which other energy uses may be added. How-
basal energy usage might serve. ever, the extra energy use induced by cold, feeding, mus-
Standard metabolic rate (SMR) is the steady-state cle use, growth, or reproduction is not necessarily addi-
rate of heat production by a whole organism under a set tive with SMR. This is because the metabolic processes
of “standard” conditions; in mammals, these conditions underlying SMR may decrease or increase in the condi-
are that the individual is an adult and is awake but resting, tions of extra energy demand (see Ref. 216). For example,
stress free, not digesting food (prior food intake being at during exercise, the metabolic rate of some organs (e.g.,
or around maintenance level), and maintained at a temper- kidney and gut) decreases, whereas gluconeogenesis in-
ature that elicits no thermoregulatory effect on heat pro- creases in the liver, and protein synthesis may decrease
duction (see Ref. 15 for review). Standard metabolic rate in skeletal muscle. Thus the processes underlying SMR
is measured either directly as heat production, or indi- cannot be regarded as fixed, but rather change with condi-
rectly as oxygen consumption from which it can be accu- tions.
rately predicted (189). Basal metabolic rate is similar to The standard and field metabolic rates are important
SMR, although the term is now most usually applied to because they determine the calorific nutrient require-
human metabolism only. Resting metabolic rate is also ments and growth rates of animals and humans. Not all
similar to SMR, except that the metabolic rate is measured the nutrient energy intake of an individual organism is
while the animal is still digesting food. lost as heat, since a proportion is lost as feces and urine
Metabolic rate changes under different conditions, (-4 and 5%, respectively, in humans) (114), a proportion
and (at least) three other conditions are generally distin- is saved as growth of the individual or passed on during
guished: minimal metabolic rate, field metabolic rate, and birth or lactation, and a small proportion is saved as work
maximal metabolic rate. Minimal metabolic rate is the on the environment (Fig. 1; see Refs. 15 and 114 for energy
metabolic rate in various conditions that cause the rate requirements). However, the majority of an individual’s
to fall below that of SMR; causes of this depression in- calorific intake is lost as heat, and thus accounted for by
clude sleep, anesthesia, and continued starvation. Meta- metabolic rate. This heat production is due to the cellular
bolic rate falls by -10% below SMR during sleep and metabolism of all the tissues of an individual organism.
may fall by 40% (a decrease of 20% in metabolic rate per In an adult organism that is not growing, reproducing,
kilogram body weight) during long-term starvation in hu- lactating, or doing work on the environment, all the energy
mans (15). Field metabolic rate is the average rate over taken in and metabolized is lost as heat by cellular meta-
an extended period of time when the animal is living in bolic reactions. All the free energy is dissipated, and all
its natural environment and may be assessedeither by the chemical potential energy is converted to heat.
direct measurements of heat production (as is often the In this article we address the following questions with
case in human studies) or by the doubly labeled water regard to mammalian SMR. 1) What is the quantitative
method. Field metabolic rate is higher than SMR as a contribution of the major organs and tissues to SMR? 2)
result of the energy requirements made by feeding, cold What is the quantitative contribution of the major meta-
exposure, and muscle use (Fig. 1). Each of these condi- bolic pathways and cellular processes to SMR? 3) What
tions can increase metabolic rate above the standard rate, are the main ATP-using reactions in mammalian cells, and
and the extent depends on conditions. The maximal meta- what is their contribution to total ATP consumption in a
bolic rate is the maximum steady-state metabolic rate at- cell? -4) At which cellular reactions is the heat lost, and
tainable during hard exercise over a defined period of at which reactions is the free energy dissipated? 5) Which
time and is often considered to be simply 10 times SMR reactions control SMR, and which are responsible for
in mammals. However, large variations in the ratio of max- physiological changes in SMR? 6) What is the function of

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 733

FIG. 1. Schematic illustration of how


Total Total Standard
energy taken into human body is used.
energy energy metabolic
First column represents total energy in-
used loss rate take, second column total energy used,
and third column total heat produced.
Proportions given are for adult humans
Total Total and are only roughly to scale. Note that
Energy ---e-B--
Heat field metabolic rate represents daily (24
Intake output h) energy expenditure and does not in-
clude energy required for reproduction.
Muscle Contribution to total heat output arising
use from cold-induced thermogenesis will ob-
viously depend on ambient temperature
-----I at which metabolic rate measurements
-Growth \\ -------- are made (see text for explanation). [Data
w--B--- ----- \\ from Blaxter (15).]
\
\ - Repro- Food
Urine & \
duction \\
\
Feces
\
\
\\ --------
‘4 Cold
“4

SMR? We first assessthe contribution of a tissue or pro- oxygen consumption occurs at all in the standard state is
cess to SMR by quantifying its coupling to oxygen con- related to processes that “uncouple” metabolism (Fig. a>.
sumption. Standard metabolic rate cannot be equated with The cell consists of a number of low-entropy structures
oxygen consumption, but oxygen consumption accurately (for example, macromolecules and ion gradients) whose
predicts SMR, and the extent to which a process is cou- production requires free energy. In animals, this free en-
pled to oxygen consumption quantifies the energy flux ergy comes from the oxidation of food molecules: carbo-
through the process. hydrates, fats, and proteins. This oxidation is coupled to
the reduction of NAD. Oxidation of NADH by the mito-
chondrial electron transport chain is coupled to the pro-
II. QUANTITATIVE ASSESSMENT OF THE
duction of an electrochemical proton gradient across the
ORIGIN OF STANDARD METABOLIC RATE
IN MAMMALS mitochondrial inner membrane. Synthesis of ATP is cou-
pled to the “downhill” channeling of these protons
Metabolism has often been conceptually divided into through the ATP synthase. Processes such as protein syn-
the reactions that produce energy or ATP (catabolism) thesis, maintenance of sodium and potassium gradients,
and the reactions that use energy or ATP (anabolism). and muscular contraction are coupled to ATP hydrolysis
This has been a useful dichotomy, but is misleading in (see Fig. 2). However, if this was all that there was to
several ways. 1) The processes of ATP production are not metabolism, then the whole system would come to equi-
energy producing (as according to the first and second librium and there would be no metabolic rate. There is a
laws of thermodynamics energy is conserved and free significant flux through the system because of reactions
energy dissipated by all reactions), but rather function to that uncouple metabolism, such as protein degradation,
distribute free energy. 2) Free energy is distributed to ion leaks, and muscle relaxation. Thus, in a sense, the
energy-requiring reactions using intermediates other than uncoupling processes account for SMR.
the ATP, for example, GTP, NADH, NADPH, the mitochon- As mentioned above, energy metabolism has pre-
drial proton gradient, and the plasma membrane Na’ gra- viously been thought to be fully coupled to ATP produc-
dient. 3) The ATP-using reactions are wrongly assumed tion, but it is now recognized that 1) not all oxygen con-
to be responsible for all the heat production and free sumption is mitochondrial and therefore coupled to oxida-
energy dissipation of metabolism and to be solely respon- tive phosphorylation, and 2) a significant proportion of
sible for control of metabolic rate. mitochondrial respiration is not coupled to ATP synthesis.
Because ATP utilization has been equated with cellu- Hence, some of the reactions responsible for SMR occur
lar energy utilization, the origin of SMR has been thought before oxidative ATP synthesis. Thus energy metabolism
to be solvable simply by identifying the major reactions can be regarded as a leaky hose pipe (Fig. Z), and in the
responsible for ATP utilization. However, for the purposes standard metabolic state, all the energy must leak out
of locating the origin of SMR, it will be more useful to somewhere, since none of the energy is conserved, rather
divide metabolism into coupling and uncoupling reac- all is lost as heat. However, the site of heat production
tions. Because no net work is done by SMR, the fact that cannot simply be identified with either the uncoupling or

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
734 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

Proteins
I
0 *
Amino acids
Substrate NADH, H’out
I Na+out x out
I
I
I I I
v V
V Na+in x X in
e2+H2O x NAD x H+LIl ADP
Actinomyosin
+ Pi (contracted)
H;O I
Actinzmyosin
(relaxed)

FIG. 2. Simplified scheme of cellular energy metabolism in standard state. Curved unbroken arrows represent
some of coupling reactions of energy metabolism. Broken arrows indicate some of uncoupling reactions. Depicted
coupling reactions are dehydrogenation, mitochondrial proton pumping, and ATP synthesis and ATP consumption by
protein synthesis and Na’-K’-ATPase plus Na’-linked transport of many different substances (X in this figure) across
plasma membrane. Note that transport of X may be via Na+ antiport as shown (e.g., if X = Ca2’), or by Na+ symport
(e.g., if X = amino acids). Uncoupling reactions are mitochondrial proton leak, protein degradation, and Na+ and X
leaks at plasma membrane. Note that uncoupling of substrate dehydrogenation may also occur, for example, due to
peroxisomal fatty acid oxidation.

coupling reactions, because both types of reaction pro- standard metabolism have been measured by various meth-
duce heat. Also, heat production cannot be equated with ods. In vitro measurements are made on perfused organs
free energy dissipation, because these are different pro- or tissue slices; in vivo estimates are made from measure-
cessesthat must be evaluated separately. In addition, we merits of arteriovenous differences in oxygen concentration
cannot equate the uncoupling reactions with the reactions and of blood flow. However, even in studies where tissue
that control metabolic rate, since control is also located oxygen consumption rates were measured in vivo, the total
in coupling reactions. oxygen consumption of the organs does not always add
Different tissue preparations from different animals, up to the measured body oxygen consumption, and this
in different metabolic conditions, have been used to esti- presumably reflects technical difficulties in making the mea-
mate the relative contribution of processes to tissue SMR. surements (1, 85, 183). An estimate of the contribution of
Thus, in using these data to estimate the overall contribu- various organs to the oxygen consumption of an adult hu-
tion of a process to SMR, care must be taken in extrapolat- man and rat in the standard metabolic state is given in
ing to the in vivo metabolic rate. Furthermore, it is essen- Table 1. The relative contribution of these organs to body
tial to distinguish between the metabolic rate of isolated mass and SMR varies somewhat with species (183). No
tissues or cells in a minimal medium and the rate under one organ is responsible for the majority of the minimal
standard physiological conditions. The former rate may be metabolic rate, but some organs (brain, kidney, heart, and
lower than the latter, because, for example, in a minimal gastrointestinal tract) contribute a much larger fraction of
medium isolated brain and heart will be missing electrical the total SMR than their fractional mass or volume of the
excitation, kidney tissue will lack metabolites to trans- body, whereas others (bone, white adipose tissue, skin, and
port, and liver will be missing substrates for glucose, urea, muscle) contribute much less. Note, however, that some
and protein synthesis. However, if care is taken, oxygen tissues (e.g., skeletal muscle) make a significant contribu-
consumption rates measured on tissue slices in vitro are tion to SMR by virtue of their size, even though their meta-
not dramatically different from those of the same tissues bolic rate per gram is relatively low. Thus the pathways
in vivo. For example, organ metabolic rates estimated contributing to the metabolic rate of these tissues are often
from tissue slice data (e.g., 37 pmol O2 min-’ whole l l of more relevance to the study of SMR than those in more
liver-’ and 0.606 ,umol O2 min-’ g skeletal muscle-‘; Ref.
l l metabolically active tissues such as kidney and brain.
71) are similar to in vivo measurements of tissue meta- Further geographical localization of oxygen con-
bolic rate (e.g., 61 pmol O2 rnin whole liver-‘, Ref. 116;
l l sumption and thus heat production has been attempted
and 0.787 pmol O2 min-’ g skeletal muscle-‘, Ref. 173).
l l in some tissues by measuring oxygen consumption or Z-
deoxyglucose uptake, or mitochondrial or cytochrome ox-
A. Contribution of Different Tissues to Standard idase density (which are thought to be proportional to
Metabolic Rate metabolic rate). The metabolic rate in brain as estimated
The oxygen consumption rates and thus heat produc- by Z-deoxyglucose uptake is 25- 50%lower in white matter
tion by the various organs of the mammalian body during than in gray matter, and cytochrome oxidase, mitochon-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 735

TABLE 1. Contribution of the major oxygen-consuming acid oxidation. Estimates of nonmitochondrial oxygen
organs of the body to body mass and standard consumption can be made by adding specific inhibitors
metabolic rate of the mitochondrial respiratory chain and measuring the
Body Mass, % Body Oxygen Use, %
residual oxygen consumption of the tissue. Nonmitochon-
drial respiration accounted for -2 1% in isolated resting
Organ Human Rat Human Rat rat liver cells (20, 41), 20% in perfused liver (187), 14%in
Liver 2 5 17 20 perfused resting rat skeletal muscle (63, 176), and 20% in
Gastrointestinal tract 2 5 10 5 rabbit reticulocytes (192), but only 2% in isolated thymo-
Kidney 0.5 0.9 6 7 cytes (134) and 3% of the estimated in vivo oxygen con-
Lung 0.9 0.6 4 1
Heart 0.4 0.5 11 3 sumption of rat heart (48). If we assume that nonmito-
Brain 2 1.5 20 3 chondrial oxygen consumption is 20, 14, and 3% of in vivo
Skeletal muscle 42 42 20 30 oxygen consumption of liver, skeletal muscle, and heart,
Total 49.8 55.5 88 69
respectively, and multiply these fractions by the respec-
Organ masses and oxygen consumption rates are mean values tive contributions of each organ to the total oxygen con-
from data on humans (1, 3, 72, 116, 135, 183) and rats (10, 71, 116, 117, sumption of rat and human (listed in Table l), then we
173, 181, 199). All data are for adult individuals except for data of Field can estimate that nonmitochondrial oxygen consumption
et al. (71), which are for 150-g animals. Data for oxygen consumption
rates are taken from in vivo tissue respiration measurements except within these organs alone contributes 8.3 and 6.5%of total
some data for rat skeletal muscle which are from tissue slices (71) and oxygen consumption in rats and humans, respectively.
perfused hindlimb preparations (10, 181, 199). Data for skeletal muscle Because these organs contribute -50% of total oxygen
tissue slice respiration are corrected for any observed decline in slice
respiration rate with time (71). consumption, and other organs are likely to have a smaller
but still significant proportion of nonmitochondrial oxy-
gen consumption, then the total contribution of nonmito-
dria, and the Na’-Kf-ATPase colocalize to dendrites and chondrial oxygen consumption to whole body oxygen
presynaptic terminals (see Refs. 68, 193, 220). For exam- consumption is likely to be -10%.
ple, in primate visual cortex, 62% of mitochondria are A small proportion of mitochondrial oxygen con-
found within dendrites, 21% in presynaptic terminals, with sumption is not coupled to proton pumping, due to leak-
relatively little in axons (12%) or glia (2%) (220). In skeletal age of electrons from the electron transport chain to re-
muscle, oxygen consumption and mitochondrial content duce oxygen to superoxide and hydrogen peroxide. This
are higher in red muscle than white muscle and decrease proportion has been estimated to be maximally l-2% of
in the order type I > type IIa > type IIb muscle fibers. In the oxygen consumption rate of isolated mitochondria by
kidney, the density of mitochondrial volume (mitochon- measuring the rate of hydrogen peroxide production
drial volume per unit volume of cytoplasm) is similar to (17, 205).
that of mitochondrial membrane density and basolateral
membrane surface density and decreases from 33 to 22% 2. lWi tochondrial pro ton leak
along the proximal tubule. Along the distal portion of the
nephron, it decreases from 44% in the medullary thick As stated in the introduction to section II, not all
ascending limb to 31% in the distal convoluted tubule. mitochondrial oxygen consumption in the standard state
In the thin limbs, the density is only 6-8%, whereas the is coupled to ATP synthesis. In brown adipose tissue, a
collecting tubule has a mitochondrial density of -20 and specific protein (thermogenin) enhances the proton con-
10%in the cortical and medullary segments, respectively ductance of the mitochondrial inner membrane, uncou-
(161). The liver is fairly homogeneous: 80% of liver cells pling mitochondrial oxygen consumption from ATP syn-
have the same or similar mitochondrial content, but the thesis (155). Brown adipose tissue (BAT) contributes to
density of inner mitochondrial membrane is slightly lower SMR in neonatal mammals and small mammals adapted
in pericentral regions of the lobules (140). to cold environments (105, 155). Foster and Frydman (74)
found that over 60% of the excess heat produced in re-
sponse to cold in rats is generated in BAT. Brown adipose
B. Contribution of Non-ATP-Consuming Reactions
tissue thermogenesis is also stimulated by overfeeding
in the Major Oxygen-Consuming Tissues to
(105, 155). However, BAT does not make a significant
Standard Metabolic Rate
contribution to SMR in adult rats and humans (73).
Thermogenin is restricted to BAT mitochondria; how-
1. Nonmitochondrial oxygen consumption
ever, mitochondria isolated from other organs have a sig-
Not all oxygen consumption by tissues is mitochon- nificant passive leak for protons (proton leak), which is
drial or coupled to proton pumping and, therefore, in- apparently not mediated by a specific protein (reviewed
volved in ATP production. A large number of oxidases in Refs. 20,35). Proton leak is not an artifact of mitochon-
function throughout metabolism, e.g., in peroxisomal fatty drial isolation, since it has been demonstrated in mito-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
736 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

chondria within isolated hepatocytes (22, 41, 97, 156) and 58%) (41, 97, 156, 165, 169). Similarly, the in vitro oxygen
thymocy-tes (44). consumption rates of perfused muscle preparations are
With the assumption that mitochondrial oxygen con- within the same range as estimates of the oxygen con-
sumption and proton pumping are obligatorily coupled sumption rate of skeletal muscle in vivo. Of course, this
under physiological conditions (see section I&!?), oxygen argument does not preclude overestimates in proton leak
consumption in the absence of phosphorylation may be measurements in vitro, since stimulations of ATP con-
used as an indirect measure of proton leak activity (38, sumption would also cause decreased proton leak flux
154). The contribution of proton leak to the resting oxygen and therefore may not significantly alter the overall oxy-
consumption of a particular tissue may be thus estimated gen consumption rate while altering the balance of ATP
by completely inhibiting mitochondrial ATP production. consumption and proton leak. However, it is one indica-
This inhibition will result in a drop in the tissue oxygen tion that perfused muscle and liver cells may be reason-
consumption rate and an increase in the mitochondrial able models for the whole tissue in vivo, and thus the
membrane potential. Because of the strong dependence experiments outlined above show that proton leak may
of the mitochondrial proton leak on membrane potential, be an important contributor to SMR, at least in rats.
the mitochondrial potential must then be returned to its With a knowledge of the contribution to rat SMR of
resting value by sequentially inhibiting the mitochondrial skeletal muscle (estimates range from 13 to 42%; mean
respiratory chain. Once the mitochondrial potential has 30%) and liver (lo-20%; mean 15%), the contribution of
been returned to its resting value, the remaining oxygen proton leak in liver and skeletal muscle to rat SMR can
consumption is a measure of the magnitude of the proton be estimated to be, on average, 20% (30 X 0.52 + 15 x
leak flux (plus any nonmitochondrial oxygen consump- 0.26) [range 10 (13 X 0.52 + 10 x 0.26) to 27 (42 x 0.52 +
tion) that had been occurring under resting conditions. 20 x 0.26)%]. If the same proportions of liver and skeletal
The nonmitochondrial oxygen consumption rate is deter- muscle oxygen consumption were uncoupled in humans,
mined by completely inhibiting mitochondrial oxygen then the proton leak in these tissues would make a smaller
consumption, and hence, the proton leak flux can be cal- contribution to SMR (- 15%) due to the smaller contribu-
culated. tion of skeletal muscle in humans. The contribution of
Experiments (of the type described above) to esti- proton leak in other tissues, apart from heart above, has
mate the proton leak rate have been carried out using not been estimated, although proton leak is present in
isolated rat liver cells and perfused rat skeletal muscle mitochondria isolated from brain and kidney and has simi-
and indicate that proton leak accounts for 26% of oxygen lar activity to that found in isolated liver mitochondria
consumption in resting rat liver cells (22, 41, 97, 156; re- (178). It might be argued that, because these organs have
viewed in Ref. 20) and 52% in resting rat skeletal muscle a higher demand for ATP relative to that in skeletal muscle
(20, 176). In the heart, Challoner (48) found that 84% of in the standard state, there may be a relatively smaller
the estimated in vivo oxygen consumption was inhibited contribution of proton leak in these tissues, since the
by arresting the heart; of the remaining 16%, 94% was proton leak and ATP synthesis compete for the same inter-
insensitive to oligomycin, but only 20% was insensitive to mediate. Thus a reasonable estimate of the overall contri-
cyanide (an inhibitor of the respiratory chain). Thus a bution of proton leak to SMR is -20%.
maximum of lo- 13% of the in vivo oxygen consumption
may be coupled to the proton leak in heart. This will 3. P-to-O Ratio and H’ stoichiometries
probably overestimate the actual contribution of proton The P-to-O ratio (P/O) is the ratio of ATP synthesized
leak (as explained above), but the correction of this over- to oxygen consumed (expressed in moles of 0 rather than
estimate in liver and skeletal was not greater than 50%, O,), but we may distinguish the mechanistic ratio from
so we may estimate the contribution of proton leak in the effective ratio. The mechanistic P/O is the maximal
heart to be -5-10% of SMR. Note that the experiments P/O in the absence of proton leak, other uncoupling reac-
outlined in References 41 and 176 may give an upper limit tions, and oxygen consumption other than at cytochrome
to proton leak activity in the intact animal, since the skele- oxidase and is equal to the product of the Hf/O and ATP/
tal muscle and liver may have a greater ATP demand in Hf ratios. The effective P/O of a tissue is the ratio of ATP
the standard state in vivo than in vitro, for example, due to synthesized (by mitochondria) to oxygen consumed by
muscular contraction and additional liver work functions. the whole tissue. The effective P/O is related to the mecha-
However, the oxygen consumption rate of isolated hepa- nistic P/O of the mitochondrial ATP synthesizing reactions
tocytes may be scaled to allow comparison with the in by multiplying the mechanistic ratio by the fraction of the
vivo rates in whole liver (14). The oxygen consumption total tissue oxygen consumption rate that is used to drive
rate of whole liver in vivo is 2.84 pmol O2 min-’ g liver-’
l l mitochondrial ATP synthesis. The effective P/O differs
(116). The oxygen consumption rates of resting hepato- from the mechanistic P/O mainly because of nonmito-
cytes isolated from fed rats incubated at 37°C in Krebs- chondrial oxygen consumption and mitochondrial proton
Henseleit bicarbonate buffer is between 33 and 79% (mean leak.

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 737

The mechanistic P/O was previously thought to be 1 systems is used to drive ATP synthesis. Thus it would
per site of the mitochondrial respiratory chain, i.e., 3 for appear that the data presented here for the contribution
mitochondria respiring on NADH and 2 for mitochondria of proton leak to resting liver cell and skeletal muscle
respiring on succinate. These estimates were based on oxygen consumption contradict the current perceptions
measurements of the effective P/O in mitochondria but of the tightness of coupling between oxygen consumption
overestimated the effect of the proton leak on these mea- and ATP synthesis in cells and organs. If the contribution
surements. More recent measurements quantitatively ac- of oxidative phosphorylation to resting oxygen consump-
counting for the proton leak estimate the mechanistic tion is 53%in liver cells (41,97, 156) and 34%in muscle (20,
P/O to be 2.5 for mitochondria respiring on NADH and 176), then the effective P/O in these systems (assuming a
1.5 respiring on succinate (19, 106). This fits well with the mechanistic P/O of 2.5) would be 1.3 (0.53 x 2.5) for liver
present consensus estimates of the Hf/O ratios of the and 0.9 (0.34 X 2.5) for muscle. However, it has been
respiratory chain (10 for NADH oxidation and 6 for succi- pointed out (29, 27, 127) that 31P-NMR saturation transfer
nate oxidation) and the H+/ATP ratio of ATP synthesis measurements of Pi+ATP flux may significantly overesti-
and transport (3 H+ for ATP synthesis and 1 Hf equivalent mate the net Pi+ATP flux catalyzed by the mitochondrial
for transport of ADP and Pi; see Refs. 16, 106). With the ATP synthase because 1) the unidirectional forward rate
use of these figures, the mechanistic P/O has been esti- of the ATP synthase may be much greater than the net
mated (16) to be 2.42 when respiring on pyruvate (with 1 rate because the backward rate is significant, and 2) a
FADH and 4 NADH oxidized and 1 GTP synthesized per proportion of the measured Pi+ATP flux is catalyzed by
pyruvate), 2.58 with glucose as substrate (with an extra glyceraldehyde-3-phosphate dehydrogenase and phospho-
2 NADH and 2 ATP produced per glucose by glycolysis), glycerate kinase in yeast (28) and perfused heart (127).
and 2.26 with palmitate as substrate (with 2 ATP equiva- Inhibition of the glyceraldehyde-3-phosphate dehydroge-
lents required for activation to palmitoyl-CoA, 7 NADH, nase and phosphoglycerate kinase exchange reaction in
and 7 FADH produced by P-oxidation, 8 x 3 NADH, 8 beating heart lead to a drop in the estimated P/O from 6.4
FADH, and 8 GTP from the tricarboxylic acid cycle). to 2.2 (127). Overexpression of yeast adenine nucleotide
There has been some doubt as to whether the mito- translocase, the kinetic properties of which are thought
chondrial oxygen consumption and proton pumping have to influence the backward (ATP+PJ rate of the mitochon-
a fixed coupling ratio (i.e., the H+/O ratio might vary or drial ATP synthase, caused the apparent mitochondrial
“slip”) (reviewed in Refs. 35, 149). If the mitochondrial P/O to increase from 0.9 to 4.2 (26). Thus it seems likely
proton pumps did slip, this would result in large changes that previous estimates of the effective P/O in resting cells
in thermogenesis, free energy dissipation, and efficiency and tissues by “lP-NMR saturation transfer were erron-
of the mitochondrial proton pumps themselves. However, eously high.
more recent evidence indicates that, under approximately
physiological conditions, the ratio is fixed (20, 21, 32, 92,
125, 221). C. Contribution of ATP-Consuming Reactions in
We can estimate the effective P/O in vivo by combin- the Major Oxygen-Consuming Tissues to
ing estimates of the mechanistic P/O (which we will as- Standard Metabolic Rate
sume to be -2.5) with estimates of the fraction of tissue
oxygen consumption 1) used by nonmitochondrial oxygen We now move on to consider the contribution of
consumption (O-20% depending on tissue) and 2) coupled the different ATP-consuming processes within the cell
to proton leak (O-50% depending on tissue). If the com- to SMR. Estimates of the relative ATP use by a process
bined nonmitochondrial respiration rate of heart, skeletal are usually made either 1) by measuring the rate of the
muscle, and liver accounts for -10% of SMR (see sect. process and estimating the oxygen consumption required
IIBI) and the proton leak in these organs accounts for to account for that rate, or 2) by specifically inhibiting
-20% of SMR (see sect. 11B2),then this allows us to esti- the process and measuring the resultant fractional
mate the effective whole body P/O as 1.8 (0.7 X 2.5). If change in oxygen consumption. Estimates based on the
we exclude the contribution of nonmitochondrial oxygen first method need to assume a value of the P/O that has
consumption, then we can estimate the effective mito- in the past been overestimated (see sect. II.@. In the
chondrial P/O as 2.0 (0.8 X 2.5). absence of significant glycolytic ATP production (see
Considerable interest has been shown in the effective sect. IIG), 1) the ATP synthesized by the tissue = (the
P/O of intact tissues (e.g., Refs. 29, 46, 76, 127). The effec- oxygen consumption of the tissue X the effective P/O of
tive P/O of yeast (29, 46), perfused kidney (76), and heart the tissue), 2) the fraction of tissue ATP production used
(127) estimated from 31P-nuclear magnetic resonance by the particular ATP user = (the ATP use by a particular
(NMR) saturation transfer measurements of Pi+ATP flux ATP user/the oxygen consumption of the tissue X the
were close to 3, 2.45, and 2.36, respectively. These data effective P/O), and 3) the fraction of tissue oxygen con-
imply that virtually all the oxygen consumption in these sumption coupled to the particular ATP user = (the ATP

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
738 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

use by a particular ATP user/the oxygen consumption of B-36% for contracting heart muscle, and 2% in resting
the tissue X the mechanistic P/O). liver. A significant contribution of this ATPase in brain
Estimates of the contribution of the different ATP- was also proposed (53), although a specific figure was not
consuming processes to metabolic rate based on the use suggested. Using the estimates for resting skeletal muscle
of inhibitors can be unreliable because 1) the inhibitors and liver and contracting heart, and the organ oxygen
are rarely specific, 2) inhibition of functions such as ion consumption rates of Table 1, we can estimate that cal-
transport may affect energy metabolism at multiple sites, cium cycling in heart, liver, and muscle contribute 4-6%
and 3) inhibition of a major ATP user may raise the ATP/ of SMR in humans and between 3 and 4% of SMR in rats.
ADP and thus stimulate other ATP consumers. The extent This figure may rise dramatically if there is any skeletal
to which the latter is a problem depends on the relative muscle contraction.
sensitivity of the ATP-consuming process and ATP pro-
duction to the ATP/ADP (see Refs. 34,43). If the sensitivity 3. Muscle contraction
of the ATP consumers is insignificant relative to mito-
The actinomyosin ATPase probably accounts for a
chondrial ATP production, as has been shown in conca-
significant proportion of the ATP turnover even in the
navalin A-stimulated thymocytes (43), then complete and
standard metabolic state, since heart work, lung work,
specific inhibition of an ATP consumer will not lead to a
vascular, gastrointestinal, and other smooth muscle tone,
significant drop in ATP/ADP and will not underestimate
and skeletal muscle tone are still required in this state.
the contribution of that ATP-consuming process to tissue
There is no good estimate of the contribution of smooth
oxygen consumption.
muscle at present. The maintenance of muscle tone via
the actinomyosin ATPase is often considered to be an
1. Na’-K”-ATPase
important contributor to SMR. However, Webster (213)
The Na’-K+-ATPase is an important ATP consumer, summarized results indicating that the energy costs of
although its importance may have been overestimated standing and of changing position were only -1% of the
previously, due to the secondary effects on cellular oxy- total energy expenditure of a confined, fed steer. Hence,
gen consumption of ouabain, an inhibitor of the Naf-K’- the contribution of muscle tone maintenance to SMR may
ATPase widely used in studies of the ion pump (reviewed be insignificant. However, the electrical activity of resting
in Refs. 53, 124). The contribution of this pump has been skeletal muscle in anesthetized rats is high (190). Ac-
estimated in a large range of tissues either by measuring cording to Rohracher (175), the roughly lo-fold greater
ion fluxes or by determining the effect of ouabain on oxy- metabolic rate in homeotherms compared with poikilo-
gen consumption. The estimated coupling to oxygen con- therms of similar body weight is due to the existence of
sumption of the Na’-K’-ATPase is - 14% in adult rats (cal- “muscular microtremor.” However, this theory has not
culated from data in Ref. 53 and data for rat organ meta- been tested. It has been estimated that the isolated, beat-
bolic rates shown in Table 1) and -20% in humans ing but unloaded heart (doing no external work) con-
(calculated from data in Ref. 53 and data for human organ sumes 50% of the oxygen consumption of the rat heart in
metabolic rates shown in Table 1). The contribution is vivo in the standard state, and the arrested heart con-
particularly large in brain (4, 68) and kidney (98, 197), sumes -15% (48, 80, ZOO). Thus -85% of the in vivo oxy-
where the pump has been estimated to be coupled to 50- gen consumption of heart is due to activation and contrac-
60% of oxygen consumption. tion. It is thought that 50-60% of this ATP requirement is
for the actinomyosin ATPase (53, ZOO). Therefore, it seems
2. Ca”+-ATPase likely that between 43% (0.85 x 0.5) and 51% (0.85 X 0.6)
of the oxygen consumption of rat heart in vivo is due to
As for the Na+-K+-ATPase, the contribution of the
the actinomyosin ATPase, and thus this ATPase accounts
calcium ATPase to the metabolic rate of the major oxygen-
for a minimum of between 1.3% (3 X 0.43) and 1.5% (3 X
consuming tissues of many mammals has been exten-
0.51) of the whole body SMR in rats and between 5% (11 x
sively reviewed by Clausen et al. (53). The contribution
0.43) and 6% (11 x 0.51) of that in humans (calculated
of the Ca’+-ATPase is usually assessed by measurement
using the organ metabolic rates given in Table 1).
of transmembrane ion fluxes and multiplying these fluxes
by the ATP:Ca2+ stoichiometry of the ATPase. The contri-
4. Protein turnover
bution of the ATPase to total tissue ATP consumption is
then calculated by assuming a P/O of 3 for the tissue (53). The contribution of protein synthesis and break-
If the data of Clausen et al. (53) are reevaluated using a down to SMR has been estimated either by 1) measuring
mechanistic P/O of 2.5, we can calculate that the Ca2’- rates of protein synthesis or 2) measuring the inhibition
ATPase accounts for the following fractions of tissue met- of oxygen consumption when protein synthesis is inhib-
abolic rate: 6% in resting skeletal muscle, 24-58% in con- ited. Waterlow (210) and Waterlow and Millward (212)
tracting skeletal muscle, <l% for resting heart muscle, have reviewed manv measurements of whole bodv pro-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 739

TABLE 2. Contribution of major oxygen-consuming processes to oxygen consumption rate of rat tissues in
standard state

Tissue Protein Synthesis Na+-K+-ATPase Ca’+-ATPase Other Major Contributors

Liver 24 5-10 5 Gluconeogenesis (15-40%) and urea synthesis (12%)


substrate cycling (26%), oligonucleotide synthesis
(5%), proton leak (26%), nonmitochondrial (20%)
Gastrointestinal tract 74 60
Kidney 6 40-70 Gluconeogenesis (5%)
Heart 3 l-5 15-30 Actinomyosin ATPase (40-50%), proton leak (15% max)
Brain 5 50-60 Significant
Skeletal muscle 17 5-10 5 Proton leak (50%), nonmitochondrial (14%)

Data for protein synthesis are calculated using data for whole body protein turnover and fractional contribution of each tissue to this figure
(210) and assuming that the average molecular weight of a peptide bond is 110, ATP peptide bond stoichiometry is 4, and ATP/O is 2.5. Data for
Na’-K+-ATPase and Ca’+-ATPase are from Clausen et al. (53). Data for substrate cycling are from Rabkin and Blum (169). Data for glucose and
urea synthesis are from Kusaka and Ui (132), Berry et al. (14), and Haussinger et al. (100). Estimates of contribution of actinomyosin ATPase
activity to ATP turnover are from data of Challoner (48), Suga (200), and Clausen et al. (53). Note that, where it was necessary to assume a value
for P/O to calculate data presented here, a figure of 2.5 has been used. Data for rat tissue oxygen consumption rates in the resting state are taken
from Table 1. Note also that because data used to compile this table are taken from different sources; the sum of energy-requiring processes in
some organs may exceed 100%.

tein turnover using labeled amino acids and estimate presence of the inhibitors despite the fact that proteolysis
that protein synthesis contributes 15-18% of SMR. This (measured as radiolabeled lysine release from cell pro-
estimate is calculated essentially by multiplying the pro- tein) was abolished by emetine in <30 min. Thus the
tein synthesis rate by the ATP requirement (4 ATP equiv- overall contribution of protein degradation is impossible
alents per peptide bond), dividing by the P/O (assumed to estimate at the moment, although the ATP requirement
to be 3), and then comparing it with the SMR. This calcu- of ubiquitin-dependent proteolysis, -4 ATP/protein (102),
lation will give a small underestimate of the contribution would not appear to predict a significant role for ATP-
of the SMR (using a mechanistic P/O of 2.5 the contribu- dependent proteolysis as a contributor to SMR.
tion is - 1822%).
Estimates of the contribution of protein synthesis us- 5. Substrate cycles
ing specific inhibitors and determining the change in oxy-
Newsholme (152) estimated that if the phosphofruc-
gen consumption rate in various in vitro tissue prepara-
tose-fructose bisphosphate cycle was fully active, it could
tions range from 2 to 30% in different organs, with an
account for -50% of an adult human’s daily expenditure.
overall contribution to SMR of 12-25% (reviewed in Ref.
He concluded that stimulation of many such cycles could
124). Estimates of the contribution in different organs are
result in the conversion of significant amounts of chem-
given in Table 2. Reeds and Harris (172) found that the
ical energy into heat. Rabkin and Blum (169) estimated
rate of protein synthesis has the same relationship with
the activity of five substrate cycles (glucose/glucose-6-
the body mass of an animal as its SMR; thus protein syn-
phosphate, glycogen/glucose-6-phosphate, fructose-6-phos-
thesis makes the same percent contribution to SMR in a
phatelfrmctose 1,6-bisphosphate, phosphoenolpyruvatel
variety of animals regardless of body size.
pyruvate/oxaloacetate, and acetate/acetyl CoA) using
The net energy costs of protein degradation are not
fluxes calculated by a computer model that gave the best
clear, although various authors have claimed that this pro-
fit to data from radioisotopic incorporation experiments.
cess is a significant energy sink (44, 45, 148, 182, 192).
The results indicated that the contribution of these five
The energy cost of proteolysis was assessed as the differ-
substrate cycles to resting metabolic rate in isolated hepa-
ence in the amount of cellular oxygen consumption inhib-
tocytes was -26% (although the contribution to ATP turn-
itable by emetine (an inhibitor of ATP-dependent proteol-
over may be higher due to low effective P/O in hepato-
ysis and protein synthesis) and cycloheximide (an inhibi-
cytes). However, the potential for substrate cycling is
tor of protein synthesis only). With the use of this method,
probably larger in liver than most other tissues. Modeling
estimates of the coupling of ATP-dependent proteolysis
of energy metabolism in young pigs suggests that sub-
to cell oxygen consumption have been 11- 15% in resting
strate cycling in glycolysis, triglyceride turnover, and the
rabbit reticulocytes (191, 192), 58% in Ehrlich mouse
Cori cycle account for 7.5% of SMR (171).
ascites tumor cells (148, 182), 7% in resting lymphocytes
(45), and <l% in resting thymocytes (41). However, these
6. RNA and DNA turnover
data are difficult to interpret, since the effect of emetine
only becomes significantly different from that of cyclohex- Several attempts have been made to quantify the con-
imide when the cells are incubated for over an hour in the tribution of oligonucleotide turnover to resting cell oxy-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
740 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

gen consumption using inhibitors of RNA and DNA syn- this estimate of the contribution to SMR cannot be used
thesis such as actinomycin and amanitin. With the use with any confidence.
of the inhibitor method, the coupling of oligonucleotide The contribution of DNA synthesis to SMR is thought
synthesis to cell oxygen consumption has been estimated to be much smaller than RNA, since the turnover is much
as 12% in resting lymphocytes (45) and Ehrlich mouse lower (147).
ascites tumor cells (182) but was insignificant in resting
thymocytes (44). 7. Enzyme phosphorylation and signal transduction
Another way of estimating the energy cost of oligonu-
Goldbeter and Koshland (84) have made rough esti-
cleotide turnover is to directly measure the rate of synthe-
mates of the cellular energy expenditure on protein phos-
sis by incubating cells or tissues with radiolabeled nucleo-
phorylation. They suggest that glycogen phosphorylase
tides. Although most RNA in the cell is rRNA and tRNA,
kinase could use a substantial proportion of cellular ATP
the turnover rate of mRNA (and its precursor hnRNA) is
turnover (> lo%), while adenosine 3’,5’-cyclic monophos-
much higher, so that the rate of RNA turnover in the cell
phate (CAMP)-dependent protein kinase and protein ki-
is mostly due to mRNA (or hnRNA). Different types of
nase C could make a significant contribution. They further
mRNA also turnover at widely different rates, so that esti-
suggest that since one in six mammalian proteins is re-
mating the overall rate of RNA turnover is not easy. How-
versibly phoshorylated, the total cost of protein phosphor-
ever, with the use of such methods, the rate of synthesis of
ylations may be substantial. However, these conclusions
RNA was estimated as 0.105 mg ml cells-’ h-l in rabbit
l l

were based on measured maximal rates rather than actual


reticulocytes (158) and as 0.80 mg 100 g-’ h-l in per- l l

rates in cells, and there is no good estimate of the contri-


fused rat liver (136). With the use of an ATP:RNA nucleo-
bution of protein phosphorylation to ATP turnover.
tide stoichiometry of 2 and with the assumption that the
Phospholipids, particularly those involved in signal
average molecular weight of a nucleotide in an RNA mole-
transduction, undergo rapid energy-dependent synthesis
cule is 320, the total ATP consumed by RNA synthesis in
and breakdown (12). Goldbeter and Koshland (84) state
these systems can be calculated as 0.65 [(O. 105 x 10B3)/
that phosphatidylinositol turnover in thrombin-stimulated
320 x Z] pmol ATP ml reticulocytes-’
l h-’ and 5 [(O.S x l

platelets may transiently account for a significant propor-


10-3)/320 x Z] pmol ATP 100 g liver-’l h-‘. The ATP con- l

tion of the ATP consumption in these cells. However,


sumption rate of reticulocytes and whole liver can be
Verhoeven et al. (208) concluded that the markedly en-
calculated as 95 pmol ATP ml reticulocytes-’ l h-l (from l

hanced phospholipid metabolism during secretion in


Refs. 191, 192 and using an effective P/O of 1.9) and 44
platelets involved only a small fraction of the total energy
pmol ATP 100 g liver-’
l h-l (from Ref. 116 and using
l

usage of the cell. Similarly, it was calculated (by radioiso-


an effective P/O of 1.3). Thus the contribution of RNA
topic measurement of inositol phospholipid turnover and
turnover to tissue ATP consumption in reticulocytes and
knowing the ATP requirement of inositol phospholipid
liver can be estimated as -1% (0.65/95 x 100) and 10%
synthesis) that a loo-fold stimulation in inositol phospho-
(5/44 x loo), respectively. Because the oxygen consump-
lipid turnover (compared with the basal turnover rate) in
tion driving ATP synthesis represents 50% (20) and a mini-
contracting canine trachealis smooth muscle would only
mum of -70% (191) of the total oxygen consumption rate
cause an increase in ATP consumption equivalent to 1%
of isolated hepatocytes and reticulocytes, respectively,
of the basal ATP requirement (31). Thus the contribution
then the contribution of RNA synthesis to the overall oxy-
of phospholipid turnover to SMR is unclear, but probably
gen consumption rate may be estimated to be 1% in thymo-
small.
cytes and -5% in liver.
The ATP requirement for CAMP turnover in platelets
Whole body RNA turnover has been estimated in
has been estimated to be 3-6% of the total ATP turnover
adult humans from the excretion rate of rare bases or their
(83, 84), and maximal stimulation of adenylate cyclase
metabolites in urine, as 0.5 pmol kg body wt-’ day-’ for
can completely deplete cell ATP in adipocytes (52). G
l l

mRNA, 0.5 pmol kg-’ day-’ for tRNA, and 0.03 pmol
proteins can hydrolyze GTP at a rapid rate, but again,
l l l

kg-’ day-’ for rRNA (186). If the average size of primary


there is no good estimate of this contribution to ATP
l

mRNA transcripts being turned over was 20,000 nucleo-


turnover.
tides, then the mRNA turnover rate would correspond to
10 mmol nucleotides kg-’ day-‘, equivalent to 20 mmol
8. Gluconeogenesis
l l

ATP kg-’ day-‘, equivalent to 1.4 mol ATP. day-’ hu-


l l l

man, equivalent to 0.56 mol0 day-’ human-’ (assum-


l l Gluconeogenesis requires ATP and occurs at a high
ing a P/O of 2.5), equivalent to 1.8% of basal metabolic rate under standard conditions. The flux through this path-
rate (with a basal oxygen consumption of 32 mol O/day). way is generally measured in vivo by infusing small quanti-
However, the average size of primary transcript being ties of a radiolabeled substrate of de novo glucose synthe-
turned over is very uncertain, and the accuracy of the sis such as alanine and observing the rate of transfer of
estimated mRNA turnover rate is also uncertain (186), so the radiolabel to glucose. This method may underestimate

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 741

the total rate of gluconeogenesis, since only glucose aris- fluorosuccinate, Gregory and Berry (86) found that a 72%
ing from radiolabeled alanine itself (or other gluconeoge- inhibition of glucose production resulted in a 21% reduc-
netic substrates such as lactate which might also become tion in oxygen consumption rate, and thus we might guess
labeled as a result of the alanine infusion) will be mea- that a 100%inhibition would produce a 30% reduction in
sured as gluconeogenesis (170). Furthermore, exchange oxygen consumption rate, i.e., that gluconeogenesis ac-
may occur between radiolabeled carbon and carbon from counts for -30% of liver metabolic rate under conditions
sources (e.g., intramitochondrial oxaloacetate), which designed to maximize de novo glucose production. How-
make no net contribution to gluconeogenesis (170). In ever, it is not clear what other effects perfluorosuccinate
overnight-fasted humans (143), the rate of gluconeogene- has on hepatocyte metabolism, so the results of this study
sis was estimated, using radiolabeled alanine, to be -0.72 are hard to interpret. Thus it appears that a variety of
mol glucose/day. This is equivalent to 4.3 mol ATP/day techniques and systems for measuring gluconeogenesis
(assuming 6 ATP/glucose) or 1.7 mol O/day assuming a give approximately the same answer: that de novo glucose
P/O of 2.5. The average adult human basal oxygen con- synthesis accounts for between 5 and 8% of SMR in hu-
sumption is 32 mol O/day (116, 184), so this represents mans and between 3 and 8% in rats.
5% of human SMR or -28% of human liver SMR assuming
liver accounts for 17%of human SMR (Table 1) and that 9. Urea synthesis
-10% of gluconeogenesis occurs in the kidney cortex. In Urea synthesis requires ATP and continues in the
young (- 170 g) rats starved for 20 h (132), the rate of standard state. Total urea synthesis is 25 g/day equivalent
gluconeogenesis was estimated to be 14 mmol/day by to 0.42 mol/day in an average human, requiring - 1.7 mol
measuring the rate of conversion of radiolabeled lactate to ATP/day (assuming 4 ATP/urea) and thus 0.67 mol O/day,
glucose. This is equivalent to -34 mmol O/day assuming 6 and thus 2% of the total oxygen consumption in humans.
ATP/glucose and a P/O of 2.5. This is roughly equivalent Almost all this urea synthesis occurs in liver, so that this
to 7% of rat SMR assuming that a 170-g rat consumes 490 represents 12% of the SMR of human liver. In perfused
mmol O/day (calculated from data in Ref. 71) or 32% of liver from fed rats or hepatocytes prepared from these
liver SMR if liver accounts for 20% of rat SMR (Table 1) livers, urea production can account for a significant pro-
and 90% of gluconeogenesis occurs in liver. portion (up to 30%) of liver SMR (14, 86, 100). Total nitro-
Data from isolated hepatocytes may enable us to gen excretion in 48-h starved rats was measured (55) to
crudely determine the significance of any underestimate be 73 mg N day-’ 100 g body wt-‘, with 75% of urinary
l l

of gluconeogenesis in the above data. Because glycogen N in urea this is equivalent to 3.9 mmol urea day-’ 100 l l

stores are completely depleted in 18 to 24-h starved rats g body wt’ (73 x 0.75/14), which would require 15.6
(14), all glucose production in hepatocytes isolated from mmol ATP day-’ 100 g-l (with 4 ATP required per urea
l l

these rats will be from gluconeogenesis. In hepatocytes synthesized), and thus 3.12 mmol O2 day-’ 100 g-l (with
l l

isolated from 18 to 24-h fasted rats, the rate of glucose a mechanistic P/O of 2.5). The same rats had an oxygen
production under conditions designed to maximize the consumption of 4.43 g O2 day-’ 100 g-’ (55), equivalent
l l

rate of gluconeogenesis has been estimated to be between to 138 mmol OZ day-’ 100 g-‘, and thus the oxygen re-
l l

0.5 and 1.6 pmol g wet wt cells-’ mir-? (14,86). Knowing


l l
quirement for urea synthesis represents 2.3% of rat SMR
the weight of cells in whole liver (14), we can estimate or 12%of rat liver SMR. Thus it appears that urea synthesis
that the rate of gluconeogenesis in rat liver is - lo-28 may also make a small but significant contribution to SMR
mm01 liver-l day-‘, which is equivalent to 24-67 mmol
l l
in rats and humans (-2%).
O/day driving gluconeogenesis in the whole rat assuming 10. Other ATP-consuming processes contributing to
6 ATP/glucose and a P/O of 2.5. An adult rat (-300 g) standard metabolic rate
consumes -840 mmol O/day (116), and hence, gluconeo-
genesis in the liver can be estimated to account for be- Exocytosis and endocytosis allow the movement of
tween 3 and 8% of rat SMR and between 15 and 40% large molecules into and across cell membranes and the
of rat liver SMR. However, the above calculation of the modulation of cell responses to hormones by altering re-
contribution of gluconeogenesis to SMR depends on the ceptor availability. Energy is required for fibroblast endo-
proportion of gluconeogenesis that is produced from glyc- cytosis (160), exocytosis in amoeba (157), and for synaptic
erol (stoichiometry = 2 ATP/glucose) compared with the vesicle cycling in nerve terminals. However, the energy
other major gluconeogenetic substrates alanine, gluta- cost of these processes would be minor if one or a few
mine, and lactate (6 ATP/glucose). Glycerol carbon has ATP molecules were expended per vesicle or large mole-
been estimated to account for -5% of the total carbon cule moved (147).
used for gluconeogenesis in 16- to 18-h fasted adult dogs
D. Service Functions
(94). If this is also true for rats, then the above calculation
for hepatocyte gluconeogenesis would overestimate the Another way of looking at the contribution of various
true gluconeogenetic flux by -3%. Using the inhibitor per- processes to whole body oxygen consumption in the stan-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
742 DAVID F.S. ROLFE AND GUY C. BROWN Volume 77

dard state is to divide metabolism into service and nonser- E. Relative ATP Use by Cellular ATPases
vice functions (7). Service functions are supracellular
functions that serve the whole body, such as kidney excre- Section IIC estimated the extent to which the major
tion, nervous function, and pressure-volume work done ATP-consuming processes were coupled to cellular oxy-
by the heart and lungs, whereas nonservice functions are gen consumption, which closely approximates their con-
for maintenance of the cell itself (e.g., maintenance of ion tribution to SMR. Clearly, if all the oxygen consumption
gradients). Service functions have been estimated (see of a tissue was used to produce ATP (i.e., the effective
references in Refs. 7, 71) to be between 25 and 50% of the P/O = 25), then the contribution of these processes to
SMR of an animal. However, in practice, it is not clear ATP consumption would be the same as their coupling to
to what extent the service and nonservice functions are oxygen consumption or SMR. However, as mentioned in
distinguishable (e.g., kidney and neuronal work is mainly section II&?, the effective P/O may be less than maximal
ion transport), and hence, an assessment of the relative in liver and skeletal muscle and perhaps in other tissues
contribution of the service and nonservice functions to also. We can obtain an estimate of the contribution of a
SMR is difficult. particular ATP consumer to whole body ATP turnover by
One way to estimate the contribution is to compare multiplying the coupling of the ATP consumer to oxygen
the oxygen consumption of tissues incubated in a minimal consumption by the maximum P/O (2.5), then dividing by
medium in vitro (where service functions are likely to be the whole body oxygen consumption rate multiplied by
minimal) to the oxygen consumption measured in vivo in the effective whole body P/O (1.8; see sect. rB3). In this
the standard state. Measurements of summed tissue slice way we can estimate the contribution to whole body ATP
oxygen consumption in rats (71), mice (142), and dogs consumption of protein synthesis (25-30%), Na’-K+-
(142) indicate that an average of -80% of the SMR of these ATPase (19~ZS%), Ca2+-ATPase (4-8%), actinomyosin
animals can be accounted for by resting tissue oxygen ATPase (2-s%>, gluconeogenesis (7-lo%), and ureagen-
consumption. Thus on the order of 20% of SMR may be esis (3%). The sum of these estimated contributions is 60-
concerned with service functions by this estimate. How- 87%. Other major contributors are substrate cycling and
ever, the difference between in vitro and in vivo oxygen RNA synthesis. Obviously, the contribution of different
consumption rates might have many other causes unre- ATP users differs dramatically in different organs.
lated to service functions, so this is unlikely to be a reli-
able estimate.
A potentially better estimate may be obtained by con- F. Reactions That Uncouple Standard Metabolism
sidering the individual organs. It has been estimated that
the arrested heart consumes -15% of the oxygen con- We have, until now, discussed many of the contribu-
sumption of the rat heart in vivo (48, 80, ZOO), thus -85% tors to SMR in terms of coupling reactions such as the
of the oxygen consumption is concerned with the service Na’-K’- and Ca2+-ATPases and protein synthesis. How-
function. In the brain, light anesthesia or sleep reduces ever, it is important to consider which reactions uncouple
brain oxygen consumption by lo-30%, whereas deep an- these processes, since without the uncoupling reactions
esthesia, removing almost all electrical activity, reduces there would be no SMR.
oxygen consumption by -4O-50% (4, 68); thus -40% of In the brain, about one-half of the ATP production is
oxygen consumption is concerned with service functions coupled to the Na+-K’-ATPase and ion cycling at the pre-
in brain. In the kidney, 50-60% of oxygen consumption is and postsynaptic membranes (4, 68, 220), and thus the
thought to be coupled to sodium transport (53, 98, 197), reactions uncoupling most of brain metabolism are the
and up to 5% is coupled to gluconeogenesis (88,197); most voltage- and neurotransmitter-gated ion channels in these
of this probably has a service function. Resting skeletal membranes. In the kidney, most of energy metabolism
muscle performs no service function (unless we include (50-60%; Refs. 53, 98, 197) is estimated to be involved in
heat production, which all tissues produce). The gut per- sodium reabsorption, so the uncoupling reactions of this
forms no obvious service function in the standard state, organ are the sodium permeability of the luminal mem-
since this state is postabsorptive. However, in the liver, branes of the tubule and glomerular filtration membranes.
gluconeogenesis and urea synthesis (and protein synthe- In heart and muscle, a smaller but still significant propor-
sis for export) may be classified as service functions and tion of metabolism is uncoupled by voltage- and ligand-
have been estimated above to require between 40 and 50% gated ion channels in their membranes. Protein degrada-
of liver SMR in humans (see sects. IIC?? and IIC9). If 40% tion uncouples protein synthesis in all tissues and thus
of brain, 60% of kidney, 50% of liver, and 85% of heart accounts for -20% of SMR (see sect. IIC~). Muscle con-
oxygen consumption was used for service functions in traction in heart alone may account for between 1 and 6%
humans, the sum of the oxygen consumption driving these of SMR (see sect. 11c3), and thus muscle relaxation, cross-
functions would account for -30% of the total SMR in bridge cycling of actinomyosin complexes, must account
humans and -20% in rats. for a small but significant uncoupling of metabolism.

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
J2dy mu’ ORIGIN OF STANDARD METABOLIC RATE 743

The mitochondrial proton leak uncouples very tion, an insignificant amount from net carbohydrate oxida-
roughly 20% of SMR and may be particularly important in tion, and -10% from amino acid oxidation (55).
skeletal muscle (see sect. IBZ). This proton flux appears What proportion of total ATP production is produced
to be mediated by the bilayer (39). Around Z-50% of the by glycolysis rather than mitochondria in the standard
proton leak of isolated mitochondria is apparently due to state? During glucose oxidation, 2 ATP/glucose are pro-
the protonophoric activity of free fatty acids in the bilayer duced by glycolysis and -23 ATP/glucose by the mito-
(39). However, it is unclear whether this free fatty acid chondria (assuming a maximal ATP production of 31 ATPI
effect is present in vivo. glucose with 20% uncoupling by the proton leak, see Ref.
Nonmitochondrial oxygen consumption in heart, 19); during glycogen oxidation, the glycolytic ATP pro-
liver, and skeletal muscle alone may account for -10% duction would be 4 ATP/glucose unit. Thus, if 30% of SMR
of rat SMR (see sect. IIN). This oxygen consumption is is derived from carbohydrate oxidation, then 2.4-4.4%
catalyzed by a heterogeneous group of enzymes and so is (0.3 X 2/25 or 0.3 X 4/27) of total ATP production is de-
uncoupled by a large number of different reactions. rived from glycolysis during glucose or glycogen oxida-
Cellular energy metabolism may be uncoupled not tion. However, glycolysis may also produce lactate, which
just by uncoupling processes in the same cell, but also by may recycle to glucose via gluconeogenesis in the liver.
uncoupling processes elsewhere in the body or outside Because no lactate (or alanine) accumulates in the stan-
the body in the environment. These uncoupling processes dard state, the rate of this (anaerobic) glycolysis can be
are often transport processes, where cellular energy me- estimated from the rate of gluconeogenesis. In section
tabolism is coupled to a transcellular transport process 11c8, we saw that the rate of glucose turnover in the stan-
and the transport is reversed either in the environment dard state has been measured to be 0.72 mol glucose/day
or by other transport processes in the body. For example, in humans and 14 mmol glucose/day in rats, and this is
Na+ is transported from one extracellular space to an- equivalent to 1.44 mol glycolytic ATP/day in humans and
other through intervening cells in the gut and kidney. This 28 mmol glycolytic ATP/day in rats (assuming 2 glycolytic
process requires a continuous input of free energy, but ATP/glucose). The total oxygen consumption is 32 mol O/
no free energy is stored, because Na’ derived from the day in humans and 490 mmol O/day in rats (see sect. 11C8),
environment (as food) is returned to the environment (as and with the assumption of an effective P/O of 1.8, this
urine). The pumping of the blood around the body by the is equivalent to 57.6 mol mitochondrial ATP/day in hu-
heart is another process that does not come to equilibrium mans and 882 mmol mitochondrial ATP/day in rats. Thus
because the transport is in a loop (and thus the pump we can estimate that anaerobic glycolytic ATP production
is short circuited), which requires a continuous input of is 2.4% of total (glycolytic and mitochondrial) ATP pro-
energy. The movement of limbs and the body by the alter- duction in humans and 3.1% in rats in the standard state.
nate contraction and relaxation of antagonistically ar- The total contribution of glycolytic ATP production (aero-
ranged skeletal muscles is a process that again conserves bic and anaerobic) is thus roughly 4.6-7.7%. Of course,
little free energy, but results in transport against frictional the proportion varies with tissue; most glycolysis is oc-
forces (although this is not involved in SMR). Gluconeo- curring in the brain in the standard state.
genesis in liver and kidney is uncoupled by glycolysis in
other tissues, and this cycle functions to transport energy
between tissues. These uncoupling processes at the supra- H. Summary
cellular level are related to the service functions discussed
in section ID.
Contributions to SMR can be quantified in terms of
coupling to oxygen consumption, ATP turnover, or uncou-
G. Substrate Utilization and Glycolytic ATP pling. In the standard state, -90% of mammalian oxygen
Production in Standard Metabolic Rate consumption is used by the mitochondria; of this 90%,
-20% is uncoupled by the mitochondrial proton leak,
The relative contributions of carbohydrate, fat, and whereas the other 80% is coupled to ATP synthesis. Of
protein oxidation to SMR can be estimated from the respi- this ATP production, -28% is used by protein synthesis,
ratory quotient (ratio of COB produced to O2 consumed) 19-28% by the Na+-Kf-ATPase, 4-8% by the Ca’+-ATPase,
and urea production (189). In humans in the postabsorp- 2-8% by actinomyosin ATPase, 7- 10% by gluconeogene-
tive state, the respiratory quotient is -0.82 (189), and the sis, and 3% by ureagenesis, with mRNA synthesis and sub-
relative contributions can be estimated to be -60,30, and strate cycling also contributing. Protein synthesis is un-
10% for fat, carbohydrate, and protein oxidation, respec- coupled by protein degradation, the Na’ pump by Na+
tively. With longer fasting, the respiratory quotient falls channels and Na+-coupled transport, the calcium pump
further and has been measured to be 0.7 in 48-h starved by Ca” channels, and gluconeogenesis by glycolysis. Con-
rats, indicating that 90% of SMR is derived from fat oxida- tributions differ in different tissues (Table 2 and Fig. 3).

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
744 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

Total Total
oxygen ATP
Total
consumption consumption
mitochondrial
in the in the
oxygen
Standard Standard
consumption
State State

FIG. 3. Estimated contribution of processes


Protein to energy utilization in standard state. First col-
synthesis umn shows contribution of mitochondrial (80%)
and nonmitochondrial (10%) oxygen consumption
to total respiration rate in standard state. Second
column shows proportion of mitochondrial respi-
ration used to drive ATP synthesis (80%) and pro-
Na+/K+ ton leak (20%) in standard state. Data shown in
Coupled to ATPase first and second columns were calculated using
Mitochondrial data for proton leak and nonmitochondrial oxy-
ATP synthesis me------ gen consumption from liver, heart, and skeletal
Ca2+ muscle only and ignoring possible contribution of
ATPase these processes in other tissues (see section II@.
Third column represents contribution of ATP-con-
Gluconeo- suming processes to total ATP consumption in
genesis
m-B----- standard state calculated as outlined in section
HE.
Ureagenesis
-----a--
Actinomyosin
--ew AVasie - -
-------a Others
\
\
(including RNA
Uncoupled by \ synthesis and
\
proton leak substrate cycling)
Non- \
\
Mitochondrial 4

III. CONTRIBUTION OF FEEDING, COLD, AND A. Cold-Induced Thermogenesis


MUSCLE USE TO FIELD AND MAXIMAL
METABOLIC RATE Exposure to temperatures below thermoneutrality
causes increases in metabolic rate (105), and increases
In this section, we very briefly summarize the major of up to five times SMR (16) have been noted in cold-
processes involved in field and maximal metabolic rate. acclimatized mammals in response to catecholamines, the
This section is intended to put the above discussion in its principal hormones involved in cold-induced thermogene-
context rather than to review the subject in detail, a task sis. In addition, long-term cold exposure may also alter
which has alrea.dy been fulfilled by others (e.g., Refs. 15, standard metabolism; for example, the rate of oxygen con-
110, 133). The value of the field metabolic rate (assessed, sumption of cold-acclimatized rats is >25% higher at ther-
using the doubly labeled water method, as the 24-h meta- moneutral temperature than that of warm-acclimated rats
bolic rate of nonbreeding free-living animals) has been (estimated from the data of Ref. 75). The response to cold
calculated to be roughly three times greater than SMR is divided into shivering and nonshivering thermogenesis.
(measured under laboratory conditions) for mammalian Shivering thermogenesis results from muscle tremor and
species in the wild (151) and roughly twice SMR for do- thus increased activity of the actinomyosin ATPase, Na+-
mestic animals (15). However, in humans, the field meta- Kf-ATPase, and Ca2+-ATPase, leading to increased oxida-
bolic rate (based on measurements of heat production tive phosphorylation in skeletal muscle. However, if cold
made while the subject carried out normal daily activities) exposure occurs for longer periods (several weeks), heat
may be between 1.5fold (for office workers) and 2. l-fold production by shivering is gradually replaced by nonshiv-
(for heavy occupational work) greater than the SMR (15). ering thermogenic processes. Although a large proportion
The major components of these increases in metabolic of this nonshivering thermogenesis is the result of in-
rate are muscular activity, shivering and nonshivering creased activity of BAT (75), other tissues, particularly
thermogenesis, and the thermogenic effect of food. The skeletal muscle, may contribute (54, 75, 117). Cold accli-
relative contributions of these processes will obviously matization studies in mammals indicate that there is a
depend on the level of locomotion and exposure to cold. general increase in the ion permeability of membranes in

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 745

a variety of tissues, resulting in increased activity of the sect. I). The extra oxygen consumption and heat produc-
Na’-K’-ATPase and Ca’+-ATPase (47,53). It has been esti- tion is the result of muscle contraction in heart and skele-
mated that -4O-50% of the response to cold (6°C) in cold- tal muscle. About 87% of body oxygen consumption is
acclimatized rats is due to nonshivering thermogenesis in consumed by skeletal muscle during heavy work in hu-
BAT (73). The increase in oxygen consumption of adult mans and 5% by the heart (145). The estimated relative
cold-acclimatized rats in response to cold (6°C) is around ATP consumption by different ATP users in isometrically
two times SMR (75, 117). Adult humans lack BAT and contracting muscle is 6580% by the actinomyosin
may have relatively little cold exposure depending on con- ATPase, lo-25% by the Ca’+-ATPase, 5-10% by the Nat-
ditions so that the contribution of cold-induced thermo- K’-ATPase, and -5% by myosin phosphorylation (53, 109,
genesis to field metabolic rate is probably smaller in hu- 110). In shortening muscle, the oxygen consumption is
mans than other mammals, although this obviously de- higher, and calcium cycling may require 20-50% of the
pends on environmental conditions. ATP turnover (53, 109). During isometric contraction of
the heart, 50-60% of ATP turnover has been attributed to
the actinomyosin ATPase, 15-35% to the Ca2+-ATPase,
B. Feeding-Induced Thermogenesis
and 15%to noncontractile functions (48, 53, 200). Muscle
contraction also does some work including the compres-
Feeding has both short-term (i.e., after a meal) and
sion and acceleration of blood in the heart and limbs, and
long-term (over- or underfeeding) effects on metabolic
air in the lungs, and this energy is dissipated as heat in
rate. The short-term effect (formerly known as specific
the blood vessels and lungs/external air, respectively. The
dynamic action) typically stimulates metabolic rate by
overall efficiency of the heart is -20% but rises to 40%
-25%, with peak stimulation l-2 h after a meal in hu-
or higher if the energy requirement for activation and
mans. This thermogenic effect of feeding contributes
noncontractile functions is subtracted (80,200). Up to 20-
- 15%of field metabolic rate in omnivores and up to 30%
30% of the calorific input can be used for external work
in ruminants (15). The mechanisms of this stimulation are
by skeletal muscle during walking, running, and cycling
still not entirely clear but include 1) the stimulation of
in humans (144). The efficiency of performing external
pathways leading to the temporary synthesis of storage
work apparently increases with the body size of species
compounds, 2) the use of the ingested food as substrate
(15). The uncoupling reactions of maximal metabolic rate
for generating ATP, and 3) hormonal changes, in particu-
are thus mainly 1) calcium and sodium channels, 2) mus-
lar, activation of the sympathetic nervous system. One
cle relaxation, 3) cross-bridge cycling of the actinomyosin
important component of the thermogenic response is the
during maintained contraction, and 0) dissipation of the
metabolism of ingested amino acids in the liver resulting
physical work either inside or outside the organism.
in glucose, fat, urea, and protein synthesis, accompanied
by amino acid oxidation (119). Swallowing, hydrolysis/
fermentation and absorption of the food, and enzyme se- IV. HEAT PRODUCTION AND FREE ENERGY
cretion all make very small additional contributions (15). DISSIPATION
Baldwin and Smith (6) estimated that after a carbohydrate
meal the increased heat production could be attributed
A. Thermogenesis and Enthalpy Dissipation
to conversion of glucose to fat (30% of the increase), con-
version of glucose to glycogen (14%), hydrolysis of the
The cellular production of heat and dissipation of free
carbohydrate in the gut (5%), synthesis of digestive pro-
energy have sometimes been confused. The free energy
teins (IS%), and active transport during absorption (24%).
change of a reaction @G) is made up of two components:
Milligan and Summers (147) and Kelly and McBride (124)
the enthalpy change (LH) and the entropy change @S),
attributed most of the increase in energy turnover to in-
where AG = LW-TAS (and T is the absolute tempera-
creases in protein turnover and Na’-Kf-ATPase after feed-
ture). For cellular reactions and transport processes, the
ing. Nutritional status (as distinct from the thermogenic
entropy component (-TLS) of AG is often large so that
effect of food) also affects SMR (179). Fasting lowers
the enthalpy change cannot be equated with the free en-
metabolic rate by up to 40% (15), and overeating may
ergy dissipation (AG). The heat production of an enzyme
increase it (179, HO), although this effect is more contro-
or transport reaction is given by -AH of the overall reac-
versial (15, 103, 180).
tion so that the heat production by a cellular reaction is
given by -AH X J, where J is the net rate of the reaction.
C. Muscle Use-Induced Thermogenesis Prusiner and Poe (168) pointed out that most of cellu-
lar heat production must occur in the process of ATP
Maximal metabolic rate attained by exercise for short production as opposed to ATP usage. They showed this
periods is roughly 10 times SMR (204), although this figure by comparing the enthalpy of substrate (glucose and fatty
varies considerably between and even within species (see acid) oxidation to that of ATP hydrolysis (after accounting

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
746 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

TABLE 3. Estimated standard molar enthalpy change of IV (succinate oxidation) evolves -62 kJ/mol 0 (sub-
cellularr reactions tracting the AH of 6 Hf from the AH of succinate oxida-
tion, assuming 6 H+/O for succinate), the ATP synthase
-AH”
evolves -24 kJ/mol ATP (subtracting the AH for ATP
Reaction kJ/mol kJ/mol 0 hydrolysis from the AH for 3 Ht assuming 3 H+/ATP)
equivalent to 48 kJmo1 0 (assuming an effective mito-
a: Glucose + 12 0 + 6 CO2 + 6 HZ0 2,803 234
b: Palmitate + 46 0 -+ 16 COP + 16 HZ0 10,014 218 chondrial P/O of Z), the adenine nucleotide translocator
c: Lactate + 6 0 --) 3 CO2 + 3 HZ0 1,367 228 evolves -21 kJ/mol ATP (assuming one net charge trans-
d: NADH2 + 0 + NAD’ + HZ0 256 256 located) equivalent to 42 kJ/mol 0 (assuming an effective
e: ATP + Hz0 + ADP + Pi 20 40
f: H,=, + H,+, (mitochondria) 15 150 mitochondrial P/O of Z), and the proton leak evolves -21
g: Succinate + 0 -+ fumarate + HZ0 \ 152 152 kJmo1 H’ equivalent to 42 kJ/mol 0 (assuming 10 H’ are
pumped per 0 and 20% of these protons return via the
Enthalpy changes (m) at 25°C are calculated for reactions as
shown and per mole of atomic oxygen. All enthalpies are negative. Val- proton leak). These estimates are crude because the mito-
ues do not include buffer ionization or metal binding. Enthalpy changes chondrial membrane potential in vivo is not known accu-
for reactions a-c are from Blaxter (15), and those for reaction d are rately. A total of -80% of cellular heat production is
from Poe et al. (163) and Burton (42). Standard enthalpy changes for
reaction e are from Podolsky and Morales (162) and are corrected for evolved by the proton pumps and transporters of the mito-
buffer ionization. Enthalpy of ATP hydrolysis per atom of oxygen is chondrial inner membrane, as pointed out by Prusiner
calculated assuming a ATP/O of 2. AH including buffer ionization has and Poe (168). However, expressing cellular heat produc-
been estimated to be 47 kJ/mol ATP in muscle (58) and 15 kJ/mol (11).
For reaction f, values for mitochondrial proton transport are enthalpy tion in percentage terms can be misleading because many
equivalent of mitochondrial membrane potential taken as 150 mV, and cellular reactions may be endothermic, and considerable
value per mole of atomic oxygen is calculated assuming a H’/O of 10 amounts of heat may be taken up or produced during
(see text).
binding or release of metabolites, protons, and metal ions
by other molecules (58, 209).
for the ATP/O). Table 3 gives literature values of some If we assume that the average plasma membrane po-
enthalpy changes and the potential energy change esti- tential is 60 mV (51, 107, 156), then the Na’-Kf-ATPase
mated for protons crossing the mitoc hondrial inner mem- evolves 15 kJ/mol ATP [subtracting the enthalpy equiva-
brane (excluding enthalpies of buffer ionization). lent of one charge transfer (6 kJ) from the enthalpy equiv-
The enthalpy change for oxidation of a particular sub- alent of ATP hydrolysis (21 kJ)], equivalent to 7 kJ/mol 0
strate by NAD can be estimated by comparing the en- [multiplying the heat production per ATP (15 kJ/mol ATP)
thalpy of combustion of the substrate to that of NADH by the fraction of total ATP consumption used by the Na+
per oxygen atom consumed. From Table 3 we can see pump (0.23; see sect. IL%‘) and by the effective mitochon-
that little or no heat production occurs during glycolysis drial P/O (2; see sect. II&?)]. The return of Na’ back into
or the oxidation of glucose or fatty acids by NAD. How- the cell is associated with -6 kJ/mol Na’ (one charge
ever, high ratios of glycolysis to oxidative metabolism can moved through 60 mV) equivalent to 8 kJmo1 0 [multi-
result in a significant heat production from glycolysis in plying the AHINa’ (6) by the Na’/ATP stoichiometry of
isolated cells, partly due to the net pH change (82). Glycol- the pump (3) and the fraction of total ATP consumption
ysis from glucose to lactate produces 63-80 kJ/mol lactate used by the pump (0.23) and the effective mitochondrial
when buffer ionization is included (82) (equivalent to lo- P/O (a)] if the return is not coupled to other reactions,
14 kJ/mol 0). The heat production by individual steps in and -5.5 kJ/mol 0 (6 x 2 x 2 x 0.23) with the return of
glycolysis has been considered in Reference 58. In con- K’ to the exterior of the cell. These estimates are crude
trast to NAD reduction, the oxidation of NADH evolves a because the contribution of the Na’-K+-ATPase to ATP
considerable amount of heat, mostly produced mitochon- turnover and the level of the plasma membrane potential
drial respiration, oxidative phosphorylation, and proton varies from tissue to tissue. The ATP (and GTP) consump-
leak. The enthalpy of combustion of NADH is about -250 tion associated with protein synthesis can be estimated
kJ/mol 0, whereas the equivalent potential heat produc- to produce 12 kJ/mol 0 [multiplying the AH of ATP hydro-
tion from the mitochondrial membrane potential is 150 kJ/ lysis (21 kJ/mol ATP) by the fraction of total ATP con-
mol0 (assuming an H’/O of 10 and membrane potential of sumption used by protein synthesis (0.28; see sect. IL%‘)
150 mV), and the enthalpy of ATP hydrolysis is roughly and the effective mitochondrial P/O (a)]. Because rela-
-40 kJ/mol 0 (assuming an effective mitochondrial P/O tively little energy is stored in the peptide bond (-5%),
of 2). The missing enthalpy is lost as heat. From Table 2 most of this heat will be evolved in protein synthesis
and the estimated H+ stoichiometries, we can estimate rather than protein breakdown. The ATP (and GTP) is
that complex I evolves 20-40 kJ/mol 0 (subtracting the used in several different steps of peptide chain elongation,
AH of succinate oxidation from the AH of NADH oxida- and heat evolution is also likely to be spread through
tion and then subtracting the AH of 4H+/O assuming an several steps, but the enthalpy values of the intermediates
H’/2e- of 4, see Ref. 37 and sect. IBM), complex III and are not generally known. Heat evolution during muscle

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 747

contraction has been studied in some detail, and the en- TABLE 4. Estimated free energy change of selected
thalpy of a number of actinomyosin intermediates has cellular reactions
been estimated (58, 109). If Z-7% of total ATP turnover
-AG
is via the actinomyosin ATPase in the standard state (as -AG"',
very roughly estimated in sects. IL!? and 11c3), then the Reaction kJ/mol kJ/mol kJ/mol 0

steady-state heat production by the actinomyosin ATPase


a: Glucose + 12 0 + 6 COz + 6 Hz0 2,880 2,808 234
and cross-bridge cycling of intermediates is small in this b: Glucose + 2 ADP -+ lactate +
state. Of course, this proportion is much higher during 2 ATP 136 92 8
c: NADH:! + 0 + NAD+ + Hz0 219 202 202
maximal metabolism when 50-80% of the oxygen con-
d: H,=, -+ HL (mitochondria) 15 18 180
sumed is coupled to the actinomyosin ATPase, and thus e: ATP + Hz0 + ADP + Pi 31 63 126
up to 40 kJ/mol 0 may be generated by the actinomyosin
ATPase. For shortening muscle, a proportion of the en- Free energy changes given are standard at pH 7 (AGO’), actual
(AG), and actual per atom of oxygen associated with reaction, at 25°C.
thalpy change may not be lost as heat but rather passed on All AG values are negative. Reaction a: AGo’ value from Blaxter (15);
as internal or external work. The heart obviously performs AG calculated for cellular concentrations: 10 mM glucose, 1 mM CO2
work in accelerating blood through the vessels, and this and 20 PM Oz. Reaction b: AGo’ value from reactions a and e; AG
calculated for 10 mM glucose, 2 mM lactate and AG ATP from reaction
work is dissipated as heat within the blood vessels as the e. AG per 0 is nG/12. Reaction c: values were calculated using E”’ for
blood is decelerated by the vessel walls or by turbulence. oxygen/water couple of +815 mV and Eh of 788 mV for 20 ,uM O2 (37);
The ratio of work done to AH of ATP or PCr has been an E”’ for NADIUNAD couple of -320 mV and Eh of -260 mV (Eh may
be as low as -320 mV in heart) (69). AE,, was converted to AG by
estimated to be 40- 100% in heart so that in some condi- multiplyng by 2F. Reaction d: values quoted are median values (see
tions the heat evolved by the cross-bridges themselves text); AG per 0 is calculated on basis of 10 H+/O (see text). Reaction
may be smaller than expected (80). Also, the work of e: cytoplasmic AG for ATP has been estimated to be 65 kJ/mol in muscle
(go), 63 kJ/mol in liver (113), 60-65 kJ/mol in heart (62), and 62 kJ/mol
intestinal or uterine smooth muscle is dissipated inter- in brain (68). AG per 0 is calculated assuming ATP/O of 2.
nally. In contrast, skeletal muscle may perform external
work (up to ZO-30% of calorific input) (144, 219).
Many of the reactions of metabolism produce or con- in muscle may involve significant transient heat changes
sume protons, and these protons are buffered by intracel- (109) that do not contribute to the steady state. The reac-
lular buffering groups. The binding and debinding of pro- tions of Table 3 have been corrected for buffer ionization
tons on these various groups produce or take up heat. and metal binding.
The amounts of heat involved are large and vary with
the type of buffer (58, 209). The M of ATP hydrolysis
corrected for buffer ionization is about -20 kJ/mol ATP, B. Free Energy Dissipation and Efficiency
whereas the observed LW of ATP hydrolysis including
buffer protonation varies from -47 to -15 kJ/mol ATP Most of the internal energy (U) or enthalpy (H) of
and depends on which intracellular buffers are present substrates is dissipated by the processes of ATP produc-
(11, 58, 168). If we used the largest of these values (LW = tion. In contrast, most of the free energy (G) is dissipated
-47 kJ/mol ATP) to quantify where the heat production by the processes of ATP consumption and the uncoupling
occurs, we would estimate that -40% of heat production reactions. Table 4 gives literature estimates of the free
(2 x 47/230, where we assume an effective P/O of 2 and energy dissipation by various reactions in energy metabo-
a LUY of glucose or NADH oxidation of 230 kJ/mol 0) lism. These estimates depend on the measured metabolite
occurs in the processes of ATP utilization (plus coupled concentrations and thus potentially vary with tissue and
and uncoupling processes and proton binding to the metabolic state; however, the measured variation is small
buffer), whereas -60% occurs during the processes of in terms of this analysis.
ATP synthesis (including heat uptake due to ionization of If a (driving) reaction A+C is coupled to another
the buffer). If the intracellular buffers take up a large (driven) reaction B+D, then the efficiency of the coupling
amount of heat on ionization (as appears to happen in can defined as
skeletal muscle), then part of the heat released during net
ATP synthesis is taken up by the buffers as they ionize to JBmDx AGDmB
provide the 0.8 Hf required per ATP synthesized, and this JAeC x AGAmC
heat is released again during net ATP hydrolysis. How-
ever, in the steady state, there is no net uptake or release where JASCis the rate of conversion of A to C and JBwDis the
of these protons, and therefore, the heat exchanges in- rate of conversion of B to D, and nGAsC is the difference in
volved in this buffering are not relevant to the steady free energy between A and C and aGDmB is the free energy
state. Similarly, the binding and unbinding of metal ions difference between D and B. From this definition, it fol-
with metabolites and proteins is associated with heat up- lows that an efficiency of less than one can arise either
take or release. For example, calcium binding and release from 1) an uncoupling reaction or slip reaction from A to

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
748 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

C, so that some of A is converted to C without B being the Na’ pump), and protein synthesis dissipates -14%
converted to D, or 2) a fully coupled reaction displaced (assuming protein synthesis uses 28% of total ATP at 5%
from equilibrium. Efficiency is not directly related to heat efficiency). This leaves -33% of the free energy dissipa-
production, since efficiency is defined in terms of free tion unaccounted for, part of which will be due to the
energy rather than enthalpy. Ca’+-ATPase, RNA turnover, and substrate cycling. These
The efficiency of mitochondrial oxygen consumption calculations assume that fatty acids or glucose is being
and oxidative phosphorylation depends on the effective oxidized. Oxidation of amino acids is a little less efficient
P/O and is maximally 78% (assuming a P/O of 2.5 from (15, 119), but the relative contribution of amino acid oxi-
NADH, a AG of 202 kJ/mol 0 for oxidation of NADH, and dation to energy metabolism is small (see sect. IIG).
a AG of 63 kJ/mol ATP for ATP hydrolysis, see Table 4, Most of the free energy dissipation by glycolysis is
then efficiency = 2.5 X 63/202) and minimally 62% (assum- thought to occur at hexokinase, phosphofructokinase,
ing an effeetive mitochondrial P/O of 2.0). Most of this and pyruvate kinase (129). Most of the free energy dissipa-
inefficiency and free energy dissipation is due to the pro- tion of substrate dehydrogenation is thought to occur at
ton leak, whereas the component processes of oxidative pyruvate, isocitrate, and oxoketoglutarate dehydrogenase
phosphorylation are relatively efficient (see below). (24, 129). Most of the free energy dissipation by the respi-
In contrast to ATP production, the ATP-utilizing reac- ratory chain occurs at cytochrome oxidase (36, 37, 150).
tions are relatively inefficient. The efficiency of the Na’- Mitochondrial ATP synthesis and transport appears to be
K’-ATPase has been estimated to be -57% in brain neu- close to equilibrium so that little free energy is dissipated
rons (35 kJ/mol ATP transferred to the Na+, Kf, and elec- by these reactions. If 20% of the respiratory protons return
trical gradients of the plasma membrane with AGATP of via the proton leak, then the leak acts as a major free
62 kJ/mol) and 67% in glial neurons (42 kJ/mol transferred energy dissipating reaction. Gunter and Jensen (89, 118)
for the same assumed AGATP) (68). Similar efficiencies have measured the efficiencies of component steps of oxi-
have been calculated for the Na+-K’-ATPase of squid ax- dative phosphorylation in isolated mitochondria and esti-
ons and mammalian erythrocytes (202). However, an effi- mated the power losses in state 3 (i.e., the maximal rate
ciency of 85% has been estimated for resting heart cells of ATP synthesis) to be 12% by electron transport, 10% by
(59). The efficiency of the plasma membrane Ca”-ATPase proton leakage, 19% by substrate and product transport,
has been estimated to be -42% in brain based on a and 10% by phosphorylation, resulting in the transfer of
ATP:Ca?H+ stoichiometry of 1:l:Z and a plasma mem- 48% of the initial free energy to ATP. However, as we have
brane AG(Ca2’) of 26 kJ/mol(68). During protein synthe- seen, the in vivo efficiencies are probably higher at least
sis, 4 mol ATP (AG -63 kJ/mol) are hydrolyzed per mole during standard metabolism, because the AGATP is
of peptide bonds formed (AGO -13 kJ/mol) (15). Thus higher than in state 3. The efficiency may be lower during
the efficiency of protein synthesis is -5%. This does not maximal metabolism.
take into account the relative concentrations of amino
acids and protein or the sequence of amino acids in the
peptide chain, but this is unlikely to alter the conclusion C. Distribution of Energy Dissipation
that protein synthesis is very inefficient. The gross effi-
ciency of muscle contraction (that is, mechanical work Most (but not all) of the internal energy dissipation
output divided by AG of substrate oxidation) is maximally and heat production from metabolism occurs at the reac-
lo-30% in heart (SO, 200) and 20-30% in skeletal muscle tions of the mitochondrial inner membrane, whereas most
(80, 219), whereas the contractile efficiency (that is, total (but not all) of the free energy dissipation occurs at the
mechanical energy output divided by the AGATP) has ATP utilizing and uncoupling reactions (Fig. 4). Why is
been estimated to be 40-80% in cardiac muscle (80, 200) this? One way to rationalize the distribution of heat pro-
and 45-77% in skeletal muscle (219). duction is to note that virtually all the free energy avail-
We can calculate the proportions of the total free able from the oxidation of NADH or glucose is enthalpic,
energy available for the oxidation of glucose dissipated with a very small entropy production, whereas roughly
by various cellular processes using the data in Table 4. two-thirds of the free energy available from ATP hydroly-
Thus glycolysis dissipates -3% of the available free en- sis is entropic (see Tables 3 and 4). Thus, even if oxidative
ergy, substrate dehydrogenation (pyruvate transport, py- phosphorylation were at equilibrium, the transfer of the
ruvate dehydrogenase, and the tricarboxylic acid cycle) free energy from the NADH2 + 0, -+ NAD+ + H20 reaction
dissipates -lo%, the respiratory chain dissipates -lo%, to the ADP + Pi -+ ATP reaction must involve the loss of
the proton leak dissipates -15%, and ATP synthesis and two-thirds of the internal energy of the former reaction
transport dissipate -4%, the Na’-Kf-ATPase plus coupled (note, however, that the enthalpy and free energy values
and uncoupling transport of Na+ and Kf dissipates - 11% used here do not include buffer ionization). It is conceiv-
(assuming ATP consumption dissipates 54% of the total able that reactions that produce a large amount of heat
AG, and 20% of the total ATP consumption is used by are confined to membranes (or other relatively rigid struc-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July kxu’ ORIGIN OF STANDARD METABOLIC RATE 749

A H loss A G loss classed as short term or long term. Short-term changes


Heat Production Free Energy dissipation include those due to cold, feeding, and muscle use (dis-
cussed in sect. III). Long-term changes in SMR include
those due to nutritional status (under- or overfeeding) or
Substrate supply
age. Differences in SMR between species are mainly due
to differences in body mass or phylogeny. Standard meta-
Respiratory chain Respiratory chain bolic rate is fairly constant for a particular individual (ig-
noring daily and seasonal variations; see Ref. 15) and is
Proton leak similar for different individuals of the same species, age,
and weight, although small intraspecies differences do
Proton leak ATP synthesis exist and several authors have proposed differences that
account for these, such as variations in lean body mass
(78, 79, 146). Standard metabolic rate changes with age;
ATP synthesis it increases from birth to maturity and falls slightly in old
age. The mass-specific SMR (metabolic rate per kg mass
of organism) declines with age (15). Standard metabolic
ATP usage rate varies with thyroid status. Individuals with lower or
higher than normal thyroid levels have lower or higher
ATP usage than normal metabolic rates, respectively (e.g., Refs. 91,
203). Standard metabolic rate varies with body mass. Big-
ger animals have lower metabolic rates per unit body
mass; for example, the metabolic rate per gram of mouse
is -25 times that of an elephant (30, 183). There are also
differences in the SMR of species of the same mass but
FIG. 4. Estimated contribution of processes to heat production
(MY) and free energy dissipation (AG). Contribution of each process different phyla, although note that these differences may
to total LW loss and total LIG loss is represented by size of correspond- be due to differences in thyroid status (112). Warm-
ing box in bar chart. ATP synthesis consists of mitochondrial ATP syn-
thase and adenine nucleotide and phosphate translocators. ATP usage
blooded mammals have roughly 4-5 times the SMR of
includes ATPases and all coupled processes and uncoupling reactions cold-blooded reptiles of the same body mass and pre-
downstream of ATP production. Substrate supply includes all processes ferred body temperature (112).
from substrates to mitochondrial NADH. Heat production is open ended,
because although most processes produce heat, some processes may
The latter part of this section deals with the molecu-
also take up heat. lar origin of the differences in SMR between species of
different body size and phylogeny. However, before that,
we examine which cellular reactions control SMR, be-
tures) to prevent the proteins involved from denaturation cause only changes in reactions that control SMR will
by the released thermic energy. actually affect SMR.
Why is such a large proportion of the free energy of
metabolism dissipated by the ATP-utilizing reactions, i.e.,
why are these reactions apparently so inefficient? One A. Control of Cellular Respiration
possible explanation is that the products of these reac-
tions, for example, proteins, the Na+, Kf, and Ca2’ gradi- The concepts of “control” and “regulation” are often
ents, and cell membrane potentials, need to vary in level confused. Ambiguity can be avoided by taking the propo-
to control cell function, and this variation must be inde- sition “x control y” to mean that a change in x causes a
pendent of ATP/ADP. An alternative explanation is that significant change in y. Whereas “x regulates y” means
the relatively low enthalpic contribution of ATP hydroly- that in vivo changes in y are brought about by changes
sis means that there is relatively little energy available per in x. Of course, x can only regulate y if x has significant
reaction, and this may cause relatively slow kinetics if the control over y in vivo. In the metabolic control theory,
ATP-utilizing reactions were endothermic. the extent to which an enzyme (or transporter) controls
a steady-state rate is quantified as a flux control coefficient
(see Refs. 40, 120, 121). This coefficient is the percent
V. CHANGES AND DIFFERENCES IN change in the steady-state rate divided by the (small) per-
METABOLIC RATE cent change in the concentration of the enzyme. For a
particular system in a particular condition, the sum of the
Differences in metabolic rate may be either within control coefficients for all enzymes over a particular flux
an organism or species or between different species or is equal to 1.
phvla. Changes in metabolic rate within an animal can be It had previously been assumed that all metabolic

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
750 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

pathways would have a single, rate-limiting step, but ex- and part may be due to differences in the ratio of relatively
tensive investigations of control within isolated mitochon- active to inactive cells in organs (e.g., neurons/glia in
dria have shown that control is shared by a number of brain; Ref. 193). However, the differences in SMR between
reactions, and the distribution of control changes with animals of varying body mass, thyroid status, and phylog-
conditions (see Refs. 34, 87, 93, 138, 217). Within cells, it eny have been shown to correlate with differences in the
is often assumed that metabolic rate is solely controlled metabolic rate of liver cells isolated from those animals
by the ATP-utilizing processes, to which the ATP-produc- (23, 97, 165) so that the decrease in SMR per kilogram in
ing processes passively respond. However, recent evi- larger animals is partly due to 1) relatively more bone,
dence suggests that processes other than those involved skin, and muscle and relatively less of the more active
in ATP consumption have significant control over oxygen internal organs; 2) a reduced proportion of relatively ac-
consumption rate. Control over cellular oxygen consump- tive cell types within organs; and 3) for a given cell type
tion has been examined in isolated liver cells (41, 97) and less active metabolism.
perfused rat skeletal muscle (177) in “resting” conditions, What metabolic changes underlie the differences in
which may approximate the standard state. Control over cellular metabolism in different species and phyla? It is
resting mitochondrial oxygen consumption in rat liver unlikely that change in the level of a single cell reaction
cells has been shown to be shared between the “substrate or process is sufficient to explain a large change in SMR,
oxidation” (29-35% of the control), “proton leak” (ZZ- because the relevant reactions would need to have high
25%) and “ATP turnover” (41-49%) pathways (41, 97). control coefficients over SMR over a large range of rates.
Note here that, because these experiments were carried This is unlikely to be true, because what evidence we
out in isolated cells, any control by substrate supply to have (see sect. VA) suggests that none of the relevant
the tissue is not included in the analysis. If significant reactions has a control coefficient over oxygen consump-
control was invested in tissue substrate supply, as some tion approaching 1. If an enzyme has a control coefficient
consider to be the case (50, 61), then the amount of con- of < 1, raising the enzyme’s activity will bring the reaction
trol invested in the cellular pathways would be less sig- closer to equilibrium and the enzyme will lose control. It
nificant. In perfused skeletal muscle, the same analysis has been shown that large increases in the rate of meta-
gave a similar control distribution: substrate oxidation bolic pathways can only be brought about by proportional
was found to have 44 t 28% of the control, proton leak activation of all the enzymes with significant control over
had 38 t 21%, and ATP turnover had 21 t 9% (177). There the pathway (70). Thus it is unlikely that a large difference
are indications in many other mammalian tissues that the in metabolic rate could result from a change in the level
ATP-utilizing reactions do not have all the control over of a single enzyme.
cellular oxygen consumption (34, 101). This is indicated A more likely hypothesis for large differences in met-
by the fact that respiratory substrates or increased mito- abolic rate is that all the major metabolic pathways
chondrial free calcium can stimulate oxygen consumption change in proportion, i.e., that all enzymes and transport-
in these tissues. In many tissues, a large amount of control ers change in concentration to the same extent as the
does lie in the ATP-utilizing processes, but this control is minimal metabolic rate. This hypothesis is discussed be-
not necessarily in the ATP-utilizing reactions themselves. low with particular reference to differences in metabolic
For example, the rate of the Na+-K+-ATPase is usually rate related to body mass, thyroid status, and phylogeny.
determined by the Na+ permeability, and the actinomyo-
sin ATPase by Ca” permeability, whereas protein synthe- 1. Body mass-related changes in metabolic rate
sis and the associated ATP usage are controlled by a large
number of processes (34). Thus it is clear that the control The SMR of adult mammals of different species is
of oxygen consumption rate is widely distributed among related to their body weight (M,) by the equation SMR
metabolic pathways and is not solely a property of the cx(MJ”.75. The following components have been shown to
ATP-consuming processes. scale in proportion to species metabolic rate in species
of different body mass: total number of mitochondria in
liver [ cy(Mb)0*72; Ref. 1961; total mitochondrial inner mem-
B. Causes of Species Differences in Standard brane area from liver, heart, kidney, brain, and skeletal
Metabolic Rate Related to Body Size muscle [c~(M~)‘*~~; Ref. 651; total liver and skeletal muscle
and Phylogeny cytochrome oxidase [cY(M~)‘*~‘-‘*~~; Refs. 115, 1301; whole
body cytochrome c content [a(Mb)Oa7; Ref. 641; whole body
What changes underlie these differences in metabolic protein turnover [cY(M~)‘-~~; Ref. 1721; whole body RNA
rate between species and phyla? Part of the differences turnover [ (x(MJ’*~~; Ref. 1851; and Na’-K+-ATPase activity
in SMR mentioned above are due to differences in the in liver and kidney slices [cY(MJ’*~~; Ref. 571. This tends to
mass of the metabolically most active organs (e.g., liver, support the proposition that changes in SMR are brought
kidney, and brain) as a proportion of body mass (15, 183), about by changes in all major cellular reactions.

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 751

In liver cells isolated from mammals of widely rang- various cellular processes to be maintained is that this
ing body mass (20-750 kg), the proportion of hepatocyte would preserve a particular pattern of control distribu-
respiration used to drive proton leak (-ZO%), ATP turn- tion. Analysis of the distribution of control over oxygen
over (-65%), and nonmitochondrial oxygen consumption consumption in mitochondria isolated from the livers of
(- 15%) was the same in all cases (167), although the different animals (from data in Ref. 167), from different
proton permeability of liver mitochondria isolated from tissues of the same animal (178), and even from plants
the same mammals varied considerably at the same pro- (60, 126) indicates that the distribution of control over
ton-motive force (164). The relationship between the oxy- mitochondrial oxygen consumption (and other variables)
gen consumption rate (per unit mass) of the liver cells appears to be a conserved property in different tissues
and the body mass (Mb) of the mammals they were iso- and species. This might be taken to indicate the impor-
lated from [(x(MJ~*‘*] was virtually identical to the mass- tance of this pattern of control. Indeed, the need to con-
specific SMR [cY(MJ -‘*17] of the animals from which they serve a certain control pattern presumably reflects the
were taken (165). importance of certain key points in a pathway that are
If the whole of metabolism sealed with body size, subject to in vivo regulation, to effect short-term changes
we would expect all metabolic intermediate levels to not in metabolic rate that are produced by alteration of the
change with body size. However, it has been reported that activity of only some points in a metabolic pathway.
the AGATP (estimated by NMR) of heart in vivo decreases
with body size in mammals (62). If confirmed in other
tissues, this would suggest that ATP production was de- C. Summary
creased more than ATP utilization with increasing body
size. In apparent contrast to this, one of the driving forces In conclusion, it appears that differences in SMR seen
for ATP synthesis, the mitochondrial membrane potential, in animals of different body mass and phylogeny are the
increases with increasing body size in isolated liver cells result of a proportional change in the activity of most
(167), perhaps implying that ATP utilization was de- cellular processes. Large-scale and long-term activation
creased more than ATP production in liver. of metabolic rate may require such proportional activation
In summary, there is evidence that most of metabo- because control over metabolic rate is distributed
lism involved in SMR scales roughly in proportion to SMR throughout metabolism so that a large activation of meta-
in mammals of different body size, but the finding that bolic rate cannot be achieved by activating only a single
metabolite levels do not remain constant suggests that or small number of metabolic reactions.
the changes in metabolism are not exactly proportionate.
VI. FUNCTIONS OF STANDARD
2. Metabolic rate and phylogeny METABOLIC RATE

Comparisons of energy metabolism in homeotherms Several different functions could be suggested for the
and poikilotherrns have shown that there is a roughly processes underlying SMR. These include 1) maintenance
fivefold higher SMR in homeotherms than in poikilo- required due to undesigned uncoupling reactions, 2) pro-
therms of the same body size and preferred body tempera- duction of heat to maintain body temperature, 3) mainte-
ture, whereas the total mitochondrial inner membrane nance and heat production for maximal metabolic rate,
area of homeotherms is three- to fourfold higher (65, 66), -4) endowment of increased sensitivity of key metabolic
the Na’ and K+ permeabilities of the cell membrane are reactions to effecters, and 5) reduction of harmful free
five- to sevenfold higher (67), and the proton permeability radical production.
of the mitochondrial inner membrane is four- to fivefold
greater in the homeotherm (23). A comparison of the iso-
lated liver cells of rat and a poikilotherm of equivalent A. Maintenance of Cellular Composition Against
body mass has shown that although the cells of the ho- Undesigned Uncoupling Reactions
meotherm had a fourfold greater oxygen consumption
rate, the proportions of oxygen consumption attributed to Some authors (e.g., Ref. 188) have suggested that the
nonmitochondrial oxygen use, the mitochondrial proton low entropy structures of cells (e.g., macromolecules, ion
leak, and total ATP turnover are similar in the two cell gradients) are continually degrading and dissipating due
types (23). The contribution of the Na’-K+-ATPase was to uncatalyzed or undesigned reactions, e.g., protein and
also found to be the same (23). Again, this suggests that fat oxidation reactions and ion leaks, so that a continual
all the components of metabolism are increased in propor- input of free energy is required to maintain the low en-
tion to metabolic rate. tropy states. It is important to distinguish between the
It is interesting to note that one of the reasons why function of coupling and uncoupling reactions here. If the
one might expect the balance between the activity of the uncoupling reactions are undesigned reactions, then the

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
752 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

coupling reactions could be regarded as functioning in if we accept that a body temperature above ambient is
maintaining the levels and organization of components advantageous, then we must also accept that heat produc-
(e.g., proteins and DNA). tion is functional.
We have seen above that the mitochondrial proton
leak is a significant uncoupling reaction and might be
regarded as uncatalyzed and/or undesigned (35,39). How-
ever, the fact that the activity of this leak can change so C. High Standard Metabolic Rate Enables High
dramatically with body mass and phylogeny suggests that Maximal and Field Metabolic Rate
the leak is partly under the control of the organism and
thus presumably designed. The Nit+, Kf, and Ca2’ conduc-
tances at the plasma membrane are almost certainly The argument that the function of SMR in homeother-
mainly mediated by ion channels rather than undesigned mic animals is to generate heat may appear somewhat
leaks (35). The extent to which protein degradation is circular, since it is generally assumed that the major ad-
ultimately caused by uncatalyzed reactions leading to pro- vantage of homeothermy is a higher metabolic rate. How-
tein denaturation is not known but is generally thought ever, the argument is not circular if the higher body tem-
to be small (102). Overall, the contribution of undesigned perature allows a higher growth rate, faster reproductive
leaks to metabolic rate is probably small, at least in mam- cycle, and higher maximal metabolic rate.
mals. Thus most of the uncoupling reactions involved in A high SMR may enable an organism to have a higher
SMR must have functional roles. The fact that poikilother- growth rate, faster reproductive cycle, and higher maxi-
mic animals of the same size and essentially the same mal metabolic rate not just because it provides a higher
structure and metabolism as homeotherms have roughly body temperature, but also because it maintains the meta-
fivefold lower SMR at the same body temperature also bolic machinery necessary for a high maximal metabolic
tends to suggest that most of the uncoupling (and cou- rate. Thus, for example, fast running requires high levels
pling) processes in homeotherms are designed. of muscle actinomyosin to cause contraction, high rates
of muscle calcium and sodium transport, high levels of
mitochondria to supply the ATP, increased vasculariza-
B. Standard Metabolic Rate as a Means tion, increased heart and lung capacity, increased glyco-
of Heat Production gen and fat storage, and increased capacity to interconvert
metabolites. Thus a higher maximal metabolic rate or
It is possible that the level of SMR in mammals is growth rate requires more metabolic machinery (i.e., more
determined by the necessity of maintaining a constant enzymes, transporters, membranes, metabolites, organ-
body temperature, and thus a rate of heat production elles, and cells), and the maintenance of more machinery
which matches the rate of heat loss. Mammals need to may require more energy (e.g., for protein turnover) even
maintain a constant body temperature of between 33 and in the standard state. Some support for this idea comes
39°C and at thermoneutral environmental temperatures, from the fact that in most animals maximal metabolic rate
the required heat is provided by standard metabolism. is always roughly 10 times SMR even though the absolute
However, if the need for heat far exceeded the need for value of SMR varies greatly. Thus we might infer that a
free energy, then we might expect heat production to be high SMR enables a high maximal metabolic rate. How-
localized to a particular organ and/or a particular uncou- ever, this explanation echoes the argument of section VIA
pling reaction early in metabolism (cf. thermogenesis in that the function of SMR is maintenance, which we have
brown adipose tissue). We have seen that heat production already dismissed, since most of the leaks that underlie
during standard metabolism is not localized to one organ, SMR appear to be designed rather than undesigned. A
and significant amounts of heat are generated both before stronger argument might be that a high maximal meta-
and after the production of ATP. This might be taken to bolic rate requires a high SMR to maintain a high body
suggest that in mammals the function of SMR is not only temperature (see sect. VI@ or increase potential for regu-
to produce heat, but also to supply free energy to drive lation (see sect. VrD).
cellular reactions whose function is not solely to generate Even in the standard state, there are a number of
heat. The production of heat is compatible with the pro- organism functions that require a continuous supply of
duction of free energy by the very same reactions, so that free energy. For example, the processing of information in
these are not mutually exclusive functions. Conversely, it the brain continues in the standard state and is of obvious
could be argued that the sole function of SMR is to pro- survival value. Brain energy metabolism appears to be
duce and use free energy and that heat is an inevitable uncoupled mainly by plasma membrane ion channels, and
waste product of this process. However, if this heat pro- these uncoupling processes participate in the information
duction were removed, then the body temperature of processing. Thus some uncoupling processes are func-
mammals would fall to that of the environment. Thus, tional in the standard state.

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF’ STANDARD METABOLIC RATE 753

D. Standard Metabolic Rate Increases the Thus it appears that the rate of generation of superox-
Potential for Regulation ide free radicals is related to the concentration of oxygen
within the cell and the degree of reduction of the species
An alternative, but not mutually exclusive, explana- that donate electrons to O2to form superoxide. One might
tion for the presence of a large SMR in mammals is that therefore suppose that ROS production would become
the presence of futile cycles (coupling and uncoupling most problematic in the resting cell, since such a cell
processes) in cellular metabolism confers on metabolism would consume less oxygen and have a higher concentra-
1) decreased transition times to new steady states, 2) a tion of reduced electron-transport-chain components that
higher sensitivity to effecters, and 3) more sites at which could donate an electron to oxygen to form superoxide.
control is exerted, as first proposed for substrate cycles This might be overcome by 1) increasing tissue oxygen
by Newsholme and Underwood (153). The rapid synthesis consumption to decrease tissue oxygen levels and 2) in-
and degradation of proteins, mRNA, phospholipids, me- creasing mitochondrial proton leak and cellular ATP utili-
tabolites, and signal molecules does not affect the steady- zation to decrease the reduction level of the mitochondrial
state levels but does decrease the transition times to new respiratory chain. Thus SMR might function to decrease
steady states and thus enables more rapid regulation of ROS production (195). However, this would seem to be a
levels. High activity of the processes producing and con- very expensive way to decrease ROS production, given
suming an intermediate relative to the net rate of produc- that there appear to be many other ways of achieving the
tion also means that relatively small changes in the pro- same ends. For example, tissue oxygen levels can be and
duction or consumption result in large changes in the net are decreased by arterial and arteriolar resistance to
rate, and thus increase the sensitivity of control. Mito- blood flow, and the respiratory chain can be and is main-
chondrial proton leak, which uncouples mitochondrial tained partially oxidized by restriction of electron entry
ATP synthesis, may increase sensitivity and decrease re- into the chain. Also, a high SMR to decrease oxygen levels
sponse time to changes in ATP utilization in the cell. The is accompanied by a high level of mitochondria, and thus
transmembrane Na+, Kf, and Ca2+ fluxes must similarly an increased potential for producing ROS. Overall, it
respond rapidly and sensitively to demand, and here seemsunlikely that decreasing ROS production is a major
again, the high level of cycling of these ions across cellular function of SMR, but it might play a minor role.
membranes may facilitate this response. Thus the uncou-
pling of metabolism may provide mechanisms for increas-
F. Summary
ing the sensitivity and rate of response of metabolic path-
ways to effecters.
Thus the functions of SMR are not fully understood.
Undesigned uncoupling reactions probably do not make
E. Standard Metabolic Rate as a Means of a significant contribution to SMR in mammals. The greater
Regulating Harmful Free Radical Production metabolic rate of homeotherms compared with poikilo-
therms points to an important role for SMR in heat genera-
tion, but this is also compatible with SMR providing free
A further function of SMR might be to control the
energy for functional coupling reactions (such as protein
production of harmful free radicals, either by the mito-
synthesis) and uncoupling reactions (such as information
chondria themselves or by nonmitochondrial processes
transfer and processing). The uncoupling of metabolism
(195). There is ample evidence that mitochondria are a
may also serve to increase the sensitivity and rate of re-
major source of cellular superoxide and hydrogen perox-
sponse of metabolic processes to effecters.
ide (95). Experiments using isolated heart mitochondria
have shown that the production of these reactive oxygen
species (ROS) is dramatically increased when mitochon- VII. CONCLUSIONS
drial state 4 oxygen consumption is inhibited at certain
point(s) in the respiratory chain. The site of free radical We have seen that there are important differences
production is probably the semiquinone of the Q cycle and distinctions between the cellular reactions that 1)
(205), but complex I may also be involved (2). It has been couple to oxygen consumption, 2) uncouple metabolism,
shown that uncoupling or the availability of ADP prevents 3) hydrolyze ATP, -4) control metabolic rate, 5) regulate
mitochondrial free radical production (49, 139), and thus metabolic rate, 6) produce heat, and r) dissipate free en-
increased ATP utilization or proton leak can decrease ergy. The quantitative contribution of different cellular
ROS production. Nonmitochondrial ROS production also reactions to these processes is difficult to estimate accu-
appears to be important, particularly via electron trans- rately at present, because of the vast amount of metabolic
port chains in the endoplasmic reticulum (95). The rate information required. However, we believe a useful first
of ROS production is roughly proportional to ambient oxy- estimate can now be made, and this will help to focus
gen concentrations (96). attention and research on outstanding problems. We esti-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
754 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

mate that -90% of mammalian oxygen consumption in methyl)aminomethane and phosphate buffers. J. Biol. Chem. 218:
961-969, 1956.
the standard state is mitochondrial, of which -20% is 12. BERRIDGE, M. J. Inositol triphosphate and diacylglycerol as sec-
uncoupled by the mitochondrial proton leak and 80% is ond messengers. Biochem. J. 220: 345-360, 1984.
coupled to ATP synthesis. Of the total ATP synthesis, 13. BERRY, E. A., AND P. C. HINKLE. Measurement of the electrochem-
ical proton gradient in submitochondrial particles. J. Biol. Chem,
-21530% is used by protein synthesis, 19-28% by the 258: 1474- 1486, 1983.
Na+-K+-ATPase, 4-S% by the Ca2+-ATPase, 2-8% by the 14. BERRY, M. N., A. M. EDWARDS, AND G. J. BARRITT. Isolated He-
actinomyosin ATPase, 7- 10% by gluconeogenesis, and 3% patocytes: Preparation, Properties and Applications. Amsterdam:
Elsevier, 1991, p. 121-178.
by ureagenesis, with mRNA synthesis and substrate cy- 15. BLAXTER, K. Energy Metabolism in Animals and Man. Cam-
cling also making significant contributions. The main cel- bridge, UK Cambridge Univ. Press, 1989.
lular reactions that uncouple standard energy metabolism 16. BLOCK, B. A. Thermogenesis in muscle. Annu. Rev. Physiol. 56:
535-577, 1994.
are the Nit+, Kf, H’, and Ca2’ channels and leaks of cell 17. BOVERIS, A., N. OSHINO, AND B. CHANCE. The cellular production
membranes and protein breakdown. Cellular metabolic of hydrogen peroxide. Biochem. J. 128: 617-630, 1972.
rate is controlled by a number of processes including met- 18. BRAND, M. D. The contribution of the leak of protons across the
mitochondrial inner membrane to standard metabolic rate. J.
abolic demand and substrate supply. The differences in Theor. Biol. 145: 267-286, 1990.
SMR between animals of different body mass and phylog- 19. BRAND, M. D. The stoichiometry of proton pumping and ATP syn-
eny appear to be due to proportionate changes in the thesis in mitochondria. Biochemist 16: 20-24, 1995.
20< BRAND, M. D., L.-F. CHIEN, E. K. AINSCOW, D. F. S. ROLFE, AND
whole of energy metabolism. Heat is produced by some R. K. PORTER. The causes and functions of mitochondrial proton
reactions and taken up by others but is mainly produced leak. Biochim. Biophys. Acta 1187: 132- 139, 1994.
by the reactions mitochondrial respiration, oxidative 21 BRAND, M. D., L.-F. CHIEN, AND P. DIOLEZ. Experimental discrim-
ination between proton leak and redox slip during mitochondrial
phosphorylation, and proton leak on the inner mitochon- electron transport. Biochem. J. 297: 27-29, 1994.
drial membrane. Free energy is dissipated by all cellular 22. BRAND, M. D., L.-F. CHIEN, AND D. F. S. ROLFE. Control of oxida-
reactions, but the major contributions are by the ATP- tive phosphorylation in liver mitochondria and hepatocytes. Bio-
them. Sot. Trans. 21: 757-762, 1993.
utilizing reactions and the uncoupling reactions. Several 23. BRAND, M. D., P. COUTURE, P. L. ELSE, K. W. WITHERS, AND
functions for SMR can be proposed, but the relative im- A. J. HULBERT. Evolution of energy metabolism. Biochem. J. 275:
portance of these is unclear. 81-86, 1991.
24. BRAND, M. D., AND M. P. MURPHY. Control of electron flux
through the respiratory chain in mitochondria and cells. Biol. Rev.
62: 141-193, 1987.
25. BRINDLE, K., S. FULTON, J. SHELDON, AND S. WILLIAMS. Probing
REFERENCES
the properties of enzymes in vivo using NMR. Biochem. Sot. Trans.
23: 376-381, 1994.
1. ALTMAN, P. L., AND D. S. DITTMER. Biologic& Handbooks: Metab- 26. BRINDLE, K., S. FULTON, J. SHELDON, AND S. WILLIAMS. Probing
olism. Bethesda, MD: FASEB, 1968. the properties of enzymes in vivo using NMR. Biochem. Sot. Trans.
2. AMBROSIO, G., J. L. ZWEIER, C. DUILIO, P. KUPPSAMY, G. SANT- 23: 376-381, 1995.
ORO, P. P. ELIA, I. TRITTO, P. CIRILLO, M. CONDORELLI, M. 27. BRINDLE, K. M., M. J. BLACKLEDGE, R. A. J. CHALLISS, AND G. K.
CHIARELLO, AND J. T. FLAHERTY. Evidence that mitochondrial RADDA. 31P NMR magnetisation transfer measurements of ATP
respiration is a source of potentially toxic oxygen free radicals in turnover during steady state isometric contraction in the rat hind
intact rabbit hearts subjected to ischemia and reflow. J. Biol. Chem. limb in vivo. Biochemistry 28: 4887-4893, 1989.
268: 18532- 18541, 1993. 28. BRINDLE, K. M., J. E. FULTON, A. J. KINSMAN, AND G. K. RADDA.
3. ASCHOFF, J., B. GUNTHER, AND K. KRAMER. Energiehaus- 31P NMR magnetisation transfer measurements of flux between
falt und Temperatur-regulation. Munich, Germany: Urban & inorganic phosphate and ATP in yeast cells over-producing phos-
Schwarzeberg, 1971. phoglycerate kinase. Biochem. Sot. Trans. 14: 1265, 1986.
4. ASTRUP, J., P. M. SORENSEN, AND H. R. SORENSEN. Oxygen and 29. BRINDLE, K. M., AND S. KRIKLER. ‘lP NMR magnetisation transfer
glucose consumption related to Na’-K’ transport in canine brain. measurements of phosphate consumption in S. cerevisiae. Bio-
Stroke 12: 726-730, 1981. chim. Biophys. Acta 847: 285-292, 1985.
5. BACHELARD, H. S. Energy utilisation by neurotransmitters. In: 30. BRODY, S. Bioenergetics and Growth. New York: Reinhold, 1945.
Brain Work: T#he Coupling of Function, Metabolism and Blood 31. BROWN, C. B., M. PRING, AND R. F. COBURN. Inositol lipid turn-
Flow in the Brain, edited by D. H. Ingvar and N. A. Lassen. Copen- over and compartmentation in canine trachealis smooth muscle.
hagen: Academic, 1975, p. 79-81. Am. J. Physiol. 256 (Cell Physiol. 25): C375-C383, 1989.
6. BALDWIN, R. L., AND N. E. SMITH. Molecular control of energy 32. BROWN, G. C. The relative proton stoichiometries of the mitochon-
metabolism. In: The Control of Metabolism, edited by J. D. Sink. drial proton pumps are independent of the proton motive force. J.
University Park: Pennsylvania State Univ. Press, 1974, p. 17-34. Biol. Chem. 264: 14704-14709, 1989.
7. BALDWIN, R. L., N. E. SMITH, J. TAYLOR, AND M. SHARP. Manipu- 33. BROWN, G. C. Total cell protein concentration as an evolutionary
lating metabolic parameters to improve growth rate and milk secre- constraint on the metabolic control distribution in cells. J. Theor.
tion. J. Anim. Sci. 51: 1416-1428, 1980. Biol. 153: 195-203, 1991.
8. BEAVIS, A. D. Upper and lower limits of the charge stoichiometry 34. BROWN, G. C. Control of respiration and ATP synthesis in mamma-
of mitochondrial electron transport. J. Biol. Chem. 262: 6165-6173, lian mitochondria and cells. Biochem. J. 284: I- 13, 1992.
1987. 35. BROWN, G. C. The slips and leaks of bioenergetic membranes.
9. BEAVIS, A. D., AND A. L. LEHNINGER. The upper and lower limits FASEB J. 6: 2961-2965, 1992.
of the mechanistic stoichiometry of mitochondrial oxidative phos- 36. BROWN, G. C., AND M. D. BRAND. Thermodynamic control of elec-
phorylation. Eur. J. Biochem. 158: 315-322, 1986. tron flux through mitochondrial bcl complex. Biochem. J. 225:
10. BERGER, M., S. A. HAGG, AND N. B. RUDERMAN. Glucose metabo- 399-405, 1985.
lism in perfused skeletal muscle. Interaction of insulin and exercise 37. BROWN, G. C., AND M. D. BRAND. Proton/electron stoichiometry
on glucose uptake. Biochem. J. 146: 231-238, 1975. of mitochondrial complex I estimated from the equilibrium thermo-
11. BERNARD, S. A. Ionisation constants and heats of tris(hydroxy- dynamic force ratio. Biochem. J. 252: 473-479. 1988.

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1.w?’ ORIGIN OF STANDARD METABOLIC RATE 755

38. BROWN, G. C., AND M. D. BRAND. The ratio of the H+/O and H’/ 64. DRABKIN, D. L. The distribution of the chromoproteins, haemoglo-
ATP stoichiometries in mitochondria over a range of np values bin, myoglobin and cytochrome c in the tissues of different species
and the thermodynamic H+/ATP stoichiometry. EBEC Short Report and the relationship of the total content of each chromophore to
5: 123, 1988. body mass. J. Biol. Chem. 182: 317-333, 1950.
39. BROWN, G. C., AND M. D. BRAND. On the nature of the mitochon- 65. ELSE, P. L., AND A. J. HULBERT. An allometric comparison of the
drial proton leak. Biochim. Biophys. Acta 1059: 55-62, 1991. mitochondria of mammalian and reptilian tissues: the implications
40. BROWN, G. C., R. P. HAFNER, AND M. D. BRAND. A “top-down” for the evolution of endothermy. J. Comp. Physiol. 156: 3-11, 1985.
approach to the determination of control coefficients in metabolic 66. ELSE, P. L., AND A. J. HULBERT. Mammals: an allometric study of
control theory. Eur. J. Biochem. 188: 321-325, 1990. metabolism at the tissue and mitochondrial level. Am. J. Physiol.
41. BROWN, G. C., P. L. LAKIN-THOMAS, AND M. D. BRAND. Control 248 (Regulatory Integrative Comp. Physiol. 17): R415-R421, 1985.
of mitochondrial respiration and ATP synthesis in isolated rat liver 67. ELSE, P. L., AND A. J. HULBERT. Evolution of mammalian endo-
cells. Eur. J. Biochem. 192: 355-362, 1990. thermic metabolism: “leaky” membranes as a source of heat. Am.
42. BURTON, K. The enthalpy change for the reduction of nicotin- J. Physiol. 253 (Regulatory Integrative Comp. Physiol. 22): Rl-
amide-adenine dinucleotide. Biochem. J. 143: 365-368, 1974. R7, 1987.
43. BUTTGEREIT, F., AND M. D. BRAND. A hierarchy of ATP-consum- 68. ERECINSKA, M., AND I. A. SILVER. ATP and brain function. J.
ing processes in mammalian cells. Biochem. J. 312: 163-167, 1995. Cereb. Blood Flow Metab. 9: 2-19, 1989.
44. BUTTGEREIT, F., M. D. BRAND, AND M. MULLER. ConA-induced 69. ERECINSKA, M., D. F. WILSON, AND K. NISHIKI. Homeostatic regu-
changes in energy metabolism of rat thymocytes. Biosci. Rep. 12: lation of regulation of cellular energy metabolism: experimental
381-386, 1992. characterisation in vivo and fit to a model. Am. J. Physiol. 234
45. BUTTGEREIT, F., M. MULLER, AND S. M. RAPOPORT. Quantifica- (Cell Physiol. 3): CBZ-C89, 1978.
tion of ATP-producing and consuming processes in quiescent pig 70. FELL, D. A., AND S. THOMAS. Physiological control of metabolic
spleen lymphocytes. Biochem. Int. 24: 59-67, 1991. flux: the requirement for multisite modulation. Biochem. J. 311:
46. CAMPBELL, S. L., K. A. JONES, AND R. G. SHULMAN. In vivo 31P 35-39, 1995.
NMR magnetisation transfer measurements of phosphate exchange 71. FIELD, J., H. S. BELDING, AND A. R. MARTIN. An analysis of the
reactions in the yeast S. cerevisiae. FEBS Lett. 193: 189- 193, 1985. relation between basal metabolism and summated tissue respira-
47. CHAFFEE, R. R. J., AND J. C. ROBERTS. Temperature acclimation tion in the rat. J. Cell. Comp. Physiol. 14: 143-155, 1939.
in birds and marnmals. Annu. Rev. Physiol. 33: 155-202, 1971. 72. FOLKE, B., AND E. NEIL. Circulation. New York: Oxford Univ.
48. CHALLONER, D. R. Respiration in myocardium. Nature 217: 78- Press, 1971.
79, 1968. 73. FOSTER, D. 0. Quantitative role of brown adipose tissue in thermo-
49. CHANCE, B., H. SIES, AND A. BOVERIS. Hydroperoxide metabo- genesis. In: Brown Adipose Tissue, edited by P. Trayhurn and D. G.
lism in mammalian organs. Physiol. Rev. 59: 527-605, 1979. Nicholls. London: Arnold, 1986, p. 31-52.
50. CHINET, A. Control of respiration in skeletal muscle at rest. Expe- 74. FOSTER, D. O., AND M. L. FRYDMAN. Nonshivering thermogenesis
rientia 46: 1194-1196, 1990. in the rat. II. Measurements of blood flow using microspheres point
51. CHINET, A., T. CLAUSEN, AND L. GIRARDIER. Microcalorimetric to brown adipose tissue as the dominant site of the calorigenesis
determination of energy expenditure due to active Na/K transport induced by noradrenaline. Can. J. Physiol. Pharmacol. 56: llO-
in the soleus muscle and brown adipose tissue of the rat. J. Physiol. 112, 1977.
(Lond.) 265: 43-61, 1977. 75. FOSTER, D. O., AND M. L. FRYDMAN. Tissue distribution of cold-
52. CHUNG, F.-Z., H. W. WEBER, AND M. M. APPLEMAN. Extensive induced thermogenesis in conscious warm and cold-acclimated rats
reversible depletion of ATP via adenylate cyclase in rat adipocytes. re-evaluated from changes in tissue blood flow: the dominant role
Proc. Natl. Acad. Sci. USA 82: 1614-1617, 1985. of brown adipose tissue in the replacement of shivering by nonshiv-
53. CLAUSEN, T., C. VAN HARDEVELD, AND M. E. EVERTS. Signifi- ering thermogenesis. Can. J. Physiol. Pharmacol. 57: 257-270,
cance of cation transport in control of energy metabolism and 1979.
thermogenesis. Physiol. Rev. 71: 733-774, 1991. 76. FREEMAN, D., S. BARTLETT, G. K. RADDA, AND B. ROSS. Energet-
54 . COLQUHOUN, E. Q., M. G., AND CLARK. Open question: has ther- its of sodium transport in the kidney: saturation transfer 31P NMR.
mogenesis in muscle been overlooked and misinterpreted? News Biochim. Biophys. Acta 762: 325-336, 1984.
Physiol. Sci. 6: 256-259, 1991. 77. GALE, C. C. Neuroendocrine aspects of therrnoregulation. Annu.
55 . CORI, C., AND G. T. CORI. The fate of sugar in the animal body. II. Rev. Physiol. 35: 391-430, 1973.
The relation between sugar oxidation and glycogen formation in
78. GARBY, L., AND 0. LAMMERT. Estimation of the sources of the
normal and insulinized rats during the absorption of glucose. J.
between-subjects variation in energy expenditure. Eur. J. Clin.
Biol. Chem. 70: 557-576, 1926.
Nutr. 48: 303-304, 1994.
56. COTTLE, W. H., AND L. D. CARLSON. Regulation of heat production
79 GARBY, L., AND 0. LAMMERT. Between-subjects variation in en-
in cold-adapted rats. Proc. Sot. Exp. Biol. Med. 92: 845-849, 1956.
ergy expenditure: estimation of the effect of variation in organ size.
57. COUTURE, P., AND A. J. HULBERT. Relationship between body
Eur. J. Clin. Nutr. 48: 376-378, 1994.
mass, tissue metabolic rate and sodium pump activity in mamma-
80 GIBBS, C. L., AND C. J. BARCLAY. Cardiac efficiency. Cardiovasc.
lian liver and kidney. Am. J. Physiol. 268 (Regulatory Integrative
Res. 30: 627-634, 1995.
Comp. Physiol. 37): R641-R650, 1995.
58. CURTIN, N. A., AND R. C. WOLEDGE. Energy changes and muscu- 81 GILL, M., J. FRANCE, M. SUMMERS, B. W. MCBRIDE, AND L. P.
lar contraction. Physiol. Rev. 58: 690-761, 1978. MILLIGAN. Simulation of the energy costs associated with protein
59. DAUT, J. The living cell as an energy transducing machine. A mini- turnover and Na+, K’ transport in growing lambs. J. Nutr. 119:
mal model of myocardial metabolism. Biochim. Biophys. Acta 895: 1287- 1299, 1989.
41-62, 1987. 82. GNAIGER, E., AND R. B. KEMP. Anaerobic metabolism in aerobic
60. DIOLEZ, P., A. K. KESSELER, F. HARAUX, M. VALERIO, K. BRINK- mammalian cells: information from the ratio of calorimetric heat
MANN, AND M. D. BRAND. Regulation of oxidative phosphorylation flux and respirometric oxygen flux. Biochim. Biophys. Acta 1016:
in plant mitochondria. Biochem. Sot. Trans. 21: 769-773, 1993. 328-332, 1990.
61. DIPRAMPERO, P. E. Metabolic and circulatory limitations to 83. GOLDBERG, N. D., T. F. WALSETH, S. J. EIDE, T. P. KRICK, B. L.
VozMAx at the whole animal level. J. Exp. Biol. 115: 319-331, 1985. KUCHN, AND J. E. GANDER. Cyclic AMP metabolism in intact plate-
62. DOBSON, G. P., AND J. P. HEADRICK. Bioenergetic scaling: meta- lets determined by “0 incorporation into adenine nucleotide alpha-
bolic design and body-size constraints in mammals. Proc. Natl. phosphoryls. Adv. Cyclic Nucleotide Protein Phosphorylation Res.
Acad. Sci. USA 92: 7317-7321, 1995. 16: 363-379, 1984.
63. DORA, K. A., S. M. RICHARDS, S. RATTIGAN, E. Q. COLQUHOUN, 84. GOLDBETER, A., AND D. E. KOSHLAND. Energy expenditure in
AND M. G. CLARK. Serotonin and norepinephrine vasoconstriction the control of biochemical systems by covalent modification. J.
in rat hindlimb have different oxygen requirements. Am. J. Physiol. Biol. Churn. 262: 4460-4471, 1987.
262 (Heart Circ. Physiol. 31): H698-H703, 1992. GRANDE. F. Energv expenditure of organs and tissues. In: Assess-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
756 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

ment of Energy Metabolism in Health and Disease, edited by S. 109. HOMSHER, E. Muscle enthalpy production and its relationship to
Calvert. Columbus, OH: Ross Laboratories, 1980, p. 88-92. actinomyosin ATPase. Annu. Rev. Physiol. 49: 672-690, 1987.
86. GREGORY, R. B., AND M. G. BERRY. The characterisation of per- 110. HOMSHER, E., AND C. J. KEAN. Skeletal muscle energetics and
fluorosuccinate as an inhibitor of gluconeogenesis in isolated rat metabolism. Annu. Rev. Physiol. 40: 93-131, 1978.
hepatocytes. Biochem. Pharmacol. 38: 2867-2872, 1989. 111. HULBERT, A. J. The evolution of energy metabolism in mammals.
87. GROEN, A. K., R. J. A. WANDERS, H. V. WESTERHOFF, R. VAN In: Comparative Physiology: Primitive Mammals, edited by K.
DER MEER, AND J. M. TAGER. Quantification of the contribution Schmidt-Nielsen, L. Bolis, and C. R. Taylor. Cambridge, UK Carn-
of various steps to the control of mitochondrial respiration. J. Biol. bridge Univ. Press, 1980, p. 129-139.
Chem. 257: 2754-2757, 1982. 112. HULBERT, A. J. A comparative study of thyroid function in reptiles
88. GULLANS, S. R., P. C. BRAZY, V. W. DENNIS, AND L. J. MANDEL. and mammals. In: The Endocrine System and the Environment,
Interactions between gluconeogenesis and sodium transport in rab- edited by B. K. Follit, S. Ishii, and A. Chandola. Tokyo: Japan Sci.
bit proximal tubule. Am. J. Physiol. 246 (Renal Fluid Electrolyte Sot. Press, 1985, p. 105-115.
Physiol. 15): F859-F869, 1984. 113. ILES, R. A., A. N. STEVENS, G. R. GRIFFITHS, AND P. G. MORRIS.
89. GUNTER, T. E., AND B. D. JENSEN. The efficiencies of the compo- Phosphorylation status of liver by 31P-NMR spectroscopy and its
nent steps of oxidative phosphorylation. I. A simple steady state implications for metabolic control. Biochem. J. 229: 141- 151, 1985.
model. Arch. Biochem. Biophys. 248: 289-304, 1986. 114. JAMES, W. P. T., AND E. C. SCHONFIELD. Human Energy Require-
90. GUPTA, R. K., AND R. D. MOORE. 31P NMR studies of intracellular ments. New York: Oxford Univ. Press, 1990.
Mg”’ in intact frog skeletal muscle. J. Biol. Chem. 255: 3987-3993, 115. JANSKY, L. Total cy-tochrome oxidase activity and its relation to
1980. basal and maximal metabolism. Nature 189: 921-922, 1961.
91. GUSTAFSSON, R., J. R. TATA, 0. LINDBERG, AND L. ERNSTER. 116. JANSKY, L. Adaptability of heat production mechanisms in ho-
The relationship between the structure and activity of rat skeletal meotherms. Acta Univ. Carol. Biol. 1: 1-91, 1965.
muscle mitochondria after thyroidectomy and thyroid hormone 117. JANSKY, L., AND J. S. HART. Participation of skeletal muscle and
treatment. J. Cell Biol. 26: 555-578, 1965. kidney during nonshivering thermogenesis in cold-acclimated rats.
92. HAFNER, R. P., AND M. D. BRAND. Effect of the proton motive Can. J. Biochem. Physiol. 41: 953-964, 1963.
force on the relative proton stoichiometries of the mitochondrial 118. JENSEN, B. D., K. K. GUNTER, AND T. E. GUNTER. The efficiencies
proton pumps. Biochem. J. 275: 75-80, 1991. of the component steps of oxidative phosphorylation. II. Experi-
93. HAFNER, R. P., G. C. BROWN, AND M. D. BRAND. Analysis of the mental determination of the efficiencies and examination of the
control of respiration rate, phosphorylation rate, proton leak rate equivalence of membrane potential and pH gradient in phosphoryla-
and proton motive force in isolated mitochondria using the top- tion. Arch. Biochem. Biophys. 248: 305-323, 1986.
down approach of metabolic control theory. Eur. J. Biochem. 188: 119. JUNGAS, R. L., M. L. HALPERIN, AND J. T. BROSMAN. Quantitative
313-319, 1990. analysis of amino acid oxidation and related gluconeogenesis in
94. HALL, S. E. H., R. GOEBEL, I. BARNES, G. HETENYI, AND M. BER- humans. Physiol. Rev. 72: 419-448, 1992.
MAN. The turnover and conversion to glucose of alanine in new- 120. KACSER, H., AND J. A. BURNS. The control of flux. Symp. Sot.
born and grown dogs. Federation Proc. 36: 239-244, 1977. Exp. Biol. 32: 65-104, 1973.
95. HALLIWELL, B. Reactive oxygen species and the central nervous 121. KACSER, H., AND J. A. BURNS. Molecular democracy: who shares
system. In: Free Radicals in the Brain, edited by L. Packer, L. the control? Biochem. Sot. Trans. 7: 1149- 1160, 1979.
Prilipko, and Y. Christen. Berlin: Springer-Verlag, 1992, p. 21-40. 122. KAUPPINEN, R. A. Proton electrochemical potential of the inner
96. HALLIWELL, B., AND J. M. C. GUTTERIDGE. Free Radicals in Biol- mitochondrial membrane in isolated perfused rat hearts as mea-
ogy and Medicine. Oxford, UK: Clarendon, 1989. sured by exogenous probes. Biochim. Biophys. Acta 725: 131- 137,
97. HARPER, M.-E., AND M. D. BRAND. The quantitative contributions 1983.
of mitochondrial proton leak and ATP turnover reactions to the 123. KAUPPINEN, R. A., J. K. HILTUNEN, AND I. E. HASSINEN. Mito-
changed respiration rates of hepatocytes isolated from rats of dif- chondrial membrane potential, transmembrane difference in NAD+
ferent thyroid status. J. Biol. Chem. 268: 14850-14860, 1993. redox potential and the equilibrium of the glutamate-aspartate
98. HARRIS, S. I., R. S. BALABAN, L. BARRETT, AND L. J. MANDEL. translocase in the isolated perfused rat heart. Biochim. Biophys.
Mitochondrial respiratory capacity and Na+- and K’-dependent Acta 725: 425-433, 1983.
adenosine triphosphatase-mediated ion transport in the intact renal 124. KELLY, J. M., AND B. W. MCBRIDE. The sodium pump and other
cell. J. Biol. Chem. 256: 10319-10328, 1981. mechanisms of thermogenesis in selected tissues. Proc. Nutr. Sot.
99. HASSELBACH, W., AND H. OETLIKER. Energetics and electrogeni- 49: 203-215, 1990.
city of sarcoplasmic reticulum calcium pump. Annu. Rev. Physiol. 125. KESSELER, A., AND M. D. BRAND. The mechanism of stimulation
45: 325-339, 1983. of state 4 respiration by cadmium in potato tuber (Solanum tubero-
100. HAUSSINGER, D., W. GEROCK, AND H. SIES. Hepatic role in pH sum) mitochondria. Plant Physiol. Biochem. 33: 519-528, 1995.
regulation: role of the intercellular glutamine cycle. Trends Bio- 126. KESSELER, A. K., P. DIOLEZ, K. BRINKMANN, AND M. D. BRAND.
them. Sci. 9: 300-302, 1984. Characterisation of the control of respiration in potato tuber mito-
101. HEINEMANN, F. W., AND R. S. BALABAN. Control of mitochondrial chondria using the top-down approach of metabolic control analy-
respiration in the heart in vivo. Annu. Rev. Physiol. 52: 523-542, sis. Eur. J. Biochem. 210: 775-784, 1992.
1990. 127. KINGSLEY-HICKMAN, P. B., E. Y. SAKO, P. MOHANAKRISHNAN,
102. HERSHKO, A., AND A. CIECHANOVER. The ubiquitin system for P. M. L. ROBITAILLE, A. H. L. FROM, J. E. FOKER, AND K. UGUR-
protein degradation. Annu. Rev. Biochem. 61: 761-807, 1992. BIL. 31P NMR studies of ATP synthesis and hydrolysis kinetics in
103. HERVEY, G. R., AND G. TOBIN. Luxuskonsumption, diet induced the intact myocardium. Biochemistry 26: 7501-7510, 1987.
thermogenesis and brown fat. Clin. Sci. 64: 7- 18, 1983. 128. KREBS, H. A. Body size and tissue respiration. Biochim. Biophys.
104. HIMMS-HAGEN, J. Cellular thermogenesis. Annu. Rev. Physiol. 38: Acta 4: 249-269, 1950.
315-351, 1976. 129. KREBS, H. A., AND H. L. KORNBERG. Energy Transformations in
105. HIMMS-HAGEN, J. Brown adipose tissue metabolism and thermo- Living Matter. Berlin: Springer-Verlag, 1957.
genesis. Annu. Rev. Nutr. 5: 69-94, 1985. 130. KUNKEL, H. O., J. F. SPALDING, G. FRANCISCIS, AND M. F. FU-
106. HINKLE, P. C., M. A. KUMAR, A. RESETAR, AND D. L. HARRIS. TRELL. Cytochrome oxidase activity and body weight in rats and
Mechanistic stoichiometry of mitochondrial oxidative-phosphoryla- three species of large animals. Am. J. Physiol. 186: 203-206, 1956.
tion. Biochemistry 30: 3576-3582, 1991. 131. KUNZ, W., F. N. GELLERICH, L. SCHILD, AND P. SCHONFELD.
107. HODGKIN, A. L., AND P. HOROWICZ. The influence of potassium Kinetic limitations in the overall reaction of mitochondrial oxida-
and chloride ions on the membrane potential of single muscle fi- tive phosphorylation. FEBS Lett. 233: 17-21, 1988.
bres. J. Physiol. (Land.) 148: 127-160, 1959. 132. KUSAKA, M., AND M. UI. Tracer kinetic analysis of Cori cycle activ-
108. HOEK, J. B., D. G. NICHOLLS, AND J. R. WILLIAMSON. Determina- ity in the rat: effect of feeding. Am. J. Physiol. 232 (Endocrinol.
tion of the mitochondrial protonmotive force in isolated hepato- Metab. Gastrointest. Physiol. 1): E136-E144, 1977.
cytes. J. Biol. Chem. 255: 1458-1464, 1980. 133. KUSHMERICK, M. J. Energetics of muscle contraction. In: Hand-

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
July 1997 ORIGIN OF STANDARD METABOLIC RATE 757

boolc of Physiology. Skeletal Muscle. Bethesda, MD: Am. Physiol. 158. PARK, E. A., AND H. E. MORGAN. Energy dependence of RNA deg-
Sot., 1983, sect. 10, chapt. 7, p. 189-236. radation in rabbit reticulocytes. Am. J. Physiol. 247 (Cell Physiol.
134. LAKIN-THOMAS, P. L., AND M. D. BRAND. Stimulation of respira- 16): C390-C395, 1984.
tion by mitogens in rat thymocytes is independent of mitochondrial 159. PAUL, R. J. Smooth muscle energetics. Annu. Rev. Physiol. 51:
calcium. Biochem. J. 256: 167-173, 1988. 331-349, 1989.
135. LAMBERTSON, C. J. Elements of normal renal function. In: Medi- 160. PEARSE, B. M. F., AND M. S. BRETSCHER. Membrane recycling by
cal Physiology (11th ed.), edited by P. Bard. St. Louis, MO: Mosby, coated vesicles. Annu. Rev. Biochem. 50: 85- 101, 1981.
1961. 161. PFALLER, W., AND M. RITTINGER. Quantitative morphology of the
136. LARDEUX, B. R., AND G. E. MORTIMORE. Amino acid and hor- rat kidney. Int. J. Biochem. 12: 17-22, 1980.
monal control of macromolecular turnover in perfused rat liver: 162. PODOLSKY, R. J., AND M. F. MORALES. The enthalpy change of
evidence for selective autophagy. J. Biol. Chem. 262: 14514- 14519, adenosine triphosphate hydrolysis. J. Biol. Chem. 218: 945-956,
1987. 1956.
137. LEMASTERS, J. J. The ATP-to-oxygen stoichiometries oxidative 163. POE, M., H. GUTFREUND, AND R. W. ESTABROOK. Kinetic studies
phosphorylation by rat liver mitochondria. J. Biol. Chem. 259: of temperature changes and oxygen uptake in a differential calorim-
13123- 13130, 1984. eter in the heart. Arch. Biochem. Biophys. 122: 204-211, 1967.
138. LETELLIER, T., M. MALGAT, AND J.-P. MAZAT. Control of oxidative 164. PORTER, R. K., AND M. D. BRAND. Body mass dependence of H’
phosphorylation in rat muscle mitochondria: implications for mito- leak in mitochondria and its relevance to metabolic rate. Nature
chondrial myopathies. Biochim. Biophys. Acta 1141: 58-64, 1993. Lond. 362: 628-630, 1993.
139. LOSCHEN, G., L. FLOHE, AND B. CHANCE. Respiratory-chain- 165. PORTER, R. K., AND M. D. BRAND. Cellular oxygen consumption
linked Hz02 production in pigeon heart mitochondria. FEBS Lett. depends on body mass. Am. J. Physiol. 269 (Regulatory Integrative
18: 261-264, 1971. Comp. Physiol. 38): R226-R228, 1995.
140. LOUD, A. D. A quantitative stereological description of the ultra- 166. PORTER, R. K., AND M. D. BRAND. Mitochondrial proton conduc-
structure of normal rat liver parenchymal cells. J. Cell Biol. 37: tance and H’/O ratio are independent of electron transport rate in
27-46, 1968. isolated hepatocytes. Biochem. J. 310: 379-382, 1995.
141. MAGNUSSON, I., D. J. ROTHMAN, B. JUCKER, G. W. CLINE, R. G. 167. PORTER, R. K., AND M. D. BRAND. Causes of the differences in
SHULMAN, AND G. I. SHULMAN. Liver glycogen turnover in fed respiration rate of hepatocytes from mammals of different body
and fasted animals. Am. J. Physiol. 266 (Endocrinol. Metab. 29): mass. Am. J. Physiol. 269 (Regulatory Integrative Comp. Physiol.
E796-E803, 1994. 38): R1213-R1224, 1995.
142. MARTIN, A. W., AND F. A. FUHRMAN. The relationship between 168. PRUSINER, S., AND M. POE. Thermodynamic considerations of
summated tissue respiration and metabolic rate in the mouse and mammalian thermogenesis. Nature 220: 235-237, 1968.
dog. Physiol. 2001. 26: 18-34, 1955. 169. RABKIN, M., AND J. J. BLUM. Quantitative analysis of intermediary
143. MARTINEAU, A., L. LECAVALIER, P. FALARDEAU, AND J.-L. CHI- metabolism in hepatocytes incubated in the presence and absence
ASSON. Simultaneous determination of glucose turnover, alanine of glucagon with a substrate mixture containing glucose, ribose,
turnover and gluconeogenesis in humans using a double stable- fructose, alanine and acetate. Biochem. J. 225: 761-786, 1985.
isotope-labelled tracer infusion and gas chromatography-mass 170. RADZIUK, J. Developments in the tracer measurement of gluconeo-
spectrometry analysis. Anal. Biochem. 151: 495-503, 1985. genesis in vivo: an overview. Federation Proc. 41: 88-90, 1982.
144. McARDLE, W. D., K. I. KATCH, AND V. L. KATCH. Exercise Physiol- 171. REEDS, P. J., M. F. FULLER, AND B. A. NICHOLSON. Metabolic
ogy (2nd ed.). Philadelphia, PA: Lea & Febiger, 1986, p. 149. basis of energy expenditure with particular reference to protein.
145. McGILVERY, R. W. Biochemistry: A Functional Approach. Phila- In: Recent Advances in Substrate and Energy Metabolism, edited
delphia, PA: Saunders, 1979. by J. S. Garrow and W. Halliday. London: Libbey, 1985, p. 46-57.
146. MILLER, A. T., AND C. S. BLYTH. Lean body mass as a metabolic 172. REEDS, P. J., AND C. I. HARRIS. Protein turnover in animals: man
reference standard. J. Appl. Physiol. 5: 311-316, 1953. in his context. In: Nitrogen Metabolism in Man, edited by J. C.
147. MILLIGAN, L. P., AND M. SUMMERS. The biological basis of mainte- Waterlow and J. M. L. Stephen. Essex, UK Appl. Sci., 1981, p. 391-
nance and its relevance to assessing responses to nutrients. Proc. 408.
Nutr. Sot. 45: 185-193, 1986. 173. RENNIE, M. J., AND J. 0. HOLLOSZY. Inhibition of glucose uptake
148. MULLER, M., W. SIEMS, F. BUTTGEREIT, R. DUMDEY, AND S. M. and glycogenolysis by availability of oleate in well oxygenated per-
RAPOPORT. Quantification of ATP-producing and consuming pro- fused skeletal muscle. Biochem. J. 168: 161-170, 1977.
cesses in Ehrlich ascites tumour cells. Eur. J. Biochem. 161: 701- 174. REYNAFARJE, B., A. ALEXANDRE, P. DAVIES, AND A. L. LEH-
705, 1986. NINGER. Proton translocation stoichiometry of cytochrome oxi-
149. MURPHY, M. P. Slip and leak in mitochondrial oxidative phosphor- dase. Proc. Natl. Acad. Sci. USA 79: 7218-7222, 1982.
ylation. Biochim. Biophys. Acta 977: 123-141, 1989. 175. ROHRACHER, H. Permanente rhythmische Mikrobewegungen des
150. MURPHY, M. P., G. C. BROWN, AND M. D. BRAND. Thermodynamic Warmbluter-Organismus (“Mikrovibration”). Naturwiss. 49: 145-
limits to the stoichiometry of H’ pumping by mitochondrial cyto- 147, 1962.
chrome oxidase. Biochem. J. 252: 473-479, 1985. 176. ROLFE, D. F. S., AND M. D. BRAND. The contribution of mitochon-
151. NAGY, K. A. Field metabolic rate and food requirement scaling in drial proton leak to the rate of respiration in rat skeletal muscle
mammals and birds. Ecol. Monogr. 57: 112- 128, 1987. and to SMR. Am. J. Physiol. 271 (Cell Physiol. 40): C1380-C1389,
152. NEWSHOLME, E. A. Sounding board: a possible basis for the con- 1996.
trol of body weight. N. Engl. J. Med. 302: 400-405, 1980. 177. ROLFE, D. F. S., AND M. D. BRAND. Proton leak and control of
153. NEWSHOLME, E. A., AND A. H. UNDERWOOD. The control of gly- oxidative phosphorylation in perfused, resting rat skeletal muscle.
colysis and gluconeogenesis in the kidney cortex. Biochem. J. 99: Biochim. Biophys. Acta 1276: 45-50, 1996.
24-26, 1966. 178. ROLFE, D. F. S., A. J. HULBERT, AND M. D. BRAND. Characteris-
154. NICHOLLS, D. G. The influence of respiration and ATP hydrolysis tics of mitochondrial proton leak and control of oxidative phos-
on the proton-electrochemical gradient across the inner membrane phorylation in the major oxygen consuming tissues of the rat. Bio-
of rat-liver mitochondria as determined by ion distribution. Eur. chim. Biophys. Acta 1188: 405-416, 1994.
J. Biochem. 50: 305-315, 1974. 179. ROTHWELL, N. J., AND M. J. STOCK. A role for brown adipose
155. NICHOLLS, D. G., AND R. M. LOCKE. Thermogenic mechanisms in tissue in diet-induced thermogenesis. Nature 281: 31-35, 1979.
brown fat. Physiol. Rev. 64: l-64, 1984. 180. ROTHWELL, N. J., AND M. J. STOCK. Diet-induced thermogenesis.
156. NOBES, C. D., G. C. BROWN, P. N. OLIVE, AND M. D. BRAND. Non- In: Mammalian Therrnogenesis, edited by L. Girardier and M. J.
ohmic conductance of mitochondrial inner membrane of hepato- Stock. London: Chapman & Hall, 1983, p. 208-233.
cytes. J. Biol. Chem. 265: 12903-12909, 1990. 181. RUDERMAN, N. B., C. R. S. HOUGHTON, AND R. HEMS. Evaluation
157. OATES, J., D. PAPAHDJOPOULOS, AND A. LOYTER. Application of the isolated perfused rat hindquarter for the study of muscle
of induced fusion to problems in biology. Trends Pharmacol. Sci. metabolism. Biochem. J. 124: 639-651, 1971.
2: 222-224, 1982. 182. SCHMIDT, H., W. SIEMS, M. MULLER, R. DUMDEY, M. JAKSTADT,

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.
758 DAVID F. S. ROLFE AND GUY C. BROWN Volume 77

AND S. M. RAPOPORT. Balancing of mitochondrial and glycolytic 203. TATA, J. R. Basal metabolic rate and thyroid hormones. In: Ad-
ATP production and of the ATP-consuming processes of Ehrlich vances in Metabolic Disorders, edited by R. Levine and R. Luft.
mouse ascites tumour cells in a high phosphate medium. Biochem. New York: Academic, 1964, vol. 1, p. 153-189.
Int. 19: 985-992, 1989. 204. TAYLOR, C. R. Structural and functional limits to oxidative metabo-
183. SCHMIDT-NIELSEN, K. Scaling: Why Is Animal Size So Im- lism: insights from scaling. Annu. Rev. Physiol. 49: 135-146, 1987.
portant? Cambridge, UK: Cambridge Univ. Press, 1984. 205. TURRENS, J. F., A. ALEXANDRE, AND A. L. LEHNINGER. Ubisemi-
184. SCHMIDT-NIELSEN, K. Animal Physiology: Adaptation and Envi- quinone is the electron donor for superoxide formation by complex
ronment. Cambridge, UK: Cambridge Univ. Press, 1990, p. 193. III of heart mitochondria. Arch. Biochem. Biophys. 237: 408-414,
185. SCHOCH, G., G. SANDER, E. FUCHS, 0. JOHREN, G. HELLER- 1985.
SCHOCH, AND H. TOPP. The whole body degradation rates of cyto- 206. UGURBIL, K., P. B. KINGSLEY-HICKMAN, E. Y. SAKO, S. ZIMMER,
plasmic tRNA, rRNA and mRNA are correlated with basal metabolic AND P. MOHANAKRISHNAN. 31P NMR studies of the kinetics and
rate in mammals of various sizes. Biol. Chern. Hoppe-Seyler 372: regulation of oxidative phosphorylation in the intact myocardium.
749, 1991. Ann. NYAcad. Sci. 508: 265-286, 1987.
186. SCHOCH, G., H. TOPP, A. HELD, G. HELLER-SCHOCH, A. BAL- 207. VERCESI, A., B. REYNAFARJE, AND A. L. LEHNINGER. Stoichiom-
LAUFF, F. MANZ, AND G. SANDER. Interrelation between whole- etry of H+ ejection and Ca2’ uptake coupled to electron transport
body turnover rates of RNA and protein. Eur. J. Clin. Nutr. 44: in rat heart mitochondria. J. Biol. Chem. 253: 6379-6395, 1978.
647-658, 1990. 208. VERHOEVEN, A. J. M., G. GORTER, M. E. MOMMERSTEEG, AND
187. SCHOLZ, R., AND T. BUCHER. Hemoglobin-free perfusion of rat J. W. M. AKKERMAN. The energetics of early platelet responses:
liver. In: Control of Energy Metabolism, edited by B. Chance. New energy-consumption during shape change and aggregation with
York: Academic, 1965, p. 393-414. special reference to protein-phosphorylation and the polyphospho-
188. SCHRODINGER, E. What Is Life? Cambridge, UK: Cambridge Univ. inositide cycle. Biochem. J. 228: 451-462, 1985.
Press, 1944. 209. VON STOCKAR, U., L. GUSTAFSSON, C. LARSSON, I. MARISON,
189. SCHUTZ, Y. The basis of direct and indirect calorimetry and their P. TISSOT, AND E. GNAIGER. Thermodynamic considerations in
potentials. Diabetes Metab. Rev. 11: 383-408, 1995. constructing energy balances for cellular growth. Biochim. Bio-
190. SELLERS, E. A., J. W. SCOTT, AND N. THOMAS. Electrical activity phys. Acta 1183: 221-240, 1993.
of skeletal muscle of normal and acclimatized rats on exposure to WATERLOW, J. C. Protein turnover with special reference to man.
cold. Am. J. Physiol. 177: 372-376, 1954. Q. J. Exp. Physiol. 69: 409-438, 1984.
191. SIEMS, W., W. DUBIEL, R. DUMDEY, M. MULLER, AND S. M. RAPO- 211. WATERLOW, J. C., P. J. GARLICK, AND D. J. MILLWARD. Protein
PORT. Accounting for the ATP-consuming processes in rabbit retic- Turnover in Mammalian Tissues and the Whole Body. New York:
ulocytes. Eur. J. Biochem. 139: lOl- 107, 1984. North-Holland, 1978.
192. SIEMS, W., M. MULLER, R. DUMDEY, H.-G. HOLZHUTTER, J. 212. WATERLOW, J. C., AND D. J. MILLWARD. Energy cost of turnover
RATHMANN, AND S. M. RAPOPORT. Quantification of pathways of protein and other cellular constituents. In: Proceedings of the
of glucose utilisation and balance of energy metabolism of rabbit Tenth Conference of the European Society for Comparative Physi-
reticulocytes. Eur. J. Biochem. 124: 567-576, 1982. ology and Biochemistry Innsbruck 1989. Stuttgart: Georg Thieme
193 SIESJO, B. K. Brain Energy Metabolism. Chichester, UK: Wiley, Verlag, 1989, p. 277-282.
1978. 213 WEBSTER, A. J. F. Prediction of energy requirements for growth
in beef cattle. World Rev. Nutr. Dietetics 30: 189-226, 1978.
194 SILVA, J. E. Hormonal control of thermogenesis and energy dissipa-
WHITTAM, R. Active cation transport as a pacemaker of respira-
tion. Trends Endocrinol. Metab. 4: 25-32, 1993.
tion. Nature 191: 603-604, 1961.
195. SKULACHEV, V. P. Nonphosphorylating respiration as a mecha-
215. WHITTAM, R. The dependence of the respiration of brain cortex
nism minimising the formation of reactive oxygen species in the
on active cation transport. Biochem. J. 82: 205-212, 1962.
cell. Mol. Biol. 29: 709-715, 1995. 216. WIESER, W. Energy allocation by addition and by compensation:
196. SMITH, R. E. Quantitative relations between liver mitochondrial an old principle revisited. In: Energy Transformations in Cells
metabolism and total body weight in mammals. Ann. NYAcad. Sci. and Organisms, edited by W. Wieser and E. Gnaiger. Stuttgart:
62: 403-422, 1956. Georg Thieme Verlag, 1989, p. 98-105.
197. SOLTOFF, S. P. ATP and the regulation of renal cell function. 217. WISNIEWSKI, E., W. S. KUNZ, AND F. N. GELLERICH. Phosphate
Annu. Rev. Physiol. 48: 9-31, 1986. affects the distribution of flux control among the enzymes of oxida-
198. SORGATO, M. C., F. GALIAZZO, L. PANATO, AND S. J. FERGUSON. tive phosphorylation in rat skeletal muscle mitochondria. J. Biol.
Estimation of H+-translocation stoichiometry of mitochondrial Chem. 268: 9343-9346, 1993.
ATPase. Biochim. Biophys. Acta 682: 184-188, 1982. 218. WOELDERS, H., W. J. VAN DER ZANDE, A.-M. COLEN, R. J. A.
199. SPRIET, L. L., C. G. MATSOS, S. J. PETERS, G. J. F. HEIGEN- WANDERS, AND K. VAN DAM. The phosphate potential maintained
HAUSER, AND N. L. JONES. Muscle metabolism and performance by mitochondria in state 4 is proportional to the protonmotive
in perfused rat hindquarter during heavy exercise. Am. J. Physiol. force. FEBS Lett. 179: 278-282, 1985.
248 (Cell Physiol. 17): Clog-C118, 1985. 219. WOLEDGE, R. C., N. A. CURTIN, AND E. HOMSHER. Energetic As-
200. SUGA, H. Ventricular energetics. Physiol. Rev. 70: 247-277, 1990. pects of Muscle Contraction. London: Academic, 1985, p. 267.
201. SWAMINATHAN, R., E. L. P. CHAN, L. Y. SIN, S. K. F. NG, AND 220. WONG-RILEY, M. T. T. Cytochrome oxidase: an endogenous meta-
A. Y. S. CHAN. The effects of ouabain on metabolic rate in guinea bolic marker for neuronal activity. Trends Neurosci. 12: 94-101,
pigs: estimation of energy costs of sodium pump activity. Br. J. 1989.
Nutr. 61: 467-473, 1989. 221. ZOLKIEWSKA, A., B. ZABLOCKA, J. DUSZYNSKI, AND L. WOJTC-
202. TANFORD, C. Equilibrium state of ATP driven ion pumps in rela- ZAK. Resting state respiration of mitochondria: reappraisal of the
tion to physiological ion concentration gradients. J. Gen. Physiol. role of passive ion fluxes. Arch. Biochem. Biophys. 275: 580-590,
77: 223-229, 1981. 1989.

Downloaded from www.physiology.org/journal/physrev by ${individualUser.givenNames} ${individualUser.surname} (018.218.056.169) on August 10, 2018.


Copyright © 1997 American Physiological Society. All rights reserved.

You might also like