Summation theorems (biochemistry)

Last updated

In metabolic control analysis, a variety of theorems have been discovered and discussed in the literature. [1] [2] [3] [4] The most well known of these are flux and concentration control coefficient summation relationships. These theorems are the result of the stoichiometric structure and mass conservation properties of biochemical networks. [5] [6] Equivalent theorems have not been found, for example, in electrical or economic systems.

Contents

The summation of the flux and concentration control coefficients were discovered independently by the Kacser/Burns group [7] and the Heinrich/Rapoport group [8] in the early 1970s and late 1960s.

If we define the control coefficients using enzyme concentration, then the summation theorems are written as:

However these theorems depend on the assumption that reaction rates are proportional to enzyme concentration. An alternative way to write the theorems is to use control coefficients that are defined with respect to the local rates which is therefore independent of how rates respond to changes in enzyme concentration:

Although originally derived for simple linear chains of enzyme catalyzed reactions, it became apparent that the theorems applied to pathways of any structure including pathways with complex regulation involving feedback control. [9] [10]

Derivation

There are different ways to derive the summation theorems. One is analytical and rigorous using a combination of linear algebra and calculus. [11] The other is less rigorous, but more operational and intuitive. The latter derivation is shown here.

Consider the two-step pathway:

where and are fixed species so that the system can achieve a steady-state.

Let the pathway be at steady-state and imagine increasing the concentration of enzyme, , catalyzing the first step, , by an amount, . The effect of this is to increase the steady-state levels of S and flux, J. Let us now increase the level of by such that the change in S is restored to the original value it had at steady-state.

The net effect of these two changes is by definition, .

There are two ways to look at this thought experiment, from the perspective of the system and from the perspective of local changes. For the system we can compute the overall change in flux or species concentration by adding the two control coefficient terms, thus:

We can also look at what is happening locally at every reaction step for which there will be two: one for , and another for . Since the thought experiment guarantees that , the local equations are quite simple:

where the terms are the elasticities. However, because the enzyme elasticity is equal to one, these reduce to:

Because the pathway is linear, at steady-state, . We can substitute these expressions into the system equations to give:

Note that at steady state the change in and must be the same, therefore .

Setting , we can rewrite the above equations as:

We then conclude through cancelation of since , that:

Interpretation

The summation theorems can be interpreted in various ways. The first is that the influence enzymes have over steady-state fluxes and concentrations is not necessarily concentrated at one location. In the past, control of a pathway was considered to be located at one point only, called the master reaction or rate limiting step. The summation theorem suggests this does not necessarily have to be the case.

The flux summation theorem also suggests that there is a total amount of flux control in a pathway such that if one step gains control another step most lose control.

Plot of a phenotype, such as a flux, as a function of enzyme level. The response is hyperbolic like such that a homozygote dominant individual and the heterozygote individual are phenotypically indistinguishable due to the curvature of the response. DominancePlot.png
Plot of a phenotype, such as a flux, as a function of enzyme level. The response is hyperbolic like such that a homozygote dominant individual and the heterozygote individual are phenotypically indistinguishable due to the curvature of the response.

Although flux control is shared, this doesn't imply that control is evenly distributed. For a large network, the average flux control will, according to the flux summation theorem, be equal to , that is a small number. In order for a biological cell to have any appreciable control over a pathway via changes in gene expression, some concentration of flux control at a small number of sites will be necessary. For example, in mammalian cancer cell lines, it has been shown [12] that flux control is concentrated at four sites: glucose import, hexokinase, phosphofructokinase, and lactate export.

Moreover, Kacser and Burns [13] suggested that since the flux–enzyme relationship is somewhat hyperbolic, and that for most enzymes, the wild-type diploid level of enzyme activity occurs where the curve is reaching a point in the curve where changes have little effect, then since a heterozygote of the wild-type with a null mutant will have half the enzyme activity it will not exhibit a noticeably reduced flux. Therefore, the wild type appears dominant and the mutant recessive because of the system characteristics of a metabolic pathway. Although originally suggested by Sewall Wright, [14] [15] the development of metabolic control analysis put the idea on a more sound theoretical footing. The flux summation theorem in particular is consistent with the flux summation theorem for large systems. Not all dominance properties can be explained in this way but it does offers an explanation for dominance at least at the metabolic level. [16]

Concentration summation theorem

In contrast to the flux summation theorem, the concentration summation theorem sums to zero. The implications of this are that some enzymes will cause a given metabolite to increase while others, in order to satisfy the summation to zero, must cause the same metabolite to decrease. This is particularly noticeable in a linear chain of enzyme reactions where, given a metabolite located in the center of the pathway, an increase in expression of any enzyme upstream of the metabolite will cause the metabolite to increase in concentration. In contrast, an increase in expression of any enzyme downstream of the metabolite will cause the given metabolite to decrease in concentration. [17]

See also

Related Research Articles

<span class="mw-page-title-main">Dirac delta function</span> Generalized function whose value is zero everywhere except at zero

In mathematical analysis, the Dirac delta distribution, also known as the unit impulse, is a generalized function or distribution over the real numbers, whose value is zero everywhere except at zero, and whose integral over the entire real line is equal to one.

<span class="mw-page-title-main">Stress–energy tensor</span> Tensor describing energy momentum density in spacetime

The stress–energy tensor, sometimes called the stress–energy–momentum tensor or the energy–momentum tensor, is a tensor physical quantity that describes the density and flux of energy and momentum in spacetime, generalizing the stress tensor of Newtonian physics. It is an attribute of matter, radiation, and non-gravitational force fields. This density and flux of energy and momentum are the sources of the gravitational field in the Einstein field equations of general relativity, just as mass density is the source of such a field in Newtonian gravity.

<span class="mw-page-title-main">Noether's theorem</span> Statement relating differentiable symmetries to conserved quantities

Noether's theorem or Noether's first theorem states that every differentiable symmetry of the action of a physical system with conservative forces has a corresponding conservation law. The theorem was proven by mathematician Emmy Noether in 1915 and published in 1918. The action of a physical system is the integral over time of a Lagrangian function, from which the system's behavior can be determined by the principle of least action. This theorem only applies to continuous and smooth symmetries over physical space.

In mathematical analysis, Cesàro summation assigns values to some infinite sums that are not necessarily convergent in the usual sense. The Cesàro sum is defined as the limit, as n tends to infinity, of the sequence of arithmetic means of the first n partial sums of the series.

A temperature coefficient describes the relative change of a physical property that is associated with a given change in temperature. For a property R that changes when the temperature changes by dT, the temperature coefficient α is defined by the following equation:

<span class="mw-page-title-main">Thermal expansion</span> Tendency of matter to change volume in response to a change in temperature

Thermal expansion is the tendency of matter to change its shape, area, volume, and density in response to a change in temperature, usually not including phase transitions.

In calculus, the Leibniz integral rule for differentiation under the integral sign states that for an integral of the form

<span class="mw-page-title-main">Flux balance analysis</span>

Flux balance analysis (FBA) is a mathematical method for simulating metabolism in genome-scale reconstructions of metabolic networks. In comparison to traditional methods of modeling, FBA is less intensive in terms of the input data required for constructing the model. Simulations performed using FBA are computationally inexpensive and can calculate steady-state metabolic fluxes for large models in a few seconds on modern personal computers. The related method of metabolic pathway analysis seeks to find and list all possible pathways between metabolites.

<span class="mw-page-title-main">Metabolic control analysis</span> Metabolic control

Metabolic control analysis (MCA) is a mathematical framework for describing metabolic, signaling, and genetic pathways. MCA quantifies how variables, such as fluxes and species concentrations, depend on network parameters. In particular, it is able to describe how network-dependent properties, called control coefficients, depend on local properties called elasticities or Elasticity Coefficients.

In electrochemistry, the Butler–Volmer equation, also known as Erdey-Grúz–Volmer equation, is one of the most fundamental relationships in electrochemical kinetics. It describes how the electrical current through an electrode depends on the voltage difference between the electrode and the bulk electrolyte for a simple, unimolecular redox reaction, considering that both a cathodic and an anodic reaction occur on the same electrode:

The rate of a chemical reaction is influenced by many different factors, such as temperature, pH, reactant, and product concentrations and other effectors. The degree to which these factors change the reaction rate is described by the elasticity coefficient. This coefficient is defined as follows:

<span class="mw-page-title-main">Lattice diffusion coefficient</span> Atomic diffusion within a crystalline lattice

In condensed matter physics, lattice diffusion refers to atomic diffusion within a crystalline lattice, which occurs by either interstitial or substitutional mechanisms. In interstitial lattice diffusion, a diffusant, will diffuse in between the lattice structure of another crystalline element. In substitutional lattice diffusion, the atom can only move by switching places with another atom. Substitutional lattice diffusion is often contingent upon the availability of point vacancies throughout the crystal lattice. Diffusing particles migrate from point vacancy to point vacancy by the rapid, essentially random jumping about. Since the prevalence of point vacancies increases in accordance with the Arrhenius equation, the rate of crystal solid state diffusion increases with temperature. For a single atom in a defect-free crystal, the movement can be described by the "random walk" model.

<span class="mw-page-title-main">Dynamic fluid film equations</span>

Fluid films, such as soap films, are commonly encountered in everyday experience. A soap film can be formed by dipping a closed contour wire into a soapy solution as in the figure on the right. Alternatively, a catenoid can be formed by dipping two rings in the soapy solution and subsequently separating them while maintaining the coaxial configuration.

Laser theory of Fabry-Perot (FP) semiconductor lasers proves to be nonlinear, since the gain, the refractive index and the loss coefficient are the functions of energy flux. The nonlinear theory made it possible to explain a number of experiments some of which could not even be explained, much less modeled, on the basis of other theoretical models; this suggests that the nonlinear theory developed is a new paradigm of the laser theory.

Control coefficients are used to describe how much influence a given reaction step has on the steady-state flux or species concentration level. In practice, this can be accomplished by changing the expression level of a given enzyme and measuring the resulting changes in flux and metabolite levels. Control coefficients form a central component of metabolic control analysis.

In biochemistry, a rate-limiting step is a step that controls the rate of a series of biochemical reactions. The statement is, however, a misunderstanding of how a sequence of enzyme catalyzed reaction steps operate. Rather than a single step controlling the rate, it has been discovered that multiple steps control the rate. Moreover, each controlling step controls the rate to varying degrees.

<span class="mw-page-title-main">Branched pathways</span> Common pattern in metabolism

Branched pathways, also known as branch points, are a common pattern found in metabolism. This is where an intermediate species is chemically made or transformed by multiple enzymatic processes. linear pathways only have one enzymatic reaction producing a species and one enzymatic reaction consuming the species.

Control coefficients measure the response of a biochemical pathway to changes in enzyme activity. The response coefficient, as originally defined by Kacser and Burns, is a measure of how external factors such as inhibitors, pharmaceutical drugs, or boundary species affect the steady-state fluxes and species concentrations. The flux response coefficient is defined by:

The stoichiometric structure and mass-conservation properties of biochemical pathways gives rise to a series of theorems or relationships between the control coefficients and the control coefficients and elasticities. There are a large number of such relationships depending on the pathway configuration which have been documented and discovered by various authors. The term theorem has been used to describe these relationships because they can be proved in terms of more elementary concepts. The operational proofs in particular are of this nature.

Biological cells consume nutrients to sustain life. These nutrients are broken down to smaller molecules which are then reassembled into more complex structures required for life. The breakdown and reassembly of nutrients is called metabolism. An individual cell will contain thousands of different kinds of small molecules, such as sugars, lipids, amino acids, etc. The interconversion of these molecules is carried out by catalysts called enzymes. For example, E. coli contains 2,338 metabolic enzymes. These enzymes form a complex web of reactions forming pathways by which nutrients are converted. These pathways come in various forms. For example, enzymes can form simple linear sequences of reactions forming a linear biochemical pathway. Pathways can also show other patterns such as branches or cycles. A famous cyclic pathway is Kreb's cycle or the Calvin's cycle.

References

  1. Westerhoff, Hans V. (27 May 2023). "Summation Laws in Control of Biochemical Systems". Mathematics. 11 (11): 2473. doi: 10.3390/math11112473 .
  2. Sauro, Herbert M.; Small, J. Rankin; Fell, David A. (May 1987). "Metabolic control and its analysis. Extensions to the theory and matrix method". European Journal of Biochemistry. 165 (1): 215–221. doi:10.1111/j.1432-1033.1987.tb11214.x. PMID   3569295.
  3. Hofmeyr, Jan-Hendrik S.; Kacser, Henrik; Merwe, Kirsten J. (March 1986). "Metabolic control analysis of moiety-conserved cycles". European Journal of Biochemistry. 155 (3): 631–640. doi: 10.1111/j.1432-1033.1986.tb09534.x . PMID   3956502.
  4. Rankin Small, J.; Fell, David A. (January 1989). "The matrix method of metabolic control analysis: its validity for complex pathway structures". Journal of Theoretical Biology. 136 (2): 181–197. Bibcode:1989JThBi.136..181R. doi:10.1016/S0022-5193(89)80225-5. PMID   2779266.
  5. Schuster, Stefan (October 1996). "Control Analysis in Terms of Generalized Variables Characterizing Metabolic Systems". Journal of Theoretical Biology. 182 (3): 259–268. Bibcode:1996JThBi.182..259S. doi:10.1006/jtbi.1996.0163. PMID   8944157.
  6. Hofmeyr, Jannie (2001). "Metabolic control analysis in a nutshell". S2CID   17007756.{{cite journal}}: Cite journal requires |journal= (help)
  7. Kacser, H.; Burns, J. A. (1973). "The control of flux". Symposia of the Society for Experimental Biology. 27: 65–104. PMID   4148886.
  8. Heinrich, R.; Rapoport, T. A. (1974). "A linear steady-state treatment of enzymatic chains. General properties, control and effector strength". European Journal of Biochemistry. 42 (1): 89–95. doi: 10.1111/j.1432-1033.1974.tb03318.x . PMID   4830198.
  9. Reder, Christine (November 1988). "Metabolic control theory: A structural approach". Journal of Theoretical Biology. 135 (2): 175–201. Bibcode:1988JThBi.135..175R. doi:10.1016/S0022-5193(88)80073-0. PMID   3267767.
  10. Heinrich R. and Schuster S. (1996) The Regulation of Cellular Systems, Chapman and Hall.
  11. Heinrich R. and Schuster S. (1996) The Regulation of Cellular Systems, Chapman and Hall.
  12. Tanner, Lukas Bahati; Goglia, Alexander G.; Wei, Monica H.; Sehgal, Talen; Parsons, Lance R.; Park, Junyoung O.; White, Eileen; Toettcher, Jared E.; Rabinowitz, Joshua D. (July 2018). "Four Key Steps Control Glycolytic Flux in Mammalian Cells". Cell Systems. 7 (1): 49–62.e8. doi:10.1016/j.cels.2018.06.003. PMC   6062487 . PMID   29960885.
  13. Kacser, Henrik; Burns, James A (1 March 1981). "The Molecular Basis of Dominance". Genetics. 97 (3–4): 639–666. doi:10.1093/genetics/97.3-4.639. PMC   1214416 . PMID   7297851.
  14. Wright, Sewall (January 1934). "Physiological and Evolutionary Theories of Dominance". The American Naturalist. 68 (714): 24–53. doi:10.1086/280521. S2CID   84400871.
  15. Vasseur, François; Fouqueau, Louise; de Vienne, Dominique; Nidelet, Thibault; Violle, Cyrille; Weigel, Detlef (24 April 2019). "Nonlinear phenotypic variation uncovers the emergence of heterosis in Arabidopsis thaliana". PLOS Biology. 17 (4): e3000214. doi: 10.1371/journal.pbio.3000214 . PMC   6481775 . PMID   31017902.
  16. Billiard, Sylvain; Castric, Vincent; Llaurens, Violaine (December 2021). "The integrative biology of genetic dominance". Biological Reviews. 96 (6): 2925–2942. doi:10.1111/brv.12786. PMC   9292577 . PMID   34382317.
  17. Sauro, Herbert (2013). Systems biology: an introduction to metabolic control analysis (1st, version 1.01 ed.). Seattle, WA: Ambrosius Publishing. ISBN   978-0982477366.