Comprehensive Physiology Wiley Online Library

Principles of Regulation and Control in Biochemistry: A Pragmatic, Flux‐Oriented Approach

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Flux‐Oriented Theory of Regulation
1.1 Flux‐Generating Steps
1.2 Branched Fluxes: Reversible Reactions
1.3 Other Types of Flux: Energy Fluxes and Equivalence
1.4 Internal and External Effectors
1.5 Control and Regulation of Metabolic Fluxes and Concentrations: Communication and Regulatory Sequences
1.6 Control Versus Regulation
1.7 Identification of Communication Sequences In Vivo
1.8 Need for Sensitive Regulation In Vivo
1.9 Metabolic Sensitivities
1.10 Analysis of Non‐Steady States
2 Two Other Important Approaches to Metabolic Control and Regulation
2.1 Biochemical Systems Theory (BST)
2.2 Metabolic Control Theory (MCT)
3 Summary and Future Prospects
Figure 1. Figure 1.

Simple linear flux. A‐D represent metabolic intermediates, E1‐E4 represent enzyme‐catalyzed reactions and symbol denotes saturation (here, with A).

Figure 2. Figure 2.

Metabolic carbohydrate flux, (a) Conventional pathway of glycolysis, originating with glucose entry into a cell, (b) whole‐body flux, of which (a) is a component. E1 = glycogen phosphorylase. In this, and other figures the cofactors are omitted, unless relevant to the discussions.

Figure 3. Figure 3.

Branched flux.

Figure 4. Figure 4.

Fluxes with reversible steps. (a) Hypothetical linear flux, (b) glycolysis: a = hexokinase (HK); b = phosphoglucose isomerase (PGI); c = phosphofructokinase (PFK); d = pyruvate kinase (PK); e = lactate dehydrogenase (LDH); f = pyruvate dehydrogenase (PDH).

Figure 5. Figure 5.

Energy fluxes, (a) Equivalent fluxes; JE=2xJl, (b) non‐equivalent fluxes; total energy flux = 2 × Jl + 38 × J0, which is not proportional to, and thus not equivalent to, the total carbon flux, Jo+Jl.

Figure 6. Figure 6.

Linear flux with external effectors. The broken lines denote regulatory (for example, allosteric) interactions.

Figure 7. Figure 7.

Partially external regulators. The concentrations of metabolites M and N are only partially determined by flux J. On the other hand, they are totally determined by the wider system, J + K. Regulator X (totally external for J) communicates with M, but M (partially external for J) does not communicate with X.

Figure 8. Figure 8.

Truncated flux. The reactions producing B are omitted, so that B is a partially external regulator of the truncated flux, Jt.

Figure 9. Figure 9.

Flux with feedback inhibition. (a) Simple linear system, (b) system with extended feedback (end product) inhibition. ⊖ denotes an inhibition, ⊕ denotes an activation.

Figure 10. Figure 10.

Outline of the control of glycolysis by ATP. E1 = phosphorylase b; E2 = phosphofructokinase (PFK). ⊕ denotes an activation.

Figure 11. Figure 11.

System with two potential control sites, only one of which is used. Here, as in subsequent figures, the greek letters denote relative‐change intrinsic or component sensitivities.

Figure 12. Figure 12.

Hill responses. The curves are calculated from equation 35, with K = 1 and Vm = 100.

Figure 13. Figure 13.

Reversible system spanning a membrane. XI and XO denote concentrations inside and outside a cell compartment, respectively. M is a metabolite present only on the inside.

Figure 14. Figure 14.

Substrate cycles. (a) Simple hypothetical cycle, (b) cycle between fructose‐6‐phosphate (F‐6‐P) and fructose‐1,6‐bisphosphate (FBP), (c) cycle between triacylglycerol and FFA; the 7 ATP molecules consumed per turn comprise two each for the conversion of the three molecules of FFA to their CoA esters, and one for the formation of glycerol phosphate.

Figure 15. Figure 15.

Interconversion cycle. Enzyme E exists in two forms, a and b, which are interconverted by E1 and E2. The energy consumption here is represented by a net hydrolysis of ATP to ADP and Pi. In the steady state the rates of E1 and E2 are equal, and there is no net flux across the cycle.

Figure 16. Figure 16.

Regulation of pyruvate dehydrogenase (PDH) by pyruvate. Pyruvate has a direct effect (rm + rc) and an indirect one, via inhibition of the phosphatase component of the interconversion cycle.

Figure 17. Figure 17.

Reversible reactions in situ. (a) Simple hypothetical system, (b) flux of serine biosynthesis in vivo. E1 = phosphoglycerate dehydrogenase plus phosphoserine aminotransferase; E2 = phosphoserine phosphatase. The symbol, denotes saturation.

Figure 18. Figure 18.

Basic control of glycolysis by ATP. This diagram illustrates the general principles: the presence of auxiliary regulators and other mechanisms makes the actual system much more complicated.

Figure 19. Figure 19.

Linear flux with external regulator of E2.

Figure 20. Figure 20.

Linear flux with reversible reaction at E2.

Figure 21. Figure 21.

Linear flux with partially external regulator at E2.

Figure 22. Figure 22.

Opening the system in Figure 21.

Figure 23. Figure 23.

Branched flux with two external regulators.

Figure 24. Figure 24.

The pathway of glutaminolysis.

Figure 25. Figure 25.

Relationship of serine biosynthesis to glycolysis.

Figure 26. Figure 26.

Linear flux with a regulator interacting at two sites.

Figure 27. Figure 27.

Linear non‐steady state system. Here v1, v2 and v3 are not mutually equal, so that S and P vary with time and there is no defined system flux; α‐γ denote absolute‐change sensitivities.

Figure 28. Figure 28.

Effect of the speed of response on a pulsed system. The pulsed stimulus lasts for time τ. Curves 1–4 show the effect of a coordinated increase of the initial rate of cycling between the b and a forms of an interconversion cycle. In curve 4 a steady state has been reached within time τ and any further increase in the rate of initial cycling has no further effect.

Figure 29. Figure 29.

Metabolic buffers. (a) Equilibrium. Q = Keq.S and the resulting larger pool of S is shown in parentheses. Fluxes I and O are equal in the steady state. (b) Receptor recycling. The hormone binds tightly to its receptor, but the complex is destroyed when it is internalized (via F); the free receptor is then recycled (via C). Here, fluxes I and O represent the delivery and removal of the hormone, respectively. (c) Substrate cycle. Here, in the steady state there is no net flux across the cycle, that fluxes I and O are equal. S and P may denote (1) glucose and glycogen, (2) amino acids and protein, (3) fatty acids and triacylglycerol, in which case the cycles may be regarded as storage cycles. E1 and E2 are assumed to have first‐order responses to S and P, respectively.

Figure 30. Figure 30.

Oscillations. Xs is the steady state value, or balance point around which X oscillates. The broken line shows a fading, or damped, oscillation.

Figure 31. Figure 31.

System with indirect removal of an intermediate. The indirect communication, S∼M∼E2, introduces a delay between the production of S (by E1) and its removal (by E2).

Figure 32. Figure 32.

System with a long feedback loop.

Figure 33. Figure 33.

A branched system in BST formulation.

Figure 34. Figure 34.

Linear flux in MCT formulation. A comparison with Figure 11 shows that ε 1/s = α; ε 2/s = β.

Figure 35. Figure 35.

Interconversion cycle with a single molecule of protein. a, b, and c refer to the same molecule and hence cannot coexist. k1 and k2 are classical rate constants that would give the observed rates as a function of the fractional times spent as a or b. The conversion of a into c takes a time π.

Figure 36. Figure 36.

Flux‐generating step with a particulate system. G is a glycogen particle, to which phosphorylase and other enzymes (E) are tightly bound. E does not dissociate as glucose‐1‐phosphate is released.

Figure 37. Figure 37.

Switch mechanism at an interconversion cycle. (a) Cycle, (b) response of E1 to b: substrate inhibition, (c) response of active, a form to regulator X. At X2 there is a large increase of a, which is not reversed until X is lowered to X1, a value much lower than X2. This system before shows hysteresis and enables a response to be maintained in the absence of the original stimulus 36.

Figure 38. Figure 38.

A fully reversible system. The net flux J (=AB; =FC; =DE) can reverse and hence change sign.

Figure 39. Figure 39.

Evaluating core sensitivities by graph theory, (a) Digraph for the system in Figure 11 (omitting Z). Variables J and S are nodes, (b) directed circuits and connections for the digraph, (c) directed paths and one‐connections for each variable. For farther details see text.



Figure 1.

Simple linear flux. A‐D represent metabolic intermediates, E1‐E4 represent enzyme‐catalyzed reactions and symbol denotes saturation (here, with A).



Figure 2.

Metabolic carbohydrate flux, (a) Conventional pathway of glycolysis, originating with glucose entry into a cell, (b) whole‐body flux, of which (a) is a component. E1 = glycogen phosphorylase. In this, and other figures the cofactors are omitted, unless relevant to the discussions.



Figure 3.

Branched flux.



Figure 4.

Fluxes with reversible steps. (a) Hypothetical linear flux, (b) glycolysis: a = hexokinase (HK); b = phosphoglucose isomerase (PGI); c = phosphofructokinase (PFK); d = pyruvate kinase (PK); e = lactate dehydrogenase (LDH); f = pyruvate dehydrogenase (PDH).



Figure 5.

Energy fluxes, (a) Equivalent fluxes; JE=2xJl, (b) non‐equivalent fluxes; total energy flux = 2 × Jl + 38 × J0, which is not proportional to, and thus not equivalent to, the total carbon flux, Jo+Jl.



Figure 6.

Linear flux with external effectors. The broken lines denote regulatory (for example, allosteric) interactions.



Figure 7.

Partially external regulators. The concentrations of metabolites M and N are only partially determined by flux J. On the other hand, they are totally determined by the wider system, J + K. Regulator X (totally external for J) communicates with M, but M (partially external for J) does not communicate with X.



Figure 8.

Truncated flux. The reactions producing B are omitted, so that B is a partially external regulator of the truncated flux, Jt.



Figure 9.

Flux with feedback inhibition. (a) Simple linear system, (b) system with extended feedback (end product) inhibition. ⊖ denotes an inhibition, ⊕ denotes an activation.



Figure 10.

Outline of the control of glycolysis by ATP. E1 = phosphorylase b; E2 = phosphofructokinase (PFK). ⊕ denotes an activation.



Figure 11.

System with two potential control sites, only one of which is used. Here, as in subsequent figures, the greek letters denote relative‐change intrinsic or component sensitivities.



Figure 12.

Hill responses. The curves are calculated from equation 35, with K = 1 and Vm = 100.



Figure 13.

Reversible system spanning a membrane. XI and XO denote concentrations inside and outside a cell compartment, respectively. M is a metabolite present only on the inside.



Figure 14.

Substrate cycles. (a) Simple hypothetical cycle, (b) cycle between fructose‐6‐phosphate (F‐6‐P) and fructose‐1,6‐bisphosphate (FBP), (c) cycle between triacylglycerol and FFA; the 7 ATP molecules consumed per turn comprise two each for the conversion of the three molecules of FFA to their CoA esters, and one for the formation of glycerol phosphate.



Figure 15.

Interconversion cycle. Enzyme E exists in two forms, a and b, which are interconverted by E1 and E2. The energy consumption here is represented by a net hydrolysis of ATP to ADP and Pi. In the steady state the rates of E1 and E2 are equal, and there is no net flux across the cycle.



Figure 16.

Regulation of pyruvate dehydrogenase (PDH) by pyruvate. Pyruvate has a direct effect (rm + rc) and an indirect one, via inhibition of the phosphatase component of the interconversion cycle.



Figure 17.

Reversible reactions in situ. (a) Simple hypothetical system, (b) flux of serine biosynthesis in vivo. E1 = phosphoglycerate dehydrogenase plus phosphoserine aminotransferase; E2 = phosphoserine phosphatase. The symbol, denotes saturation.



Figure 18.

Basic control of glycolysis by ATP. This diagram illustrates the general principles: the presence of auxiliary regulators and other mechanisms makes the actual system much more complicated.



Figure 19.

Linear flux with external regulator of E2.



Figure 20.

Linear flux with reversible reaction at E2.



Figure 21.

Linear flux with partially external regulator at E2.



Figure 22.

Opening the system in Figure 21.



Figure 23.

Branched flux with two external regulators.



Figure 24.

The pathway of glutaminolysis.



Figure 25.

Relationship of serine biosynthesis to glycolysis.



Figure 26.

Linear flux with a regulator interacting at two sites.



Figure 27.

Linear non‐steady state system. Here v1, v2 and v3 are not mutually equal, so that S and P vary with time and there is no defined system flux; α‐γ denote absolute‐change sensitivities.



Figure 28.

Effect of the speed of response on a pulsed system. The pulsed stimulus lasts for time τ. Curves 1–4 show the effect of a coordinated increase of the initial rate of cycling between the b and a forms of an interconversion cycle. In curve 4 a steady state has been reached within time τ and any further increase in the rate of initial cycling has no further effect.



Figure 29.

Metabolic buffers. (a) Equilibrium. Q = Keq.S and the resulting larger pool of S is shown in parentheses. Fluxes I and O are equal in the steady state. (b) Receptor recycling. The hormone binds tightly to its receptor, but the complex is destroyed when it is internalized (via F); the free receptor is then recycled (via C). Here, fluxes I and O represent the delivery and removal of the hormone, respectively. (c) Substrate cycle. Here, in the steady state there is no net flux across the cycle, that fluxes I and O are equal. S and P may denote (1) glucose and glycogen, (2) amino acids and protein, (3) fatty acids and triacylglycerol, in which case the cycles may be regarded as storage cycles. E1 and E2 are assumed to have first‐order responses to S and P, respectively.



Figure 30.

Oscillations. Xs is the steady state value, or balance point around which X oscillates. The broken line shows a fading, or damped, oscillation.



Figure 31.

System with indirect removal of an intermediate. The indirect communication, S∼M∼E2, introduces a delay between the production of S (by E1) and its removal (by E2).



Figure 32.

System with a long feedback loop.



Figure 33.

A branched system in BST formulation.



Figure 34.

Linear flux in MCT formulation. A comparison with Figure 11 shows that ε 1/s = α; ε 2/s = β.



Figure 35.

Interconversion cycle with a single molecule of protein. a, b, and c refer to the same molecule and hence cannot coexist. k1 and k2 are classical rate constants that would give the observed rates as a function of the fractional times spent as a or b. The conversion of a into c takes a time π.



Figure 36.

Flux‐generating step with a particulate system. G is a glycogen particle, to which phosphorylase and other enzymes (E) are tightly bound. E does not dissociate as glucose‐1‐phosphate is released.



Figure 37.

Switch mechanism at an interconversion cycle. (a) Cycle, (b) response of E1 to b: substrate inhibition, (c) response of active, a form to regulator X. At X2 there is a large increase of a, which is not reversed until X is lowered to X1, a value much lower than X2. This system before shows hysteresis and enables a response to be maintained in the absence of the original stimulus 36.



Figure 38.

A fully reversible system. The net flux J (=AB; =FC; =DE) can reverse and hence change sign.



Figure 39.

Evaluating core sensitivities by graph theory, (a) Digraph for the system in Figure 11 (omitting Z). Variables J and S are nodes, (b) directed circuits and connections for the digraph, (c) directed paths and one‐connections for each variable. For farther details see text.

References
 1. Akerboom, T. P. M., R. Van Der Meer, and J. M. Tager. Techniques for the investigation of intracellular compartmentation. Tech. Metabol. Res. B205: 1–33, 1979.
 2. Atkinson, D. E. Regulation of enzyme activity. Annu. Rev. Biochem. 35: 85–124, 1966.
 3. Atkinson, D. E. Limitation of metabolite concentrations and the conservation of solvent capacity in the living cell. Curr. Top. Cell. Regul. 1: 29–43, 1969.
 4. Axelrod, J. Receptor‐mediated activation of phospholipase A2 and arachidonic acid release in signal transduction. Biochem. Soc. Trans. 18: 503–507, 1990.
 5. Bahr, R., P. Hansson, and O. M. Sejersted. Triglyceride/fatty acid cycling is increased after exercise. Metabolism 39: 993–999, 1990.
 6. Batke, J. Channeling of glycolytic intermediates by temporary, stationary bi‐enzyme complexes is probable in vivo. Trends Biochem. Sci. 14: 481–482, 1989.
 7. Beck, J. S. On internalization of hormone‐receptor complex and receptor cycling. J. Theor. Biol. 132: 263–276, 1988.
 8. Beeckmans, S. Some structural and regulatory aspects of citrate synthase. Int. J. Biochem. 16: 341–351, 1984.
 9. Berry, M. N., R. B. Gregory, A. R. Grivell, D. C. Henley, J. W. Phillips, P. G. Wallace, and G. R. Welch. Linear relationships between mitochondrial forces and cytoplasmic flows argue for the organized energy‐coupled nature of cellular metabolism. FEBS Lett. 224: 201–207, 1987.
 10. Board, M., S. Humm, and E. A. Newsholme. Maximum activities of key enzymes of glycolysis, glutaminolysis, pentose phosphate pathway and tricarboxylic acid cycle in normal, neoplastic and suppressed cells. Biochem J. 265: 503–509, 1990.
 11. Bowman, R. H. Inhibition of citrate metabolism by sodium fluoroacetate in the perfused rat heart and the effect on phosphofructokinase activity and glucose utilization. Biochem. J. 93: 13C–15C, 1964.
 12. Bristow, J., D. M. Bier, and L. G. Lange. Regulation of adult and fetal myocardial phosphofructokinase. Relief of cooperativity and competition between fructose‐2,6‐bisphosphate, ATP and citrate. J. Biol. Chem. 262: 2171–2175, 1987.
 13. Brown, G. C, R. P. Hafner, and M. D. Brand. A ‘top‐down’ approach to the determination of control coefficients in metabolic control theory. Eur. J. Biochem. 188: 321–325, 1990.
 14. Bucher, T. and W. Russmann. Equilibrium and nonequilibrium in the glycolysis system. Angew. Chem. Internat. Edit. 3: 426–439, 1964.
 15. Burgoyne, R. D. and A. Morgan. The control of free arachidonic acid levels. Trends Biochem. Sci. 15: 364–366, 1990.
 16. Burnell, J. N. and M. D. Hatch. Activation and inactivation of an enzyme catalyzed by a single, bifunctional protein: a new example and why. Arch. Biochem. Biophys. 245: 297–304, 1986.
 17. Burt, C. T., T. Glonek, and M. Barany. Analysis of phosphate metabolites, the intracellular pH, and the state of adenosine triphosphate in intact muscle by phosphorus nuclear magnetic resonance. J. Biol. Chem. 251: 2584–2591, 1976.
 18. Cardenas, M. L. and A. Cornish‐Bowden. Characteristics necessary for an interconvertible enzyme cascade to generate a highly sensitive response to an effector. Biochem. J. 257: 339–345, 1989.
 19. Cascante, M., E. I. Canela, and R. Franco. Control analysis of systems having two steps catalysed by the same protein molecule in unbranched chains. Eur. J. Biochem. 192: 369–371, 1990.
 20. Cascante, M., R. Franco, and E. I. Canela. Use of implicit methods from general sensitivity theory to develop a systematic approach to metabolic control. II. complex systems. Math. Biosci. 94: 289–309, 1989.
 21. Cascante, M., R. Franco, and E. I. Canela. Use of implicit methods from general sensitivity theory to develop a systematic approach to metabolic control. I. Unbranched pathways. Math. Biosci. 94: 271–288, 1989.
 22. Challis, R. A. J., B. Crabtree, and E. A. Newsholme. Hormonal regulation of the rate of the glycogen/glucose‐1‐phosphate cycle in skeletal muscle. Eur. J. Biochem. 163: 205–210 1987.
 23. Challiss, R.J.A., J.R.S. Arch, and E. A. Newsholme. The rate of substrate cycling between fructose‐6‐phosphate and fructose‐1,6‐bisphosphate in skeletal muscle. Biochem. J. 221: 153–161, 1984.
 24. Chen, Y‐D. and H. V. Westerhoff. How do inhibitors and modifiers of individual enzymes affect steady state fluxes and concentrations in metabolic systems? Math. Modelling 7: 1173–1180, 1986.
 25. Clark, M. G., C. H. Williams, W. F. Pfeifer, D. P. Bloxham, P. C. Holland, C. A. Taylor, and H. A. Lardy. Accelerated substrate cycling of fructose‐6‐phosphate in the muscles of malignant hyperthermic pigs. Nature 245: 99–101, 1973.
 26. Cleland, W. W. The kinetics of enzyme‐catalysed reactions with two or more substrates and products. III. Prediction of initial velocity and inhibition patterns by inspection. Biochim. Biophys. Acta 67: 188–196, 1963.
 27. Cohen, P. The role of protein phosphorylation in the hormonal control of enzyme activity. Eur. J. Biochem. 151: 439–448, 1985.
 28. Cohen, S. M. Simultaneous 13C and 31P NMR studies of perfused rat liver: effects of insulin and glucagon and a 13C NMR assay of free Mg2+. J. Biol. Chem. 258: 14294–14308, 1983.
 29. Cohen, S. M., R. G. Shulman, and A. C. McLaughlin. Effects of ethanol on alanine metabolism in perfused mouse liver studied by 13C NMR. Proc. Natl. Acad. Sci. USA 76: 4808–4812, 1979.
 30. Cohen, S. R. Why does brain make lactate? J. Theor. Biol. 112: 429–432, 1985.
 31. Cornish‐Bowden, A. Metabolic control theory and biochemical systems theory: different objectives, different assumptions, different results. J. Theor. Biol. 136: 365–377, 1989.
 32. Cornish‐Bowden, A. Failure of channelling to maintain low concentrations of metabolic intermediates. Eur. J. Biochem. 195: 103–108, 1991.
 33. Cox, D. R. and H. D. Miller. The Theory of Stochastic Processes. London: Methuen, 1965.
 34. Crabtree, B. Reversible (near equilibrium) reactions and substrate cycles. Biochem. Soc. Trans. 4: 1046–1048, 1976.
 35. Crabtree, B. Theoretical considerations of the sensitivity conferred by substrate cycles in vivo. Biochem. Soc. Trans. 4: 999–1002, 1976.
 36. Crabtree, B. A metabolic switch produced by enzymically interconvertible forms of an enzyme. FEBS Lett. 187: 193–195, 1985.
 37. Crabtree, B., Metabolic regulation. In: Quantitative Aspects of Ruminant Digestion and Metabolism, edited by J. M. Forbes and J. France. Butterworths, 1992.
 38. Crabtree, B. A qualitative method for identifying external control sites in metabolic systems. Am. J. Physiol. 262 (Regulatory Integrative Comp Physiol. 31): R806–R812.
 39. Crabtree, B., G. Collins, and M. F. Franklin. A simplified method for calculating complex metabolic sensitivites by using matrix partitioning. Biochem. J. 263: 289–292, 1989.
 40. Crabtree, B. and G. E. Lobley. Measuring metabolic fluxes in organs and tissues with single and multiple tracers. Proc. Nutr. Soc. 47: 353–364, 1988.
 41. Crabtree, B. and E. A. Newsholme. Comparative aspects of fuel utilization and metabolism by muscle. In: Insect Muscle, edited by P. N. R. Usherwood. London & New York: Academic Press, 1975, p. 405–500.
 42. Crabtree, B. and E. A. Newsholme. Sensitivity of a near‐equilibrium reaction in a metabolic pathway to changes in substrate concentration. Eur. J. Biochem. 89: 19–22, 1978.
 43. Crabtree, B. and E. A. Newsholme. Reply to letter from Susan Moore on substrate cycling. Trends Biochem. Sci. 10: 387, 1985.
 44. Crabtree, B. and E. A. Newsholme. A quantitative approach to metabolic control. Curr. Top. Cell. Regul. 25: 21–76, 1985.
 45. Crabtree, B. and E. A. Newsholme. A systematic approach to describing and analysing metabolic control systems. Trends Biochem. Sci. 12: 4–12, 1987.
 46. Crabtree, B. and E. A. Newsholme. The derivation and interpretation of control coefficients. Biochem. J. 247: 113–120, 1987.
 47. Crabtree, B. and E. A. Newsholme. A method for testing the stability of a steady state system during the calculation of a response to large changes in regulator concentration. FEBS Lett. 280: 329–331, 1991.
 48. Derr, R. F. Modern metabolic control theory II. Determination of flux‐control coefficients. Biochem. Archiv. 2: 31–44, 1986.
 49. Dixon, M. and E. C. Webb. Enzymes. New York: Longmans, 1979.
 50. Easterby, J. S. Biochem. J. 199: 155–161, 1981.
 51. El‐Maghrabi, M., T. Claus, T. Pilkis, E. Fox, and S. Pilkis. J. Biol. Chem. 257: 7603–7607, 1982.
 52. England, P. J. and P. J. Randle. Effectors of rat‐heart hexokinases and the control of rates of glucose phosphorylation in the perfused rat heart. Biochem. J. 105: 907–920, 1967.
 53. Engstrom, L. The regulation of liver pyruvate kinase by phosphorylation‐dephosphorylation Curr. Top. Cell. Regul. 13: 29–52, 1978.
 54. Fell, D. A. and H. M. Sauro. Metabolic control and its analysis. Eur. J. Biochem. 148: 555–561, 1985.
 55. Fell, D. A. and K. Snell. Control analysis of mammalian serine biosynthesis. Biochem. J. 256: 97–101, 1988.
 56. Fischer, E. H., A. Pocker, and J. C. Saari. The structure, function and control of glycogen phosphorylase. Essays Biochem. 6: 23–68, 1970.
 57. Giersch, C. Control analysis of biochemical pathways: a novel procedure for calculating control coefficients, and an additional theorem for branched pathways. J. Theor. Biol. 134: 451–462, 1988.
 58. Gilman, A. G. G‐proteins: transducers of receptor‐generated signals. Annu. Rev. Biochem. 56: 615–649, 1987.
 59. Goldbeter, A. and D. E. Koshland. An amplified sensitivity arising from covalent modification in biological systems. Proc. Nat. Acad. Sci. U.S.A. 78: 6840–6844, 1981.
 60. Goldbeter, A. and D. E. Koshland, Jr.. Energy expenditure in the control of biochemical systems by covalent modification. J. Biol. Chem. 262: 4460–4471, 1987.
 61. Goldhammer, A. R. and H. H. Paradies. Phosphofructokinase: structure and function. Curr. Top. Cell. Regul. 15: 109–141, 1979.
 62. Goldstein, B. M. and A. N. Ivanova. Hormonal regulation of 6‐phosphofructo‐2‐kinase/fructose‐2,6‐bisphosphatase: kinetic models. FEBS Lett. 217: 212–215, 1987.
 63. Goldstein, B. N. and E. L. Shevelev. Stability of multienzyme systems with feedback regulation: a graph‐theoretical approach. J. Theor. Biol. 112: 493–503, 1985.
 64. Groen, A. K., C.W.T. Van Roermudn, R. C. Vervoorn, and J. M. Tager. Control of gluconeogenesis in rat liver cells. Biochem. J. 237: 379–389, 1986.
 65. Groen, A. K., R.J.A. Wanders, H.V. Westerhoff, R. Van Der Meer, and J. M. Tager. Quantification of the contribution of various steps to the control of mitochondrial respiration. J. Biol. Chem. 257: 2754–2757, 1982.
 66. Grynkiewicz, G., M. Poenie, and R. Y. Tsien. A new generation of Ca2+ indicators with greatly improved fluorescence properties. J. Biol. Chem. 260: 3440–3450, 1985.
 67. Halling, P. J. Do the laws of chemistry apply to living cells? Trends Biochem. Sci. 14: 317–318, 1989.
 68. Hanson, P. J. and D. S. Parsons. Metabolism and transport of glutamine and glucose in vascularly perfused small intestine rat. Biochem. J. 166: 509–519, 1977.
 69. Hawthorne, J. N. Phosphoinositides and metabolic control: how many messengers? Biochem. Soc. Trans. 16: 657–660, 1988.
 70. Heinrich, R. and S. M. Rapoport. The utility of mathematical models for the understanding of metabolic systems. Biochem. Soc. Trans. 11: 31–35, 1983.
 71. Heinrich, R., S. M. Rapoport, and T. A. Rapoport. Metabolic regulation and mathematical models. Prog. Biophys. Molec. Biol. 32: 1–82, 1977.
 72. Heinrich, R. and T. A. Rapoport. A linear steady state treatment of enzymatic chains: general properties, control and effector strength. Eur. J. Biochem. 42: 89–95, 1974.
 73. Hofmeyr, J‐H. S. and A. Cornish‐Bowden. Quantitative assessment of regulation in metabolic systems. Eur. J. Biochem. 200: 223–236, 1991.
 74. Irvine, D. H. Objectives, assumptions and results of metabolic control theory and biochemical systems theory. J. Theor. Biol. 143: 139–143, 1990.
 75. Jeffrey, F.M.H., A. Rajagopal, C. R. Malloy, and A. D. Sherry. 13C NMR: a simple yet comprehensive method for analysis of intermediary metabolism. Trends Biochem. Sci. 16: 5–10, 1991.
 76. Kacser, H. The control of enzyme systems in vivo: elasticity analysis of the steady state. Biochem. Soc. Trans. 11: 35–40, 1983.
 77. Kacser, H. and J. A. Burns. The control of flux. Symp. Soc. Exp. Biol. 27: 65–104, 1973.
 78. Kacser, H. and J. A. Burns. Molecular democracy: who shares the controls? Biochem. Soc. Trans. 7: 1149–1160, 1979.
 79. Kacser, H. and J. W. Porteous. Control of metabolism: what do we have to measure? Trends Biochem. Sci. 12: 5–14, 1987.
 80. Katz, J. An n.m.r. study of the tricarboxylic acid cycle. Biochem. J. 263: 997, 1989.
 81. Katz, J. and R. Rognstad. Futile cycles in the metabolism of glucose. Curr. Top. Cell. Regul. 10: 238–287, 1976.
 82. Keizer, J. L. Thermodynamic coupling in chemical reactions. J. Theor. Biol. 49: 323–335, 1975.
 83. Keleti, T. Kinetic power: basis of enzyme efficiency, specificity and evolution. J. Molec. Catal. 47: 271–279, 1988.
 84. Keleti, T. and Vértessy, B., Kinetic power, and theory of control. In Dynamics of Biochemical Systems, edited by S. Damjanovich, T. Keleti and L. Trón Amsterdam: Elsevier, 1996, p. 3–10.
 85. Keleti, T. and J. Ovadi. Control of metabolism by dynamic macromolecular interactions. Curr. Top. Cell. Regul. 29: 1–33, 1988.
 86. Keleti, T. and G. R. Welch. The evolution of enzyme kinetic power. Biochem. J. 223: 299–303, 1984.
 87. Kell, D. B. and H. V. Westerhoff. Metabolic control theory: its role in microbiology and biotechnology. FEMS Microbiol. Rev. 39: 305–320, 1986.
 88. Koerner, T.A.W., R. J. Voll, and E. S. Younathan. A proposed model for the regulation of phosphofructokinase and frustose‐1,6‐bisphosphatase based on their reciprocal anomeric specificities. FEBS Lett. 84: 207–213, 1977.
 89. Koshland, D. R., Jr., Control of enzyme activity and metabolic pathways. In: Metabolic Regulation, edited by R. S. Ochs, R. W. Hanson, and J. Hall. Amsterdam: Elsevier, 1985, p. 1–8.
 90. Krebs, H. A. The Pasteur effect and the relations between respiration and fermentation. Essays Biochem 8: 1–34, 1972.
 91. Kunz, W., F. N. Gellerich, L. Schild, and P. Schonfeld. Kinetic limitations in the overall reactions of mitochendrial oxidative phosphorylation accounting for flux‐dependent changes in the apparent ratio. FEBS Lett. 233: 17–21, 1988.
 92. La Porte, D. C. and D. E. Koshland. A protein with kinase and phosphatase activities involved in regulating the tricarboxylic acid cycle. Nature, 300: 458–460, 1982.
 93. La Porte, D. C. and D. E. Koshland. Phosphorylation of isocitrate dehydrogenase as a demonstration of enhanced sensitivity in covalent regulation. Nature, 305: 286–290, 1983.
 94. LaPorte, D. C., K. Walsh, and D. E. Koshland, Jr.. The branch point effect: ultrasensitivity and subsensitivity to metabolic control. J. Biol. Chem. 259: 14068–14075, 1984.
 95. Madsen, J., J. Bulow, and N. E. Nielsen. Inhibition of fatty acid mobilization by arterial free fatty acid concentration. Acta Physiol. Scand. 127: 161–166, 1986.
 96. Malloy, C. R., A. D. Sherry, and F.M.H. Jeffrey. Carbon flux through citric acid cycle pathway in perfused heart by 13C NMR spectroscopy. FEBS Lett. 212: 58–62, 1987.
 97. Malloy, C. R., A. D. Sherry, and F.M.H. Jeffrez. Evaluation of carbon flux and substrate selection through alternate pathways involving the citric acid cycle of the heart by 3C NMR spectroscopy. J. Biol. Chem. 263: 6964–6971, 1988.
 98. Matsuno, T. Bioenergetics of tumor cells: glutamine metabolism in tumor cell mitochondria. Int. J. Biochem. 19: 303–307, 1987.
 99. Mazur, A. K. A probabilistic view on steady‐state enzyme reactions. J. Theor. Biol. 148: 229–242, 1991.
 100. McCormack, J.G., A. Halestrap, and R. M. Denton. Role of calcium ions in the regulation of mammalian intramitochondrial metabolism. Physiol. Rev. 70: 391–425, 1990.
 101. Meinke, M. H. and R. D. Edstrom. Muscle glycogenolysis: regulation of the cyclic interconversion of phosphorylase a and phosphorylase b. J. Biol. Chem. 266: 2259–2266, 1991.
 102. Monod, J., J‐P. Changeux, and F. Jacob. Allosteric proteins and cellular control systems. J. Mol. Biol. 6: 306–329, 1963.
 103. Morris, J. G. A Biologist's Physical Chemistry. London: Edward Arnold, 1974.
 104. Morris, J. L. Computational Methods in Elementary Numerical Analysis. Chichester: Wiley & Sons, 1983.
 105. Newsholme, E. A. and M. Board. Application of metabolic‐control logic to fuel utilization and its significance in tumor cells. Adv. Enzyme Regul. 31: 225–246, 1991.
 106. Newsholme, E. A., R. A. J. Challiss, and B. Crabtree. Substrate cycles; their role in improving sensitivity in metabolic control. Trends Biochem. Sci. 9: 277–280, 1984.
 107. Newsholme, E. A. and B. Crabtree. Metabolic aspects of enzyme activity regulation. Symp. Soc. Exp. Biol. 27: 429–460, 1973.
 108. Newsholme, E. A. and B. Crabtree. Substrate cycles in metabolic regulation and in heat generation. Biochem. Soc. Symp. 41: 61–109, 1976.
 109. Newsholme, E. A. and B. Crabtree. Theoretical principles in the approaches to the control of metabolic pathways, and their application to the control of glycolysis in muscle. J. Mol. Cell. Cardiol. 11: 839–856, 1979.
 110. Newsholme, E. A. and B. Crabtree. Flux‐generating and regulatory steps in metabolic control. Trends Biochem. Sci. 6: 53–55, 1981.
 111. Newsholme, E. A. and B. Crabtree. Qualitative and quantitative approaches for control: Their current value. In: Regulation of Hepatic Function, edited by N. Grunnet and B. Quistorff. Copenhagen: Munksgaard, 1991, p. 214–226.
 112. Newsholme, E. A., B. Crabtree, and M.S.M. Ardawi. The role of high rates of glycolysis and glutamine utilization in rapidly dividing cells. Biosci. Rep. 5: 393–400, 1985.
 113. Newsholme, E. A., B. Crabtree, and M. Parry‐Billings. The energetic cost of regulation: an analysis based on the principles of metabolic‐control‐logic. In: Energy Metabolism: tissue determinants and cellular corollaries, edited by J. M. Kinney and H. M. Tucker. Raven Press, New York, 1992, p. 467–493.
 114. Newsholme, E. A. and A. R. Leech. Biochemistry for the Medical Sciences. Chichester: John Wiley & Sons, 1983.
 115. Newsholme, E. A., P. Newsholme, R. Curi, E. Challoner, and M.S.M. Ardawai. A role for muscle in the immune system and its importance in surgery, trauma, sepsis and burns. Nutrition, 4: 261–268, 1988.
 116. Newsholme, E. A. and J. C. Stanley. Substrate cycles: their role in control of metabolism with specific references to the liver. Diabetes/Metabolism Reviews. 3: 295–305, 1987.
 117. Newsholme, E. A. and C. Start. General aspects of the regulation of enzyme activity and the effects of hormones. In: Handbook of Physiology. Section 7: Endocrinology, edited by D. F. Steiner and N. Freinkel. Washington, D.C.: American Physiological Society, 1972, p. 369–383.
 118. Newsholme, E. A. and C. Start. Regulation in Metabolism. Chichester: Wiley & Sons, 1973.
 119. Newsholme, P., S. Gordon, and E. A. Newsholme. Rates of utilization and fates of glucose, glutamine, pyruvate, fatty acids and ketone bodies by mouse macrophages. Biochem. J. 242: 631–636, 1987.
 120. Nimmo, H. G. and P. Cohen. Applications of recombinant DNA technology to studies of metabolic regulation. Biochem. J. 247: 1–13, 1987.
 121. Ninio, J. Alternative to the steady state method: derivation of reaction rates from first‐passage times and pathway probabilities. Proc. Natl. Acad. Sci. USA 84: 663–667, 1987.
 122. Ovadi, J. Physiological significance of metabolic channelling. J. Theor. Biol. 152: 1–22, 1991.
 123. Ovadi, J., P. Tompa, B. Vertessy, F. Orosz, T. Keleti, and G. R. Welch. The perfection of substrate‐channelling in interacting enzyme systems: transient‐time analysis. Biochem J. 1991.
 124. Pahl‐Wostl, C. and J. Seelig. Metabolic pathways for ketone body production. 13C NMR spectroscopy of rat liver in vivo using 13C‐multilabeled fatty acids. Biochemistry 25: 6799–6807, 1986.
 125. Parry‐Billings, M. Studies in glutamine metabolism in muscle. Oxford: Thesis, Oxford University, 1989.
 126. Pryor, H. J., J. E. Smyth, P. T. Quinlan, and A. N. Halestrap. Evidence that the flux control coefficient of the respiratory chain is high during gluconeogenesis from lactate in hepatocytes from starved rats. Biochem. J. 247: 449–457, 1987.
 127. Putney, J. W., Jr.. Formation and actions of calcium‐mobilizing messenger inositol‐1,4,5‐trisphosphate. Am. J. Physiol. 252 (Gastrointest. Liver Physiol. 15): G149–G157, 1987.
 128. Randle, P. J. Fuel selection in animals. Biochem. Soc. Trans. 14: 799–806, 1986.
 129. Randle, P. J., P. J. England, and R. M. Denton. Control of the tricarboxylate cycle and its interactions with glycolysis during acetate utilization in rat heart. Biochem. J. 117: 677–695, 1970.
 130. Rapoport, T. A., R. Heinrich, and S. M. Rapoport. The regulatory principles of glycolysis in erythrocytes in vivo and in vitro: a minimal comprehensive model describing steady states, quasi‐steady states and time‐dependent processes. Biochem. J. 154: 449–469, 1976.
 131. Reder, C. Metabolic control theory: a structural approach. J. Theor. Biol. 135: 175–201, 1988.
 132. Ricard, J. Dynamics of multi‐enzyme reactions, cell growth and perception of ionic signals from the external milieu. J. Theor. Biol. 128: 253–278, 1987.
 133. Richter, E. A. and H. Galbo. High Glycogen levels enhance glycogen breakdown in isolated contracting skeletal muscle. J. Appl. Physiol. 61: 827–831, 1986.
 134. Rognstad, R. Rate‐limiting steps in metabolic pathways. J. Biol. Chem. 254: 1875–1878, 1979.
 135. Sacktor, B. Regulation of intermediary metabolism, with special reference to the control mechanisms in insect flight muscle. Adv. Insect Physiol. 7: 267–347, 1970.
 136. Sahlin, K. Control of energetic processes in contracting human skeletal muscle. Biochem. Soc. Trans. 19: 353–358, 1991.
 137. Salter, M., R. G. Knowles, and C. I. Pogson. Quantification of the importance of individual steps in the control of aromatic amino acid metabolism. Biochem. J. 234: 635–647, 1986.
 138. Savageau, M. A. Biochemical systems analysis III. Dynamic solutions using a power‐law approximation. J. Theor. Biol. 26: 215–226, 1970.
 139. Savageau, M. A. The behavior of intact biochemical control systems. Curr. Top. Cell. Regul. 6: 63–130, 1972.
 140. Savageau, M. A. A theory of alternative designs for biochemical control systems. Biomed. Biochim. Acta 6: 875–880, 1985.
 141. Savageau, M. A. Mathematics of organizationally complex systems. Biomed. Biochim. Acta 6: 839–844, 1985.
 142. Savageau, M. A., E. O. Voit, and D. H. Irvine. Biochemical systems theory and metabolic control theory: 2. he role of summation and connectivity relationships. Math. Biosci. 86: 147–169, 1987.
 143. Schiffmann, Y. Self‐organization in biological membranes. Biochem. Soc. Trans. 14: 1195–1196, 1986.
 144. Sen, A. K. Topological analysis of metabolic control. Math. Biosci. 102: 191–223, 1990.
 145. Sen, A. K. Metabolic control analysis. An application of signal flow graphs. Biochem. J. 269: 141–147, 1990.
 146. Sen, A. K. A graph‐theoretic analysis of metabolic regulation in linear pathways with multiple feedback loops and branched pathways. Biochim. Biophys. Acta 1059: 293–311, 1991.
 147. Sen, A. K. Quantitative analysis of metabolic regulation: a graph‐theoretic analysis using spanning trees. Biochem. J. 275: 253–258, 1991.
 148. Sen, A. K. Calculation of control coefficients of metabolic pathways: a flux‐oriented graph‐theoretic approach. Biochem. J. 279: 55–65, 1991.
 149. Shulman, G. I. and L. Rossetti. Influence of the route of glucose administration on hepatic glycogen repletion. Am. J. Physiol. 257 (Endocrinol. Metab. 20) E681–E685, 1989.
 150. Shulman, G. I., D. L. Rothman, T. Jue, P. Stein, R. A. Defronzo, and R. G. Shulman. Quantitation of glycogen synthesis in normal subjects and subjects with non insulin dependent diabetes by 13C nuclear magnetic resonance spectroscopy. New Engl. J. Med. 322: 223–228, 1990.
 151. Sibley, D. R. and R. J. Lefkowitz. Molecular mechanisms of receptor desensitization using the β‐adrenergic receptor‐coupled adenylate cyclase system as a model. Nature 317: 124–129, 1985.
 152. Sibley, D. R., R. H. Strasser, J. L. Benovic, K. Daniel, and R. J. Lefkowitz. Phosphorylation/ dephosphorylation of the β‐adrenergic receptor regulates its functional coupling to adenylate cyclase and subcellular distribution. Proc. Natl. Acad. Sci. USA 83: 9408–9412, 1986.
 153. Small, J. R. and D. A. Fell. Responses of metabolic systems: application of control analysis to yeast glycolysis. Biochem. Soc. Trans. 15: 238, 1987.
 154. Small, J. R. and D. A. Fell. The matrix method of metabolic control analysis: its validity for complex pathway structures. J. Theor. Biol. 136: 181–197, 1989.
 155. Snell, K. and D. A. Fell. Metabolic control analysis of mammalian serine metabolism. Adv. Enzyme Regul. 30: 13–32, 1990.
 156. Sols, A. and C. Gancedo. Primary regulatory enzymes and related proteins. In: Biochemical Regulatory mechanisms in eukaryotic cells, edited by E. Kun and S. Grisolia. New York: Wiley‐Interscience, 1972, p. 85–114.
 157. Sorribas, A. and M. A. Savageau. A comparison of variant theories of intact biochemical systems. II. Flux‐oriented and metabolic control theories. Math. Biosci. 94: 195–238, 1989.
 158. Sorribas, A. and M. A. Savageau. A comparison of variant theories of intact biochemical systems. I. Enzyme‐enzyme interactions and biochemical systems theory. Math. Biosci. 94: 161–193, 1989.
 159. Sorribas, A. and M. A. Savageau. Strategies for representing metabolic pathways within biochemical systems theory: reversible pathways. Math. Biosci. 94: 239–269, 1989.
 160. Srivastava, D. K. and S. A. Bernhard. Metabolite transfer via enzyme‐enzyme complexes. Science. 234: 1081–1086, 1986.
 161. Stadtman, E. R. and P. B. Chock. Interconvertible enzyme cascades in metabolic regulation. Curr. Top. Cell. Regul. 13: 53–95, 1978.
 162. Stephenson, G. Mathematical Methods for Science Students. London: Longmans, 1961.
 163. Stryer, L. Biochemistry. New York: Freeman, 1988.
 164. Stucki, J. W. The optimal efficiency and the economic degrees of coupling of oxidative phosphorylation. Eur. J. Biochem. 109: 269–283, 1980.
 165. Stucki, J. W. The thermodynamic‐buffer enzymes. Eur. J. Biochem. 109: 257–267, 1980.
 166. Tager, J. M., A. K. Groen, R. J. A. Wanders, J. Duszynski, H. V. Westerhoff, and R. C. Vervoorn. Control of mitochondrial respiration. Biochem. Soc. Trans. 11: 40–43, 1983.
 167. Tejwani, G. A., A. Ramaiah, and M. Ananthanaryana. Regulation of glycolysis in muscle. The role of ammonium and synergism among the positive effectors of phosphofructokinase. Arch. Biochem. Biophys. 158: 195–199, 1973.
 168. Tompa, P., J. Batke, J. Ovadi, G. R. Welch, and P. A. Srere. Quantitation of the interaction between citrate synthase and malate dehydrogenase. J. Biol. Chem. 262: 6089–6092, 1987.
 169. Tornheim, K. Activation of muscle phosphofructokinase by fructose‐2,6‐bisphosphate and fructose‐1,6‐bisphosphate is differently affected by other regulatory metabolites. J. Biol. Chem. 260: 7985–7989, 1985.
 170. Traut, T. W. Uridine‐5'‐phosphate synthesis: evidence for substrate cycling involving this bifunctional protein. Arch. Biochem. Biophys. 268: 108–115, 1989.
 171. Umbarger, H. E. Amino acid biosynthesis and its regulation. Annu. Rev. Biochem. 47: 533–606, 1978.
 172. Van Schaftingen, E. Fructose‐2,6‐bisphosphate. Adv. Enzymol. 59: 315–395, 1987.
 173. Vinay, P., G. Lemieux, and a. Gougoux. Characteristics of glutamine metabolism by rat kidney tubules: a carbon and nitrogen balance. Can. J. Biochem. 57: 346–356, 1979.
 174. Vinay, P., J. P. Mapes, and H. A. Krebs. Fate of glutamine carbon in renal metabolism. Am. J. Physiol. 234 (Renal Fluid Electrolyte Physiol. 3): F123–F129, 1978.
 175. Voit, E. O. and M. A. Savageau. Accuracy of alternative representations for integrated biochemical systems. Biochemistry 26: 6869–6880, 1987.
 176. Wajngot, A., V. Chandramouli, W. C. Schumann, K. Kumaran, S. Efendic, and B. R. Landau. Testing of the assumptions made in estimating the extent of futile cycling. Am. J. Physiol. 256 (Endocrinol. Metab. 19): E668–E6765, 1989.
 177. Walker, D. G. The nature and function of hexokinases in animal tissues. Essays Biochem. 2: 33–67, 1966.
 178. Waterlow, J. C., P. J. Garlick, and D. J. Millward. Protein Turnover in Mammalian Tissues and in the Whole Body. Amsterdam: Elsevier, 1978.
 179. Welch, G. R. Some problems in the usage of Gibbs free energy in biochemistry. J. Theor. Biol. 114: 433–446, 1985.
 180. Welch, G. R. and T. Keleti. On the “cytosociology” of enzyme action in vivo: a novel thermodynamic correlate of biological evolution. J. Theor. Biol. 93: 701–735, 1981.
 181. Welch, G. R., T. Keleti, and B. Vertessy. The control of cell metabolism for homogeneous vs. heterogeneous enzyme systems. J. Theor. Biol. 130: 407–422, 1988.
 182. Westerhoff, H. V., A. K. Groen, and R.J.A. Wanders. Modern theories of metabolic control and their applications. Biosci. Rep. 4: 1–22, 1984.
 183. Westerhoff, H. V. and D. B. Kell. Matrix method for determining steps most rate‐limiting to metabolic fluxes in biotechnological processes. Biotechnol. Bioeng. 30: 101–107, 1987.
 184. Williamson, J. R., R. H. Cooper, S. K. Joseph, and A. P. Thomas. Inositol trisphosphate and diacylglycerol as intracellular second messengers in liver. Am. J. Physiol. 248 (Cell Physiol. 17): C203–C216, 1985.
 185. Wilson, D. F, K. Nishiki, and M. Erecinska. Energy metabolism in muscle and its regulation during individual contraction‐relaxation cycles. In: Metabolic Regulation, edited by R. S. Ochs, R. W. Hanson, and J. Hall. Amsterdam: Elsevier, 1985, p. 77–86.
 186. Windmueller, H. G. and A. E. Spaeth. Uptake and metabolism of plasma glutamine by the small intestine. J. Biol. Chem. 249: 5070–5079, 1974.
 187. Windmueller, H. G. and A. E. Spaeth. Identification of ketone bodies and glutamine as the major respiratory fuels in vivo for postabsorptive rat small intestine. J. Biol. Chem. 253: 69–76, 1978.
 188. Woledge, R. C., N. A. Curtin, and E. Homsher. Energetic Aspects of Muscle Contraction. London: Academic Press, 1985.
 189. Woods, N. M., S. R. Cuthbertson, and P. H. Cobbold. Repetitive transient rises in cytoplasmic free calcium in hormone‐stimulated hepatocytes. Nature 319: 600–602, 1986.
 190. Wright, B. E. and P. J. Kelly. Kinetic models of metabolism in intact cells, tissues and organisms. Curr. Top. Cell. Regul. 19: 103–158, 1981.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

B. Crabtree, E. A. Newsholme, N. B. Reppas. Principles of Regulation and Control in Biochemistry: A Pragmatic, Flux‐Oriented Approach. Compr Physiol 2011, Supplement 31: Handbook of Physiology, Cell Physiology: 117-180. First published in print 1997. doi: 10.1002/cphy.cp140105