Skip to main content
Log in

Metabolic Free Energy and Biological Codes: A ‘Data Rate Theorem’ Aging Model

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

A famous argument by Maturana and Varela (Autopoiesis and cognition. Reidel, Dordrecht, 1980) holds that the living state is cognitive at every scale and level of organization. Since it is possible to associate many cognitive processes with ‘dual’ information sources, pathologies can sometimes be addressed using statistical models based on the Shannon Coding, the Shannon–McMillan Source Coding, the Rate Distortion, and the Data Rate Theorems, which impose necessary conditions on information transmission and system control. Deterministic-but-for-error biological codes do not directly invoke cognition, but may be essential subcomponents within larger cognitive processes. A formal argument, however, places such codes within a similar framework, with metabolic free energy serving as a ‘control signal’ stabilizing biochemical code-and-translator dynamics in the presence of noise. Demand beyond available energy supply triggers punctuated destabilization of the coding channel, affecting essential biological functions. Aging, normal or prematurely driven by psychosocial or environmental stressors, must interfere with the routine operation of such mechanisms, initiating the chronic diseases associated with senescence. Amyloid fibril formation, intrinsically disordered protein logic gates, and cell surface glycan/lectin ‘kelp bed’ logic gates are reviewed from this perspective. The results generalize beyond coding machineries having easily recognizable symmetry modes, and strip a layer of mathematical complication from the study of phase transitions in nonequilibrium biological systems.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

References

  • Atlan H, Cohen I (1998) Immune information, self-organization, and meaning. Int Immunol 10:711–717

    Article  Google Scholar 

  • Atiyah M, Singer I (1963) The index of elliptical operators on compact manifolds. Bull Am Math Soc 69:322–433

    Article  MathSciNet  Google Scholar 

  • Beck C, Schlogl F (1993) Thermodynamics of chaotic systems. Cambridge University Press, New York

    Book  Google Scholar 

  • Bredon G (1993) Topology and geometry. Springer, New York

    Book  MATH  Google Scholar 

  • Brown R (1987) From groups to groupoids: a brief survey. Bull Lond Math Soc 19:113–134

    Article  MATH  Google Scholar 

  • Budrikis Z, Costantini G, La Porta C, Zapperi S (2014) Protein accumulation in the endoplasmic reticulum as a non-equilibrium phase transition. Nat Commun 5:3620. doi:10.1038/ncomms4620www.nature.com/naturecommunications

  • Champagnat N, Ferriere R, Meleard S (2006) Unifying evolutionary dynamics: from individual stochastic process to macroscopic models. Theor Popul Biol 69:297–321

    Article  MATH  Google Scholar 

  • Chou KC, Maggiora G (1998) Domain structural class prediction. Protein Eng 11:523–528

    Article  Google Scholar 

  • Cover T, Thomas J (2006) Elements of information theory, 2nd edn. Wiley, New York

    MATH  Google Scholar 

  • Dam T, Gerken T, Cavada B et al (2007) Binding studies of \(\alpha \)-GalNAc-specific lectins to the \(\alpha \)-GalNAc(Tn-antigen) form of procine submaxillary mucin and its smaller fragments. J Biol Chem 38:28256–28263

    Article  Google Scholar 

  • Dam T, Brewer C (2008) Effects of clustered epitopes in multivalent ligan–receptor interactions. Biochemistry 47:8470–8476

    Article  Google Scholar 

  • Dam T, Brewer C (2010) Lectins as pattern recognition molecules: the effects of epitope density in innate immunity. Glycobiology 20:270–279

    Article  Google Scholar 

  • de Groot S, Mazur P (1984) Nonequilibrium thermodynamics. Dover, New York

    Google Scholar 

  • Dembo A, Zeitouni O (1998) Large deviations and applications, 2nd edn. Springer, New York

    Book  MATH  Google Scholar 

  • Ellis R (1985) Entropy, large deviations, and statistical mechanics. Springer, New York

    Book  MATH  Google Scholar 

  • Feynman R (2000) Lectures on computation. Westview Press, New York

    Google Scholar 

  • Ge H, Qian H (2011) Non-equilibrium phase transitions in mesoscopic biochemical systems: from stochastic to nonlinear dynamics and beyond. Interface 8:107–116

    Google Scholar 

  • Goldschmidt L, Teng P, Riek R, Eisenberg D (2010) Identifying the amylome, proteins capable of forming amyloid-like fibrils. Proc Nat Acad Sci 107:3487–3492

    Article  Google Scholar 

  • Hazewinkel M (2002) Encyclopedia of mathematics, ‘index formulas’. Springer, New York

    Google Scholar 

  • Hecht M, Das A, Go A, Aradely L, Wei Y (2004) De novo proteins from designed combinatorialy libraries. Protein Sci 13:1711–1723

    Article  Google Scholar 

  • Horsthemeke W, Lefever R (2006) Noise-induced transitions, vol. 15, theory and applications in physics, chemistry, and biology. Springer, New York

    Google Scholar 

  • Khasminskii R (2012) Stochastic stability of differential equations, 2nd edn. Springer, New York

    Book  MATH  Google Scholar 

  • Khinchin A (1957) Mathematical foundations of information theory. Dover, New York

    MATH  Google Scholar 

  • Landau L, Lifshitz E (2007) Statistical physics, part I. Elsevier, New York

    Google Scholar 

  • Lee J (2000) Introduction to topological manifolds, graduate texts in mathematics series. Springer, New York

    Google Scholar 

  • Lee H, Wei Y (2012) Mitochondria and aging. Chapter 14. In: Scantena et al (eds) Advances in mitochondrial medicine. Springer, New York

    Google Scholar 

  • Levitt M, Chothia C (1976) Structural patterns in globular proteins. Nature 261:552–557

    Article  Google Scholar 

  • Matsumoto Y (2002) An introduction to Morse theory, translations of mathematical monographs, vol 208. American Mathematical Society, Providence

    Google Scholar 

  • Maturana H, Varela F (1980) Autopoiesis and cognition. Reidel, Dordrecht

    Book  Google Scholar 

  • Maury C (2009) Self-propagating \(\beta \)-sheet polypeptide structures as prebiotic informational molecular entities: the amyloid world. Orig Life Evol Biosph 39:141–150

    Article  Google Scholar 

  • Motlagh H, Wrabl J, Li J, Hilser V (2014) The ensemble nature of allostery. Nature 508:331–339

    Article  Google Scholar 

  • Nair G, Fagnani F, Zampieri S, Evans R (2007) Feedback control under data rate constraints: an overview. Proc IEEE 95:108–137

  • Oyelaran O, Li Q, Farnsworth D, Gildersleeve J (2009) Microarrays with varying carbohydrate density reveal distinct subpopulations of serum antibodies. J Proteome Res 8:3529–3538

    Article  Google Scholar 

  • Pettini M (2007) Geometry and topology in Hamiltonian dynamics. Springer, New York

    Book  MATH  Google Scholar 

  • Protter P (1990) Stochastic integration and differential equations. Springer, New York

    Book  MATH  Google Scholar 

  • Ringel G, Young J (1968) Solutions of the Heawood map-coloring problem. Proc Nat Acad Sci 60:438–445

    Article  MathSciNet  MATH  Google Scholar 

  • Rockafellar R (1970) Convex analysis. Princeton University Press, Princeton

    Book  MATH  Google Scholar 

  • Sawaya M, Sambashivan S, Nelson R et al (2007) Atomic structures of amyloid cross-\(\beta \) splines reveal varied steric zippers. Nature 447:453–457

    Article  Google Scholar 

  • Shmulevich I, Dougherty E (2007) Genomic signal processing. Princeton University Press, Princeton

    Book  MATH  Google Scholar 

  • Tlusty T (2007) A model for the emergence of the genetic code as a transition in a noisy information channel. J Theor Biol 249:331–342

    Article  MathSciNet  Google Scholar 

  • Tlusty T (2008a) A simple model for the evolution of molecular codes driven by the interplay of accuracy, diversity and cost. Physical Biology 5:016001

    Article  Google Scholar 

  • Tlusty T (2008b) Casting polymer nets to optimize noisy molecular codes. Proc Natl Acad Sci 105:8238–8243

    Article  Google Scholar 

  • Tomkins G (1975) The metabolic code. Science 189:760–763

    Article  Google Scholar 

  • Tompa P, Szasz C, Buday L (2005) Structural disorder throws new light on moonlighting. Trends Biochem Sci 30:484–489

    Article  Google Scholar 

  • Van den Broeck C, Parrondo J, Toral R (1994) Noise-induced nonequilibrium phase transition. Phys Rev Lett 73:3395–3398

    Article  Google Scholar 

  • Van den Broeck C, Parrondo J, Toral R, Kawai R (1997) Nonequilibrium phase transitions induced by multiplicative noise. Phys Rev E 55:4084–4094

    Article  Google Scholar 

  • Wallace R (2005) Consciousness: a mathematical treatment of the global neuronal workspace model. Springer, New York

    Google Scholar 

  • Wallace R (2007) Culture and inattentional blindness: a global workspace perspective. J Theor Biol 245:378–390

    Article  Google Scholar 

  • Wallace R (2010) A scientific open season. Phys Life Rev 7:377–378

    Article  Google Scholar 

  • Wallace R (2011) Multifunction moonlighting and intrinsically disordered proteins: information catalysis, non-rigid molecule symmetries and the ‘logic gate’ spectrum. C R Chim 14:1117–1121

    Article  Google Scholar 

  • Wallace R (2012a) Metabolic constraints on the evolution of genetic codes: did multiple ‘preaerobic’ ecosystem transitions entrain richer dialects via serial endosymbiosis? Trans Comput Syst Biol XIV, LNBI 7625:204–232

    Article  Google Scholar 

  • Wallace R (2012b) Extending Tlusty’s rate distortion index theorem method to the glycome: do even ‘low level’ biochemical phenomena require sophisticated cognitive paradigms? BioSystems 107:145–152

    Article  Google Scholar 

  • Wallace R (2012c) Consciousness, crosstalk, and the mereological fallacy: an evolutionary perspective. Phys Life Rev 9:426–453

  • Wallace R (2012d) Spontaneous symmetry breaking in a non-rigid molecule approach to intrinsically disordered proteins. Mol BioSyst 8:374–377

    Article  Google Scholar 

  • Wallace R (2014a) Cognition and biology: perspectives from information theory. Cogn Process 15:1–12

    Article  Google Scholar 

  • Wallace R (2014b) Embodied cognition, embodied regulation, and the Data Rate Theorem. Biol Inspired Cogn Archit. doi:10.1016/j.bica.2014.02.003

  • Wallace R (2014c) Canonical failure modes of real-time control systems: insights from cognitive theory. Int J Syst Sci. doi:10.1080/00207721.2014.923951

  • Wallace R, Wallace D (2010) Gene expression and its discontents: the social production of chronic disease. Springer, New York

    Book  Google Scholar 

  • Wallace R, Wallace D (2013) A mathematical approach to multilevel, multiscale health interventions: pharmaceutical industry decline and policy response. Imperial College Press, London

    Book  Google Scholar 

  • Weinstein A (1996) Groupoids: unifying internal and external symmetry. Not Am Math Assoc 43:744–752

    MATH  Google Scholar 

  • Wilson K (1971) Renormalization group and critical phenomena I. Renormalization group and the Kadanoff scaling picture. Phys Rev B 4:3174–3183

    Article  MATH  Google Scholar 

Download references

Acknowledgments

The author thanks Dr. D. N. Wallace for useful discussions, and the reviewers for comments helpful in revision.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rodrick Wallace.

Mathematical Appendix

Mathematical Appendix

1.1 The Data Rate Theorem

The DRT, a generalization of the classic Bode Integral Theorem for linear control systems, describes the stability of feedback control under data rate constraints (Nair et al. 2007). Given a noise-free data link between a discrete linear plant and its controller, unstable modes can be stabilized only if the feedback data rate \(\mathcal {I}\) is greater than the rate of ‘topological information’ generated by the unstable system. For the simplest incarnation, if the linear matrix equation of the plant is of the form \(x_{t+1}=\mathbf {A}x_{t}+ \cdots \), where \(x_{t}\) is the n-dimensional state vector at time \(t\), then the necessary condition for stabilizability is that

$$\begin{aligned} \mathcal {I} > \log [|\hbox {det}\mathbf {A}^{u}|] \end{aligned}$$
(30)

where det is the determinant and \(\mathbf {A}^{u}\) is the decoupled unstable component of \(\mathbf {A}\), i.e., the part having eigenvalues \(\ge \)1. The determinant represents a generalized volume. Thus there is a critical positive data rate below which there does not exist any quantization and control scheme able to stabilize an unstable system (Nair et al. 2007).

The new theorem, and its variations, relate control theory to information theory and are as fundamental as the Shannon Coding and Source Coding Theorems, and the Rate Distortion Theorem for understanding complex cognitive machines and biological phenomena (Cover and Thomas 2006). Very clearly, however, the DRT speaks to the centrality of the question whether nonequilibrium phase transitions play an important role in living systems.

1.2 Morse Theory

Morse Theory explores relations between analytic behavior of a function—the location and character of its critical points—and the underlying topology of the manifold on which the function is defined. We are interested in a number of such functions, for example, information source uncertainty on a parameter space and possible iterations involving parameter manifolds determining critical behavior. An example might be the sudden onset of a giant component when ‘weak ties’ exceed a threshold. These can be reformulated from a Morse Theory perspective (Pettini 2007).

The basic idea of Morse Theory is to examine an \(n\)-dimensional manifold \(M\) as decomposed into level sets of some function \(f: M \rightarrow \mathbf {R}\) where \(\mathbf {R}\) is the set of real numbers. The \(a\)-level set of \(f\) is defined as

$$\begin{aligned} f^{-1}(a) = \{x \in M: f(x)=a\}, \end{aligned}$$

the set of all points in \(M\) with \(f(x)=a\). If \(M\) is compact, then the whole manifold can be decomposed into such slices in a canonical fashion between two limits, defined by the minimum and maximum of \(f\) on \(M\). Let the part of \(M\) below \(a\) be defined as

$$\begin{aligned} M_{a} = f^{-1}(-\infty , a] = \{x \in M : f(x) \le a\}. \end{aligned}$$

These sets describe the whole manifold as \(a\) varies between the minimum and maximum of \(f\).

Morse functions are defined as a particular set of smooth functions \(f:M \rightarrow \mathbf {R}\) as follows. Suppose a function \(f\) has a critical point \(x_{c}\), so that the derivative \(df(x_{c})=0\), with critical value \(f(x_{c})\). Then, \(f\) is a Morse function if its critical points are nondegenerate in the sense that the Hessian matrix of second derivatives at \(x_{c}\), whose elements, in terms of local coordinates are

$$\begin{aligned} \mathcal {H}_{i,j}= \partial ^{2}f/\partial x^{i} \partial x^{j}, \end{aligned}$$

has rank \(n\), which means that it has only nonzero eigenvalues, so that there are no lines or surfaces of critical points and, ultimately, critical points are isolated.

The index of the critical point is the number of negative eigenvalues of \(\mathcal {H}\) at \(x_{c}\).

A level set \(f^{-1}(a)\) of \(f\) is called a critical level if \(a\) is a critical value of \(f\), that is, if there is at least one critical point \(x_{c} \in f^{-1}(a)\).

Again following Pettini (2007), the essential results of Morse Theory are:

  1. 1.

    If an interval \([a,b]\) contains no critical values of \(f\), then the topology of \(f^{-1}[a, v]\) does not change for any \(v \in (a, b]\). Importantly, the result is valid even if \(f\) is not a Morse function, but only a smooth function.

  2. 2.

    If the interval \([a, b]\) contains critical values, the topology of \(f^{-1}[a,v]\) changes in a manner determined by the properties of the matrix \(H\) at the critical points.

  3. 3.

    If \(f :M \rightarrow \mathbf {R}\) is a Morse function, the set of all the critical points of \(f\) is a discrete subset of \(M\), i.e., critical points are isolated. This is Sard’s Theorem.

  4. 4.

    If \(f : M \rightarrow \mathbf {R}\) is a Morse function, with \(M\) compact, then on a finite interval \([a, b] \subset \mathbf {R}\), there is only a finite number of critical points \(p\) of \(f\) such that \(f(p) \in [a, b]\). The set of critical values of \(f\) is a discrete set of \(\mathbf {R}\).

  5. 5.

    For any differentiable manifold \(M\), the set of Morse functions on \(M\) is an open dense set in the set of real functions of \(M\) of differentiability class \(r\) for \(0 \le r \le \infty \).

  6. 6.

    Some topological invariants of \(M\), that is, quantities that are the same for all the manifolds that have the same topology as \(M\), can be estimated and sometimes computed exactly once all the critical points of \(f\) are known: let the Morse numbers \(\mu _{i}(i=0, \)...\(, m)\) of a function \(f\) on \(M\) be the number of critical points of \(f\) of index \(i\), (the number of negative eigenvalues of \(\mathcal {H}\)). The Euler characteristic of the complicated manifold \(M\) can be expressed as the alternating sum of the Morse numbers of any Morse function on \(M\),

    $$\begin{aligned} \chi = \sum _{i=1}^{m} (-1)^{i}\mu _{i}. \end{aligned}$$

    The Euler characteristic reduces, in the case of a simple polyhedron, to

    $$\begin{aligned} \chi = V - E + F \end{aligned}$$

    where \(V, E\), and \(F\) are the numbers of vertices, edges, and faces in the polyhedron.

  7. 7.

    Another important theorem states that, if the interval \([a,b]\) contains a critical value of \(f\) with a single critical point \(x_{c}\), then the topology of the set \(M_{b}\) defined above differs from that of \(M_{a}\) in a way which is determined by the index, \(i\), of the critical point. Then \(M_{b}\) is homeomorphic to the manifold obtained from attaching to \(M_{a}\) an \(i\)-handle, i.e., the direct product of an \(i\)-disk and an \((m-i)\)-disk.

Pettini (2007) and Matsumoto (2002) contain details and further references.

1.3 Groupoids

A groupoid, \(G\), is defined by a base set \(A\) upon which some mapping—a morphism—can be defined. Note that not all possible pairs of states \((a_{j}, a_{k})\) in the base set \(A\) can be connected by such a morphism. Those that can define the groupoid element, a morphism \(g=(a_{j},a_{k})\) having the natural inverse \(g^{-1}=(a_{k}, a_{j})\). Given such a pairing, it is possible to define ‘natural’ end-point maps \(\alpha (g)=a_{j}, \beta (g)=a_{k}\) from the set of morphisms \(G\) into \(A\), and a formally associative product in the groupoid \(g_{1}g_{2}\) provided \(\alpha (g_{1}g_{2})=\alpha (g_{1}), \beta (g_{1}g_{2})=\beta (g_{2})\), and \(\beta (g_{1})=\alpha (g_{2})\). Then, the product is defined, and associative, \((g_{1}g_{2})g_{3}=g_{1}(g_{2}g_{3})\). In addition, there are natural left and right identity elements \(\lambda _{g}, \rho _{g}\) such that \(\lambda _{g}g=g=g\rho _{g}\).

An orbit of the groupoid \(G\) over \(A\) is an equivalence class for the relation \(a_{j} \sim G a_{k}\) if and only if there is a groupoid element \(g\) with \(\alpha (g)=a_{j}\) and \(\beta (g)=a_{k}\). A groupoid is called transitive if it has just one orbit. The transitive groupoids are the building blocks of groupoids in that there is a natural decomposition of the base space of a general groupoid into orbits. Over each orbit, there is a transitive groupoid, and the disjoint union of these transitive groupoids is the original groupoid. Conversely, the disjoint union of groupoids is itself a groupoid.

The isotropy group of \(a \in X\) consists of those \(g\) in \(G\) with \(\alpha (g)=a=\beta (g)\). These groups prove fundamental to classifying groupoids.

If \(G\) is any groupoid over \(A\), the map \((\alpha , \beta ):G \rightarrow A \times A\) is a morphism from \(G\) to the pair groupoid of \(A\). The image of \((\alpha , \beta )\) is the orbit equivalence relation \(\sim G\), and the functional kernel is the union of the isotropy groups. If \(f: X \rightarrow Y\) is a function, then the kernel of \(f,\, ker(f)=[(x_{1},x_{2})\in X \times X:f(x_{1})=f(x_{2})]\) defines an equivalence relation.

Groupoids may have additional structure. For example, a groupoid \(G\) is a topological groupoid over a base space \(X\) if \(G\) and \(X\) are topological spaces and \(\alpha , \beta \) and multiplication are continuous maps.

In essence, a groupoid is a category in which all morphisms have an inverse, here defined in terms of connection to a base point by a meaningful path of an information source dual to a cognitive process.

The morphism \((\alpha , \beta )\) suggests another way of looking at groupoids. A groupoid over \(A\) identifies not only which elements of \(A\) are equivalent to one another (isomorphic), but it also parameterizes the different ways (isomorphisms) in which two elements can be equivalent, i.e., in our context, all possible information sources dual to some cognitive process. Given the information theoretic characterization of cognition presented above, this produces a full modular cognitive network in a highly natural manner.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wallace, R. Metabolic Free Energy and Biological Codes: A ‘Data Rate Theorem’ Aging Model. Bull Math Biol 77, 879–903 (2015). https://doi.org/10.1007/s11538-014-0013-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-014-0013-0

Keywords

Navigation