Physiologists and biochemists frequently ignore the importance of adjusting equilibrium constants to the ionic conditions of the cell prior to calculating a number of bioenergetic and kinetic parameters. The present study examines the effect of pH and free magnesium levels (free [Mg2+]) on the apparent equilibrium constants (K′) of creatine kinase (ATP: creatine N-phosphotransferase; EC 2.7.3.2), adenylate kinase (ATP:AMP phosphotransferase; EC 2.7.4.3) and adenosinetriphosphatase (ATP phosphohydrolase; EC 3.6.1.3) reactions. We show how K′ can be calculated using the equilibrium constant of a specified chemical reaction (Kref) and the appropriate acid-dissociation and Mg2+-binding constants at an ionic strength (I) of 0.25 mol l−1and 38 °C. Substituting the experimentally determined intracellular pH and free [Mg2+] into the equation containing a known Kref and two variables, pH and free [Mg2+], enables K′ to be calculated at the experimental ionic conditions. Knowledge of K′ permits calculation of cytosolic phosphorylation ratio ([ATP]/[ADP][Pi]), cytosolic free [ADP], free [AMP], standard transformed Gibbs energy of formation (ΔfG′ °ATP) and the transformed Gibbs energy of the system (ΔfG′ATP) for the biological system. Such information is vital for the quantification of organ and tissue bioenergetics under physiological and pathophysiological conditions.

Knowledge of the thermodynamics of creatine kinase (EC 2.7.3.2), adenylate kinase (EC 2.7.4.3) and adenosinetriphosphatase (EC 3.6.1.3) reactions is central to studying the biochemical and physiological processes of the cell. The maintenance of near-equilibrium of the creatine kinase and adenylate kinase reactions in vivo (Lawson and Veech, 1979; Teague and Dobson, 1992; Veech et al. 1979) has led to their widespread use in estimating free cytosolic [ADP], free [AMP], cytosolic phosphorylation ratio ([ATP]/[ADP][Pi]) and [PCr]/[Pi] (where PCr is phosphocreatine and Pi is orthophosphate) ratio (Chance et al. 1985, 1986; Gyulai et al. 1985; Veech et al. 1979). The cytosolic phosphorylation ratio provides an index of the energy status of the cell. Free [ADP] and [Pi] have been implicated as the primary kinetic controllers of steady-state rates of oxygen consumption (Balaban, 1990; Chance et al. 1986; Chance and Williams, 1955; Headrick et al. 1994; Lardy and Wellman, 1952; Ugurbil et al. 1987), while free [AMP] has been shown to be involved in the regulation of a number of key glycogenolytic and glycolytic enzymes (Dobson et al. 1986; Matherne et al. 1993) and IMP levels (Matherne et al. 1993) and is possibly linked to cytosolic adenosine production (Headrick and Willis, 1990). The transformed Gibbs energy of the system (ΔfG′ATP) can be calculated from knowledge of the standard transformed Gibbs energy of formation (ΔfG′ °ATP) and from the phosphorylation ratio derived from the creatine kinase equilibrium and inorganic orthophosphate concentration. ΔfG′ATP may also be used to estimate the thermodynamic efficiency in forming 3 mol of ATP along the mitochondrial respiratory chain from NADH to O2 for every 2 electrons cycled (Dobson and Headrick, 1995).

The aim of this study is to provide quantitative mathematical expressions for the adjustment of an equilibrium constant to varying pH and free [Mg2+], thereby permitting more accurate bioenergetic assessment of mammalian organs and tissues. It is argued that such parameters have little quantitative meaning without the K′ of each reaction being adjusted to the intracellular pH, free [Mg2+], temperature (T) and ionic strength (I) of the biological system under investigation (symbols are defined in Table 1). The importance of adjustment of the apparent equilibrium constant of creatine kinase (KCK) to the pH and pMg of the cell may be illustrated in the anaesthetized rat brain. The widely used value for the KCK constant for bioenergetic calculations is 166, which is specified at pH 7.0, free [Mg2+], 1.0 mmol l−1; I, 0.25 mol l−1 and T, 38 °C. However, the adjusted constant accounting for pH and pMg of anaesthetized rat brain is 122 (pH, 7.0; free [Mg2+], 0.5 mmol l−1; I, 0.25 mol l−1; T, 38 °C). In this case, if K′ is not adjusted there would be significant errors (26 %) in the calculation of free cytosolic [ADP] and the cytosolic phosphorylation ratio ([ATP]/[ADP][Pi]). Moreover, in muscle during vigorous exercise, pH may fall by up to 1 unit (from 7.2 to 6.2) (Fitts, 1994), which would require an adjustment of KCK by nearly an order of magnitude from 85 to 726.

Table 1.

Definitions of thermodynamic quantities, symbols and units

Definitions of thermodynamic quantities, symbols and units
Definitions of thermodynamic quantities, symbols and units

Biochemical versus chemical reactions

There exists much confusion in the literature regarding the thermodynamics of a number of phosphotransferase reactions and their applicability to biological systems. For example, it is common to see the creatine kinase reaction written in the following way:
formula
where each reactant represents the sum of all the ionic species and metal complexes. The problem with the above reaction is that it does not have an equilibrium constant, because it balances neither charge nor mass. It becomes particularly confusing when an apparent equilibrium constant (Kobs or K′) follows the reaction at specified pH and pMg, temperature (T), ionic strength (I) and pressure (P). Having H+ in the reaction, as above, and giving a value for K′ at specified pH and other ionic conditions, is incompatible (Alberty, 1994a,b).

In order to clarify the situation, Alberty (1992, 1994b) has defined two equation types, a biochemical equation and a chemical equation. A biochemical equation is one describing an equilibrium mixture of total reactants followed by K′ at specified pH, pMg, I, P and T. The value of K′ may be analytically measured in the laboratory or calculated from in vitro experimental data using a system of equations as described by Teague and Dobson (1992). The point of emphasis here is that a biochemical equation deals with total concentrations and therefore does not balance charge, but it must balance elements, except for H+ and Mg2+ when pH and free [Mg2+] are specified. A chemical equation, in contrast, is defined as one comprising ionic species of reactants at specified I, P and T. By definition, a chemical reaction must balance charge and atoms of elements. Furthermore, because it is a reference equation, it may be part of a mathematical expression with the appropriate metal-binding and acid-dissociation constants with which to calculateK′ with varying pH and pMg (Teague and Dobson, 1992). The equilibrium constant of a chemical equation is thus abbreviated Kref and is dependent on I, P and T (Alberty and Goldberg, 1992). In contrast to a biochemical equation, there exist numerous chemical equations, each with specified ionic species and corresponding Kref values. It is important to realize that chemical and biochemical equations are two separate systems and cannot be added or subtracted from one another to form one equation (Alberty and Goldberg, 1992).

This distinction between biochemical and chemical equations serves to promote understanding of complex physiological processes using precise thermodynamic language. What follows is the use of such nomenclature in the adjustment of K′ with varying pH and free [Mg2+] at I=0.25 mol l−1 and T=38 °C. Computations were performed using a Macintosh computer (Microsoft Excel software).

Calculation of K′for the creatine kinase reaction

Biochemical equation:
formula
formula
where PCr is phosphocreatine, ADP is adenosine 5′-diphosphate, ATP is adenosine 5′-triphosphate, Cr is creatine and all concentrations are expressed in mol l−1. Each reactant represents the sum of all the ionic and metal complex species.
Chemical equation:
formula
formula
To be more precise, the equilibrium constant Kref (equation 5), should have the standard state concentration (c°), where c°=1.0 mol l−1, in the numerator to make the constant dimensionless, but it has been omitted in all the Kref expressions of this paper to simplify the equations (Alberty and Goldberg, 1992).
The total concentration of reactants in equation 2 are defined as:
formula
formula
formula
Equation 3 may be rearranged in terms of the ionic species and expressed as a function of the acid-dissociation constants (Ka values), magnesium-binding constants (Kb values) (see Table 2), pH and free [Mg2+] to give the following equation:
Table 2.

Acid-dissociation and magnesium-binding constants at I=0.25 mol l−1 and T=38 °C

Acid-dissociation and magnesium-binding constants at I=0.25 mol l−1 and T=38 °C
Acid-dissociation and magnesium-binding constants at I=0.25 mol l−1 and T=38 °C
formula
where Kref is 3.77 ×108 at 38 °C (Teague and Dobson, 1992) and [Mg2+] refers to the free [Mg2+]. The adjustment of KCK with varying pH and free [Mg2+] is shown in Table 3.
Table 3.

Adjustment of K′CK with varying pH and pMg for the creatine kinase equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1

Adjustment of K′CK with varying pH and pMg for the creatine kinase equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1
Adjustment of K′CK with varying pH and pMg for the creatine kinase equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1

Calculation of K′ for the adenylate kinase reaction

Biochemical equation:
formula
formula
where is the apparent equilibrium constant of adenylate kinase and all concentrations are expressed in mol l−1. Each reactant represents the sum of all the ionic and metal complex species of the reactants.
Chemical equation:
formula
formula
The total concentrations of reactants in equation 10 are defined as:
formula
formula
formula
Equation 11 may be rearranged as the sum of ionic species and expressed as a function of the acid-dissociation constants, magnesium-binding constants (Table 2), pH and free [Mg2+]:
formula
where Kref is 3.74 ×10−1 at 38 °C and was calculated from the K′ of 1.05 at pH=7.0, free [Mg2+]=1.0 mmol l−1, I=0.25 mol l−1, T=38 °C (Lawson and Veech, 1979). The adjustment of with varying pH and free [Mg2+] is given in Table 4.
Table 4.

Adjustment of K′AK with varying pH and pMg for the adenylate kinase equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1

Adjustment of K′AK with varying pH and pMg for the adenylate kinase equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1
Adjustment of K′AK with varying pH and pMg for the adenylate kinase equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1

Calculation of K′ for the ATP hydrolysis reaction

Biochemical equation:
formula
formula
where is the apparent equilibrium constant of ATP phosphohydrolase, Pi is orthophosphate and all concentrations are expressed in mol l−1. By convention, H2O concentration is unity and is omitted from equilibrium expressions. Each reactant represents the sum of all the ionic and metal complex species.
Chemical equation:
formula
formula
The total concentrations of reactants in equation 18 are defined as:
formula
formula
formula
Equation 19 may be rearranged in terms of the speciated forms and expressed as a function of the acid-dissociation constants, magnesium-binding constants (Table 2), pH and free [Mg2+]:
formula
where Kref is 7.22×10−2 at 38 °C and was calculated from the K′ of 2.19×105 mol l−1 at pH=7.0, free [Mg2+]=1.0 mmol l−1, I=0.25 mol l−1 and T=38 °C (Guynn and Veech, 1973). The adjustment of for a range of pH and free [Mg2+] values is shown in Table 5.
Table 5.

Adjustment of K′ATP with varying pH and pMg for the ATP hydrolysis equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1

Adjustment of K′ATP with varying pH and pMg for the ATP hydrolysis equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1
Adjustment of K′ATP with varying pH and pMg for the ATP hydrolysis equilibrium at T=38 °C, P=0.1 MPa and I=0.25 mol l−1

Biochemical and physiological applications of thermodynamic data

The primary aim of this study is to provide biochemists and physiologists with a number of thermodynamic expressions that will enable them to adjust K′ of key equilibria to the pH and free [Mg2+] of their experimental system at I=0.25 mol l−1 and 38 °C. Such information is essential for quantifying the bioenergetic state of a tissue or organ. The reactant concentrations ATP and PCr and the parameters pH and pMg may be obtained by phosphorus magnetic resonance spectroscopy (31P MRS) and the total creatine concentration by conventional metabolic analysis methodology (Chance et al. 1988; Conway and Radda, 1991; Gadian, 1982; Gadian and Radda, 1981; Ingwall, 1982; Kushmerick and Meyer, 1985; Meyer et al. 1982). It is important to convert the tissue enzymatic measurements from units of μmol g−1 wet mass to mol l−1, which requires estimations of the total tissue water space and the intra- and extracellular water distribution of that total water space for the organ system under investigation (Dobson et al. 1992; Masuda et al. 1990).

Free cytosolic [ADP]

Free cytosolic [ADP] (in mol l−1) can be calculated from rearrangement of the equilibrium expression (equation 3):
formula
Before calculating [ADP], the apparent equilibrium constant, , needs to be adjusted to the pH and free [Mg2+] of the experimental conditions at I=0.25 mol l−1 and 38 °C using equation 9. The [ADP] calculated this way is often called the free cytosolic [ADP], which has been shown to be 20- to 50-fold less than the total measured tissue content (Bünger and Soboll, 1986; Seraydarian et al. 1962; Veech et al. 1979).

Cytosolic phosphorylation ratio

The phosphorylation ratio is often referred to as the phosphorylation potential, but it is not a potential with units of energy, but rather a ratio of the reactants of the ATP hydrolysis reaction in units of concentration (Slater, 1976). The phosphorylation ratio (in l mol−1) is calculated from an arrangement of the creatine kinase equilibrium, and the free cytosolic orthophosphate concentration, [Pi], which must be determined independently:
formula
Before calculating the phosphorylation ratio, needs to be adjusted to the pH and free [Mg2+] of the experimental conditions at I=0.25 mol l−1 and 38 °C using equation 9.

Free cytosolic [AMP]

Free cytosolic [AMP] (in mol l−1) can be calculated from rearrangement of the adenylate equilibrium expression (equation 11):
formula
Prior to calculation of free [AMP], needs to be adjusted to the pH and free [Mg2+] of the experimental conditions at I=0.25 mol l−1 and 38 °C using equation 17. The [AMP] calculated in this way is often called the free cytosolic [AMP], which has been shown to be 20- to 50-fold less than the total measured tissue content (Bünger and Soboll, 1986).

Calculation of the ΔfGof ATP hydrolysis: relevance to biological systems

Since ATP is the primary energy currency of a cell, it is the chemical potential of its hydrolysis (equations 18, 19), as opposed to its synthesis, that drives the extent and direction of the energy transformations in living systems (Krebs and Kornberg, 1957). The transformed Gibbs energy of ATP hydrolysis, LfG′ATP, can be determined from the following equation:
formula
where R is the gas constant, T is the temperature (in Kelvin; see Table 1) and ΔfG′ °ATP is the standard transformed Gibbs energy of ATP hydrolysis (ATP + H2O = ADP + Pi) at a specified pH, free [Mg2+], I, P and T (see below). The cytosolic phosphorylation ratio, [ATP]/([ADP][Pi]), is calculated from the creatine kinase equilibrium expression (equation 27). It should be noted that the phosphorus metabolite values in equations 26–29 represent their free concentrations as determined by 31P magnetic resonance spectroscopy, rather than total tissue measurements.

Calculation of the ΔfG′ ° of ATP hydrolysis with varying free [Mg2+] and pH at I=0.25 mol l-1and T=38 °C

The standard apparent Gibbs energy of ATP hydrolysis (ATP + H2O = ADP + Pi) provides a quantitative measure of the chemical potential for phosphate group transfer between the reactants ATP, ADP and Pi for specifed ionic conditions, I, P and T. This is in contrast to the ATP ‘high-energy phosphate’ concept, which refers to bond energy and not to the free energy difference between the reacting components of a specified reaction (Lipmann, 1941). ΔfG′ °ATP is calculated by equating the apparent transformed Gibbs energy (ΔfG′) to zero and solving using the following equation:
formula
where KATP is the apparent equilibrium constant of the ATP hydrolysis reaction (equation 19), R is the gas constant and T is the temperature (in Kelvin; Table 1). LfG′°ATP can also be related to the standard transformed enthalpy and standard transformed entropy of reaction where ΔfG′° = ΔfH′° − TΔfS′ ° (symbols defined in Table 1).

The present study has provided mathematical expressions for calculating the apparent equilibrium constant (K′) of the creatine kinase, adenylate kinase and ATP hydrolysis reactions in terms of Kref and the appropriate acid-dissociation and magnesium-binding constants. We have calculated Kref for each reaction and demonstrated how K′ can be adjusted to varying levels of experimental pH and free [Mg2+]. Tables of K′ as a function of pH and free [Mg2+] at I=0.25 mol l−1 and T=38 °C have also been provided for convenience. Finally, we have indicated some of the biochemical applications for using the equilibrium constants in assessing cellular bioenergetics taking place under physiological and pathophysiological conditions.

The authors would like to thank R. A. Alberty, Professor Emeritus, Department of Chemistry, Massachusetts Institute of Technology, for critical comments on the manuscript. We thank members of the Department of Molecular Sciences, Division of Chemistry (Dr P. Duggan) and Division of Biochemistry, James Cook University, for their interest and lively discussions on some aspects of this study. This work was supported by an ARC small grant 1.91380.9821 (G.P.D.).

Alberty
,
R. A.
(
1992
).
Equilibrium calculations on systems of biochemical reactions at specified pH and pMg
.
Biophys. Chem.
42
,
117
131
.
Alberty
,
R. A.
(
1994a
).
Biochemical thermodynamics
.
Biochim. biophys. Acta
1207
,
1
11
.
Alberty
,
R. A.
(
1994b
).
Recommendations for nomenclature and tables in biochemical thermodynamics
.
Pure appl. Chem.
66
,
1641
1666
.
Alberty
,
R. A.
and
Goldberg
,
R. N.
(
1992
).
Standard thermodynamic formation properties for the adenosine 5′-triphosphate series
.
Biochemistry, N.Y
.
31
,
10610
10615
.
Balaban
,
R. S.
(
1990
).
Regulation of oxidative phosphorylation in the mammalian cell
.
Am. J. Physiol.
258
,
C377
C389
.
Bünger
,
R.
and
Soboll
,
S.
(
1986
).
Cytosolic adenylates and adenosine release in perfused working heart: Comparison of whole tissue with cytosolic non-aqueous fractionation analyses
.
Eur. J. Biochem.
159
,
203
213
.
Chance
,
B.
,
Leigh
,
J. S.
,
Clark
,
B. J.
,
Maris
,
J.
,
Kent
,
J.
,
Nioka
,
S.
and
Smith
,
D.
(
1985
).
Control of oxidative metabolism and oxygen delivery in human skeletal muscle: A steady-state analysis of the work/energy cost transfer function
.
Proc. natn. Acad. Sci. U.S.A
.
82
,
8384
8388
.
Chance
,
B.
,
Leigh
,
J. S.
,
Mccully
,
K.
,
Nioka
,
S.
,
Clark
,
B. J.
,
Maris
,
J. M.
and
Graham
,
T.
(
1986
).
Multiple controls of oxidative metabolism in living tissues as studied by phosphorus magnetic resonance
.
Proc. natn. Acad. Sci. U.S.A
.
83
,
9458
9462
.
Chance
,
B.
,
Leigh
,
J. S.
,
Mclaughlin
,
A. C.
,
Schnall
,
M.
and
Sinnwell
,
T.
(
1988
).
Phosphorus-31 spectroscopy and imaging
. In
Magnetic Resonance Imaging, vol. 2
(ed.
C. L.
Partain
,
R. R.
Price
,
J. A.
Patton
,
M. V.
Kulkarni
and
A. E.
James
), pp.
1501
1520
.
New York
:
W. B. Saunders Co
.
Chance
,
B.
and
Williams
,
G. R.
(
1955
).
Respiratory enzymes in oxidative phosphorylation. I. Kinetics of oxygen utilization
.
J. biol. Chem.
217
,
383
393
.
Conway
,
M. A.
and
Radda
,
G. K.
(
1991
).
Nuclear magnetic resonance spectroscopic investigations of the human myocardium
.
Topics cardiovasc. Med
.
1
,
300
304
.
Dobson
,
G. P.
and
Headrick
,
J. P.
(
1995
).
Bioenergetic scaling: metabolic design and body-size constraints in mammals
.
Proc. natn. Acad. Sci. U.S.A. (in press)
.
Dobson
,
G. P.
,
Veech
,
R. L.
,
Passonneau
,
J. V.
,
Kobayashi
,
K.
,
Inubushi
,
T.
,
Wehrli
,
S.
,
Nioka
,
S.
and
Chance
,
B.
(
1992
).
31P NMR and enzymatic analysis of cytosolic phosphocreatine, ATP, Pi, intracellular pH in the isolated working rat heart
.
NMR in Biomedicine
5
,
20
28
.
Dobson
,
G. P.
,
Yamamoto
,
E.
and
Hochachka
,
P. W.
(
1986
).
Phosphofructokinase control in muscle: nature and reversal of pH-dependent ATP inhibition
.
Am. J. Physiol.
250
,
R71
R76
.
Fitts
,
R. H.
(
1994
).
Cellular mechanisms of muscle fatigue
.
Physiol. Rev.
74
,
49
94
.
Gadian
,
D. G.
(
1982
).
Nuclear Magnetic Resonance and its Applications to Living Systems
, pp.
117
118
.
New York
:
Oxford University Press
.
Gadian
,
D. G.
and
Radda
,
G. K.
(
1981
).
NMR studies of tissue metabolism
.
A. Rev. Biochem.
50
,
69
83
.
Guynn
,
R. W.
and
Veech
,
R. L.
(
1973
).
The equilibrium constants of the adenosine triphosphate hydrolysis and the adenosine triphosphate–citrate lyase reactions
.
J. biol. Chem.
248
,
6966
6972
.
Gyulai
,
L.
,
Roth
,
Z.
,
Leigh
,
J. S. J.
and
Chance
,
B.
(
1985
).
Bioenergetic studies of mitochondrial oxidative phosphorylation using 31phosphorus NMR
.
J. biol. Chem.
260
,
3947
3954
.
Headrick
,
J. P.
,
Dobson
,
G. P.
,
Williams
,
J. P.
,
Mckirdy
,
J. C.
,
Jordan
,
L. R.
and
Willis
,
R. J.
(
1994
).
Bioenergetics and control of oxygen consumption in the in situ rat heart
.
Am. J. Physiol.
267
,
H1074
H1084
.
Headrick
,
J. P.
and
Willis
,
R. J.
(
1990
).
Adenosine formation and energy metabolism: A 31P NMR study in isolated rat heart
.
Am. J. Physiol.
258
,
H617
H624
.
Ingwall
,
J. S.
(
1982
).
Phosphorus nuclear magnetic resonance spectroscopy of cardiac and skeletal muscles
.
Am. J. Physiol.
242
,
H729
H744
.
Krebs
,
H. A.
and
Kornberg
,
H. L.
(
1957
).
Energy Transformations in Living Matter
.
Berlin
:
Springer-Verlag
.
Kushmerick
,
M. J.
and
Meyer
,
R. A.
(
1985
).
Chemical changes in rat leg muscle by phosphorus nuclear magnetic resonance
.
Am. J. Physiol.
248
,
C542
C549
.
Lardy
,
H. A.
and
Wellman
,
H. Y.
(
1952
).
Oxidative phosphorylation: role of inorganic phosphate and acceptor systems in control of metabolic rates
.
J. biol. Chem.
195
,
215
224
.
Lawson
,
J. W. R.
and
Veech
,
R. L.
(
1979
).
Effect of pH and free Mg2+ on the Keq of the creatine kinase reaction and other phosphate hydrolysis and phosphate transfer reactions
.
J. biol. Chem.
254
,
6528
6537
.
Lipmann
,
F.
(
1941
).
Metabolic generation and utilization of phosphate bond energy
. In
Advances in Enzymology and Related Subjects, vol. 1
(ed.
F. F.
Nord
and
C. H. W.
Nord
), pp.
99
121
.
New York
:
Academic Press
.
Masuda
,
T.
,
Dobson
,
G. P.
and
Veech
,
R. L.
(
1990
).
The Gibbs–Donnan near-equilibrium system of heart
.
J. biol. Chem.
265
,
20321
20334
.
Matherne
,
G. P.
,
Headrick
,
J. P.
,
Berr
,
S.
and
Berne
,
R. M.
(
1993
).
Metabolic and functional responses of immature and mature rabbit hearts to hypoperfusion, ischemia and reperfusion
.
Am. J. Physiol.
264
,
H2143
H2153
.
Meyer
,
R. A.
,
Kushmerick
,
M. A.
and
Brown
,
T. R.
(
1982
).
Application of 31P NMR spectroscopy to the study of striated muscle metabolism
.
Am. J. Physiol.
242
,
C1
C11
.
Seraydarian
,
K.
,
Mommaerts
,
W. F. H. M.
and
Wallner
,
A.
(
1962
).
The amount and compartmentalization of adenosine diphosphate in muscle
.
Biochim. biophys. Acta
65
,
443
460
.
Slater
,
E. C.
(
1976
).
Intra- and extramitochondrial phosphorylation potentials
. In
Use of Isolated Liver Cells and Kidney Tubules in Metabolic Studies
(ed.
J. M.
Tager
,
H. D.
Soling
and
J. R.
Williamson
).
Amsterdam
:
North-Holland Publishing Company
.
Teague
,
W. E. J.
and
Dobson
,
G. P.
(
1992
).
Effect of temperature on the creatine kinase equilibrium
.
J. biol. Chem.
267
,
14084
14093
.
Ugurbil
,
K.
,
Kingsley-Hickman
,
P. B.
,
Sako
,
E. Y.
,
Zimmer
,
S.
,
Mohanakrishnan
,
P.
,
Robitaille
,
P. M. L.
,
Thoma
,
W. J.
,
Johnson
,
A.
,
Foker
,
J. E.
and
From
,
A. H. L.
(
1987
).
31P NMR studies of the kinetics and regulation of oxidative phosphorylation in the intact myocardium
. In
Annals of the New York Academy of Sciences. Physiological NMR spectroscopy, vol. 508
(ed.
S. M.
Cohen
), pp.
265
286
.
New York
:
The New York Academy of Sciences
.
Veech
,
R. L.
,
Lawson
,
J. W. R.
,
Cornell
,
N. W.
and
Krebs
,
H. A.
(
1979
).
Cytosolic phosphorylation potential
.
J. biol. Chem.
254
,
6538
6547
.