Skip to main content
Log in

Growth and Dissolution of Macromolecular Markov Chains

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

The kinetics and thermodynamics of free living copolymerization are studied for processes with rates depending on k monomeric units of the macromolecular chain behind the unit that is attached or detached. In this case, the sequence of monomeric units in the growing copolymer is a kth-order Markov chain. In the regime of steady growth, the statistical properties of the sequence are determined analytically in terms of the attachment and detachment rates. In this way, the mean growth velocity as well as the thermodynamic entropy production and the sequence disorder can be calculated systematically. These different properties are also investigated in the regime of depolymerization where the macromolecular chain is dissolved by the surrounding solution. In this regime, the entropy production is shown to satisfy Landauer’s principle.

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
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  1. Flory, P.J.: Principles of Polymer Chemistry. Cornell University Press, Ithaca (1953)

    Google Scholar 

  2. Odian, G.: Principles of Polymerization, 4th edn. Wiley, Hoboken (2004)

    Book  Google Scholar 

  3. Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K., Watson, J.D.: Molecular Biology of the Cell, 2nd edn. Garland Publishing, New York (1989)

    Google Scholar 

  4. Ring, W., Mita, I., Jenkins, A.D., Bikales, N.M.: Source-based nomenclature for copolymers. Pure Appl. Chem. 57, 1427–1440 (1985)

    Article  Google Scholar 

  5. Jenkins, A.D., Kratochvil, P., Stepto, R.F.T., Suter, U.W.: Glossary of basic terms in polymer science. Pure Appl. Chem. 68, 2287–2311 (1996)

    Google Scholar 

  6. Randall, J.C.: A \(^{13}\) C NMR determination of the comonomer sequence distributions in propylene-butene-1 copolymers. Macromolecules 11, 592–597 (1978)

    Article  ADS  Google Scholar 

  7. Randall, J.C., Rucker, S.P.: Markovian statistics for finite chains: characterization of end group structures and initiation, chain propagation, and chain-transfer probabilities in poly(ethylene-co-propylene). Macromolecules 27, 2120–2129 (1994)

    Article  ADS  Google Scholar 

  8. Seger, M.R., Maciel, G.E.: Quantitative \(^{13}\) C NMR analysis of sequence distributions in poly(ethylene-co-1-hexene). Anal. Chem. 76, 5734–5747 (2004)

    Article  Google Scholar 

  9. Osakada, K., Choi, J.-C., Yamamoto, T.: \(\pi \)-Allylic rhodium complex catalyzed living copolymerization of arylallenes with carbon monoxide to give structurally regulated polyketones. J. Am. Chem. Soc. 119, 12390–12391 (1997)

    Article  Google Scholar 

  10. Yoshida, Y., Mohri, J.-I., Ishii, S.-I., Mitani, M., Saito, J., Matsui, S., Makio, H., Nakano, T., Tanaka, H., Onda, M., Yamamoto, Y., Mizuno, A., Fujita, T.: Living copolymerization of ethylene with norbornene catalyzed by bis(pyrrolide-imine) titanium complexes with MAO. J. Am. Chem. Soc. 126, 12023–12032 (2004)

    Article  Google Scholar 

  11. Zhang, H., Nomura, K.: Living copolymerization of ethylene with styrene catalyzed by (cyclopentadienyl)(ketimide)titanium(IV) complex-MAO catalyst system. J. Am. Chem. Soc. 127, 9364–9365 (2005)

    Article  Google Scholar 

  12. Cai, Z., Nakayama, Y., Shiono, T.: Living random copolymerization of propylene and norbornene with ansa-fluorenylamidodimethyltitanium complex: Synthesis of novel syndiotactic polypropylene-b-poly(propylene-ran-norbornene). Macromolecules 39, 2031–2033 (2006)

    Article  ADS  Google Scholar 

  13. He, L.-P., Liu, J.-Y., Li, Y.-G., Liu, S.-R., Li, Y.-S.: High-temperature living copolymerization of ethylene with norbornene by titanium complexes bearing bidentate [O, P] ligands. Macromolecules 42, 8566–8570 (2009)

    Article  Google Scholar 

  14. Liu, W., Zhang, K., Fan, H., Wang, W.-J., Li, B.-G., Zhu, S.: Living copolymerization of ethylene/1-octene with fluorinated FI-Ti catalyst. J. Polym. Sci. A: Polym. Chem. 51, 405–414 (2013)

    Article  ADS  Google Scholar 

  15. Mehdiabadi, S., Soares, J.B.P., Bilbao, D., Brinen, J.: Ethylene polymerization and ethylene/1-octene copolymerization with rac-dimethylsilylbis(indenyl)hafnium dimethyl using trioctyl aluminium and borate: A polymerization kinetics investigation. Macromolecules 46, 1312–1324 (2013)

    Article  ADS  Google Scholar 

  16. Pappalardo, D., Annunziata, L., Pellecchia, C.: Living ring-opening homo- and copolymerization of \(\varepsilon \) -caprolactone and L- and D, L-lactides by dimethyl(salicylaldiminato)aluminium compounds. Macromolecules 42, 6056–6062 (2009)

    Article  ADS  Google Scholar 

  17. Duda, A., Kowalski, A.: Thermodynamics and kinetics of ring-opening polymerization, Chapter 1. In: Dubois, P., Coulembier, O., Raquez, J.-M. (eds.) Handbook of Ring-Opening Polymerization, pp. 1–51. Wiley, Weinheim (2009)

    Chapter  Google Scholar 

  18. Mayo, F.R., Lewis, F.M.: Copolymerization. I. A basis for comparing the behavior of monomers in copolymerization; The copolymerization of styrene and methyl methacrylate. J. Am. Chem. Soc. 66, 1594–1601 (1944)

    Article  Google Scholar 

  19. Alfrey Jr., T., Goldfinger, G.: The mechanism of copolymerization. J. Chem. Phys. 12, 205–209 (1944)

    Article  ADS  Google Scholar 

  20. Fink, G., Richter, W.J.: Copolymerization parameters of metallocene-catalyzed copolymerizations, Part II. In: Brandrup, J., Immergut, E.H., Grulke, E.A. (eds.) Polymer Handbook, 4th edn, pp. 329–337. Wiley, New York (1999)

    Google Scholar 

  21. Andrieux, D., Gaspard, P.: Molecular information processing in nonequilibrium copolymerizations. J. Chem. Phys. 130, 014901 (2009)

    Article  ADS  Google Scholar 

  22. Gaspard, P., Andrieux, D.: Kinetics and thermodynamics of first-order Markov chain copolymerization. J. Chem. Phys. 141, 044908 (2014)

    Article  ADS  Google Scholar 

  23. Shu, Y.-G., Song, Y.-S., Ou-Yang, Z.-C., Li, M.: A general theory of kinetics and thermodynamics of steady-state copolymerization. J. Phys.: Condens. Matter 27, 235105 (2015)

    ADS  Google Scholar 

  24. Andrieux, D., Gaspard, P.: Information erasure in copolymers. EPL 103, 30004 (2013)

    Article  ADS  Google Scholar 

  25. Landauer, R.: Irreversibility and heat generation in the computing process. IBM J. Res. Dev. 5, 183–191 (1961)

    Article  MATH  MathSciNet  Google Scholar 

  26. Nicolis, G.: Introduction to Nonlinear Science. Cambridge University Press, Cambridge (1995)

    Book  Google Scholar 

  27. Kolomeisky, A.B., Fisher, M.E.: Molecular motors: a theorist’s perspective. Annu. Rev. Phys. Chem. 58, 675–695 (2007)

    Article  ADS  Google Scholar 

  28. Gaspard, P.: Force-velocity relation for copolymerization processes. New J. Phys. 17, 045016 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  29. Feller, W.: An Introduction to Probability Theory and Its Applications, vol. I, 3rd edn. Wiley, New York (1968)

    MATH  Google Scholar 

  30. Coleman, B.D., Fox, T.G.: General theory of stationary random sequences with applications to the tacticity of polymers. J. Polym. Sci. A 1, 3183–3197 (1963)

    Google Scholar 

  31. Coleman, B.D., Fox, T.G.: Multistate mechanism for homogeneous ionic polymerization. I. The diastereosequence distribution. J. Chem. Phys. 38, 1065–1075 (1963)

    Article  ADS  Google Scholar 

  32. Prigogine, I.: Introduction to Thermodynamics of Irreversible Processes. Charles C. Thomas Publishers, Springfield (1955)

    MATH  Google Scholar 

  33. Schnakenberg, J.: Network theory of microscopic and macroscopic behavior of master equation systems. Rev. Mod. Phys. 48, 571–585 (1976)

    Article  ADS  MathSciNet  Google Scholar 

  34. Nicolis, G.: Irreversible thermodynamics. Rep. Prog. Phys. 42, 225–268 (1979)

    Article  ADS  Google Scholar 

  35. Jiang, D.-Q., Qian, M., Qian, M.-P.: Mathematical Theory of Nonequilibrium Steady States. Springer, Berlin (2004)

    Book  MATH  Google Scholar 

  36. Bennett, C.H.: Dissipation-error tradeoff in proofreading. Biosystems 11, 85–91 (1979)

    Article  Google Scholar 

  37. Andrieux, D., Gaspard, P.: Nonequilibrium generation of information in copolymerization processes. Proc. Natl. Acad. Sci. USA 105, 9516–9521 (2008)

    Article  ADS  Google Scholar 

  38. Gaspard, P.: Thermodynamics of information processing at the molecular scale. Eur. Phys. J. Special Topics 224, 825–838 (2015)

    Article  ADS  Google Scholar 

  39. Karlin, S., Taylor, H.M.: A First Course in Stochastic Processes. Academic Press, New York (1975)

    MATH  Google Scholar 

  40. Karlin, S., Taylor, H.M.: A Second Course in Stochastic Processes. Academic Press, New York (1981)

    MATH  Google Scholar 

  41. Gillespie, D.T.: A general method for numerically simulating the stochastic time evolution of coupled chemical reactions. J. Comput. Phys. 22, 403–434 (1976)

    Article  ADS  MathSciNet  Google Scholar 

  42. Gillespie, D.T.: Exact stochastic simulation of coupled chemical reactions. J. Phys. Chem. 81, 2340–2361 (1977)

    Article  Google Scholar 

  43. Satoh, K., Matsuda, M., Nagai, K., Kamigaito, M.: AAB-sequence living radical chain copolymerization of naturally occurring limonene with maleimide: an end-to-end sequence-regulated copolymer. J. Am. Chem. Soc. 132, 10003–10005 (2010)

    Article  Google Scholar 

  44. Lutz, J.-F., Ouchi, M., Liu, D.R., Sawamoto, M.: Sequence-controlled polymers. Science 341, 1238149 (2013)

    Article  Google Scholar 

  45. Roy, R.K., Meszynska, A., Laure, C., Charles, L., Verchin, C., Lutz, J.-F.: Design and synthesis of digitally encoded polymers that can be decoded and erased. Nat. Commun. 6, 7237 (2015)

    Article  ADS  Google Scholar 

  46. Ellis, R.S.: Entropy, Large Deviations, and Statistical Mechanics. Springer, New York (1985)

    Book  MATH  Google Scholar 

  47. Touchette, H.: The large deviation approach to statistical mechanics. Phys. Rep. 478, 1–69 (2009)

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgments

The author thanks David Andrieux and Yves Geerts for useful discussions. This research is financially supported by the Université Libre de Bruxelles, the FNRS-F.R.S., and the Belgian Federal Government under the Interuniversity Attraction Pole Project P7/18 “DYGEST”.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierre Gaspard.

Appendices

Appendix 1: Solving the Kinetic Equations

1.1 The Growth of 1st-order Markov Chains

The general method described in Sect. 2 to solve the master equations (5) have already been used in Ref. [22] to deduce Eqs. (41)–(43) with \(k=1\). In this case, the attachment and detachment rates only depend on the previously incorporated monomeric units \(w_{\pm m_l\vert m_{l-1}}\) and first-order Markov chains are generated.

1.2 The Growth of 2nd-order Markov Chains

Remarkably, the method of Ref. [22] extends mutatis mutandis to higher orders k. Since the equations become quite complicated, we start developing the method to the case \(k=2\) where the attachment and detachment rates depend on the last two monomeric units \(w_{\pm m_l\vert m_{l-1}m_{l-2}}\). Summing the master equations (5) over sequences \(m_1m_2\ldots m_{l-r}\) with \(r=0,1,2,\ldots \), we obtain a hierarchy of equations for the probabilities (7), (8), (9),.... This hierarchy is summarized with Eq. (14). Replacing with the particular solution and letting the wavenumber going to zero, the hierarchy reduces to Eq. (20). This hierarchy is divided by the constant \(\phi _0(\emptyset )\) to get the equations

$$\begin{aligned} ({\varvec{\mathsf A}}+{\varvec{\mathsf B}}-{\varvec{\mathsf C}})\cdot \pmb {\mu }=0 \end{aligned}$$
(94)

for the time-independent probabilities (22), (23), (24), etc. The first few equations of this hierarchy read:

$$\begin{aligned} r=1:\quad 0= & {} \sum _{m_{l-2}m_{l-1}} w_{+m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1}) \nonumber \\&+\, \sum _{m_{l-1}m_{l+1}} w_{-m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-1}m_{l}m_{l+1}) \nonumber \\&- \sum _{m_{l-2}m_{l-1}} w_{-m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1}m_{l}) \nonumber \\&- \sum _{m_{l-1}m_{l+1}} w_{+m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-1}m_{l}), \end{aligned}$$
(95)
$$\begin{aligned} r=2:\quad 0= & {} \sum _{m_{l-2}} w_{+m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1}) + \sum _{m_{l+1}} w_{-m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-1}m_{l}m_{l+1}) \nonumber \\&- \sum _{m_{l-2}} w_{-m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1}m_{l}) - \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-1}m_{l}), \nonumber \\ \end{aligned}$$
(96)
$$\begin{aligned} r=3:\quad 0= & {} w_{+m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1}) + \sum _{m_{l+1}} w_{-m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-2}m_{l-1}m_{l}m_{l+1}) \nonumber \\&- \bigg (w_{-m_l\vert m_{l-1}m_{l-2}} + \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}m_{l-1}}\bigg )\, \mu (m_{l-2}m_{l-1}m_{l}), \end{aligned}$$
(97)
$$\begin{aligned} r=4:\quad 0= & {} w_{+m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-3}m_{l-2}m_{l-1}) \nonumber \\&+ \sum _{m_{l+1}} w_{-m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-3}m_{l-2}m_{l-1}m_{l}m_{l+1}) \nonumber \\&- \bigg ( w_{-m_l\vert m_{l-1}m_{l-2}}+ \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}m_{l-1}}\bigg )\, \mu (m_{l-3}m_{l-2}m_{l-1}m_{l}), \\&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \vdots \nonumber \end{aligned}$$
(98)

Because of Eqs. (26), (27),..., summing Eq. (98) over \(m_{l-3}\) gives Eq. (97), summing Eq. (97) over \(m_{l-2}\) gives Eq. (96), and summing Eq. (96) over \(m_{l-1}\) gives Eq. (95). We notice that summing Eq. (95) over \(m_{l}\) gives \(0=0\) for \(r=0\). The key observation is that the equations for \(r=3,4,\ldots \) are all similar to each other and they can thus be solved with the assumption of factorization

$$\begin{aligned} \mu (m_1m_2m_3\ldots m_{l-2}m_{l-1}m_l)= & {} \mu (m_1\vert m_2 m_3) \, \mu (m_2\vert m_3 m_4)\nonumber \\&\ldots \mu (m_{l-2}\vert m_{l-1} m_{l}) \, \mu (m_{l-1}m_{l}), \end{aligned}$$
(99)

as for a 2nd-order Markov chain in terms of the \(M^3\) conditional probabilities \(\mu (m_{l-2}\vert m_{l-1} m_{l})\) and \(M^2\) tip probabilities \(\mu (m_{l-1}m_{l})\). Under this assumption, all the equations \(r=3,4,\ldots \) of the hierarchy reduce to the equation for \(r=3\). Therefore, the conditional and tip probabilities should satisfy the first three equations for \(r=1,2,3\), which become:

$$\begin{aligned} r=1:\qquad 0= & {} \sum _{m_{l-2}m_{l-1}} w_{+m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1}) \nonumber \\&+ \sum _{m_{l-1}m_{l+1}} w_{-m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-1}\vert m_{l}m_{l+1})\, \mu (m_{l}m_{l+1}) \nonumber \\&- \sum _{m_{l-1}}\bigg [\sum _{m_{l-2}} w_{-m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}\vert m_{l-1}m_{l})+ \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}m_{l-1}}\bigg ] \nonumber \\&\qquad \times \, \mu (m_{l-1}m_{l}), \nonumber \\&\end{aligned}$$
(100)
$$\begin{aligned} r=2:\qquad 0= & {} \sum _{m_{l-2}} w_{+m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1}) \nonumber \\&+ \sum _{m_{l+1}} w_{-m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-1}\vert m_{l}m_{l+1}) \, \mu (m_{l}m_{l+1})\nonumber \\&- \,\bigg [\sum _{m_{l-2}} w_{-m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}\vert m_{l-1}m_{l}) + \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}m_{l-1}}\bigg ]\nonumber \\&\qquad \times \, \mu (m_{l-1}m_{l}), \end{aligned}$$
(101)
$$\begin{aligned} r=3:\qquad 0= & {} w_{+m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1}) \nonumber \\&+ \,\mu (m_{l-2}\vert m_{l-1}m_{l})\Bigg [\sum _{m_{l+1}} w_{-m_{l+1}\vert m_{l}m_{l-1}}\, \mu (m_{l-1}\vert m_{l}m_{l+1})\, \mu (m_{l}m_{l+1}) \nonumber \\&- \,\bigg (w_{-m_l\vert m_{l-1}m_{l-2}}+ \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}m_{l-1}}\bigg ) \mu (m_{l-1}m_{l})\Bigg ]. \end{aligned}$$
(102)

We see that Eq. (101) should determine the tip probabilities \(\mu (m_{l-1}m_{l})\) and Eq. (102) the conditional probabilities \(\mu (m_{l-2}\vert m_{l-1}m_{l})\), if decoupling could be achieved between these quantities. This is where the partial velocities (40) are introduced. In the case \(k=2\), they read

$$\begin{aligned} v_{m_{l-2}m_{l-1}}\equiv & {} \sum _{m_l} w_{+m_l\vert m_{l-1}m_{l-2}}\nonumber \\&-\, \frac{1}{\mu (m_{l-2}m_{l-1})}\sum _{m_l} w_{-m_l\vert m_{l-1}m_{l-2}} \, \mu (m_{l-2}\vert m_{l-1}m_{l}) \, \mu (m_{l-1}m_{l}).\qquad \end{aligned}$$
(103)

We observe that the first and third terms in the bracket of Eq. (102) actually form the partial velocity \(v_{m_{l-1}m_{l}}\) multiplied by \(\mu (m_{l-1}m_{l})\). Hence, Eq. (102) gives the conditional probabilities as

$$\begin{aligned} \mu (m_{l-2}\vert m_{l-1} m_{l}) = \frac{w_{+m_l\vert m_{l-1}m_{l-2}}\, \mu (m_{l-2}m_{l-1})}{(w_{-m_l\vert m_{l-1}m_{l-2}}+v_{m_{l-1}m_{l}})\, \mu (m_{l-1}m_{l})}. \end{aligned}$$
(104)

Inserting this result into Eq. (103), we get the self-consistent equations (41) with \(k=2\) for the partial velocities:

$$\begin{aligned} v_{m_{l-2} m_{l-1}} = \sum _{m_l} \frac{w_{+m_l\vert m_{l-1}m_{l-2}}\, v_{m_{l-1}m_{l}}}{w_{-m_l\vert m_{l-1}m_{l-2}}+v_{m_{l-1}m_{l}}}. \end{aligned}$$
(105)

Summing Eq. (104) over \(m_{l-2}\) and using the normalization condition (37), we find the equations (42) with \(k=2\) for the tip probabilities:

$$\begin{aligned} \mu (m_{l-1} m_{l}) = \sum _{m_{l-2}} \frac{w_{+m_l\vert m_{l-1}m_{l-2}}}{w_{-m_l\vert m_{l-1}m_{l-2}}+v_{m_{l-1}m_{l}}} \, \mu (m_{l-2}m_{l-1}). \end{aligned}$$
(106)

We can check that Eq. (101) transformed with Eqs. (103)–(104) also yields Eq. (106), while Eq. (100) transformed with Eq. (104) becomes the trivial equation \(v=v\) with the mean growth velocity (39) here given by

$$\begin{aligned} v = \sum _{m_{l-2}m_{l-1}} v_{m_{l-2}m_{l-1}} \, \mu (m_{l-2}m_{l-1})= \sum _{m_{l-1}m_{l}} v_{m_{l-1}m_{l}} \, \mu (m_{l-1}m_{l}). \end{aligned}$$
(107)

Therefore, Eqs. (100) and (101) are satisfied once Eq. (102) is solved by Eqs. (104)-(106) in terms of the partial velocities (103), which play an essential role in providing a self-consistent solution.

1.3 The Growth of kth-order Markov Chains

The general case with attachment and detachment rates \(w_{\pm m_l\vert m_{l-1}\ldots m_{l-k}}\), depending on k previously incorporated monomeric units, can be solved with the same method. Now, the equations of the hierarchy (94) are similar to each other for \(r=k+1,k+2,\ldots \):

$$\begin{aligned} r=k+1:&\nonumber \\ 0= & {} w_{+m_l\vert m_{l-1}\ldots m_{l-k}} \, \mu (m_{l-k}\ldots m_{l-1}) \nonumber \\&+\,\sum _{m_{l+1}} w_{-m_{l+1}\vert m_{l}\ldots m_{l-k+1}} \, \mu (m_{l-k}\ldots m_{l-1}m_{l}m_{l+1}) \nonumber \\&-\bigg ( w_{-m_{l}\vert m_{l-1}\ldots m_{l-k}} + \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}\ldots m_{l-k+1}}\bigg )\, \mu (m_{l-k}\ldots m_{l-1}m_{l}), \end{aligned}$$
(108)
$$\begin{aligned} r=k+2:&\nonumber \\ 0= & {} w_{+m_l\vert m_{l-1}\ldots m_{l-k}} \, \mu (m_{l-k-1}m_{l-k}\ldots m_{l-1}) \nonumber \\&+\sum _{m_{l+1}} w_{-m_{l+1}\vert m_{l}\ldots m_{l-k+1}} \, \mu (m_{l-k-1}m_{l-k}\ldots m_{l-1}m_{l}m_{l+1}) \nonumber \\&-\bigg ( w_{-m_{l}\vert m_{l-1}\ldots m_{l-k}} + \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}\ldots m_{l-k+1}}\bigg )\, \mu (m_{l-k-1}m_{l-k}\ldots m_{l-1}m_{l}),\qquad \quad \,\, \\&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \vdots \nonumber \end{aligned}$$
(109)

so that they are solved with the assumption (36) that the Markov chain is of kth order. The equation (108) for \(r=k+1\) becomes

$$\begin{aligned} r= & {} k+1:\nonumber \\&0= w_{+m_l\vert m_{l-1}\ldots m_{l-k}} \, \mu (m_{l-k}\ldots m_{l-1})+\mu (m_{l-k}\vert m_{l-k+1}\ldots m_{l}) \nonumber \\&\qquad \times \, \Bigg [\sum _{m_{l+1}} w_{-m_{l+1}\vert m_{l}\ldots m_{l-k+1}} \, \mu (m_{l-k+1}\vert m_{l-k+2} \ldots m_{l+1})\, \mu (m_{l-k+2}\ldots m_{l+1}) \nonumber \\&\qquad -\bigg ( w_{-m_{l}\vert m_{l-1}\ldots m_{l-k}} + \sum _{m_{l+1}} w_{+m_{l+1}\vert m_{l}\ldots m_{l-k+1}}\bigg )\, \mu (m_{l-k+1}\ldots m_{l}) \Bigg ]. \end{aligned}$$
(110)

Thanks to the introduction of the partial velocities (40), Eq. (110) gives the expression (43) for the conditional probabilities \(\mu (m_{l-k}\vert m_{l-k+1}\ldots m_{l})\). On the one hand, the self-consistent equations (41) for the partial velocities are obtained by inserting Eq. (43) into the definition (40) for the partial velocities. On the other hand, Eq. (42) for the tip probabilities are given by summing Eq. (43) over \(m_{l-k}\) and using the normalization conditions (37).

Here also, we can check that the equations for \(r=1,2,\ldots ,k\) in the hierarchy are satisfied with the assumption (36) and Eqs. (40)–(43).

Appendix 2: Proof of the Fluctuation Relation

In order to prove the fluctuation relation (67), we should start from the master equation

$$\begin{aligned}&\frac{d}{dt}\, P_t(m_1\ldots m_l,\Delta \mathbf{N})\nonumber \\&\quad = w_{+m_l\vert m_{l-1}\ldots m_{l-k}} \, P_t(m_1\ldots m_{l-1},\Delta \mathbf{N}-\mathbf{1}_{m_{l-k}\ldots m_l}) \nonumber \\&\qquad +\,\sum _{m_{l+1}=1}^M w_{-m_{l+1}\vert m_{l}\ldots m_{l-k+1}} \, P_t(m_1\ldots m_{l+1},\Delta \mathbf{N}+\mathbf{1}_{m_{l-k+1}\ldots m_{l+1}}) \nonumber \\&\qquad -\, \left( w_{-m_{l}\vert m_{l-1}\ldots m_{l-k}} + \sum _{m_{l+1}=1}^M w_{+m_{l+1}\vert m_{l}\ldots m_{l-k+1}}\right) P_t(m_1\ldots m_{l},\Delta \mathbf{N}), \end{aligned}$$
(111)

for the probability \(P_t(m_1\ldots m_l,\Delta \mathbf{N})\) that the chain has the sequence \(m_1\ldots m_l\) with \(l\gg k\) at the time t and that the \(M^{k+1}\) counters of the different possible multiplets \(m_j\ldots m_{j+k}\) of length \(k+1\) take the values \(\Delta \mathbf{N} = \left\{ \Delta N_{m_j\ldots m_{j+k}}\right\} \). The notation

$$\begin{aligned} (\Delta \mathbf{N}\pm \mathbf{1}_{m_{i}\ldots m_{i+k}})_{m_j\ldots m_{j+k}}= \left\{ \begin{array}{ll} \Delta N_{m_j\ldots m_{j+k}}, &{}\text{ if } \quad m_j\ldots m_{j+k}\ne m_{i}\ldots m_{i+k} \\ \Delta N_{m_j\ldots m_{j+k}}\pm 1, &{}\text{ if } \quad m_j\ldots m_{j+k}=m_{i}\ldots m_{i+k} \end{array} \right. \qquad \end{aligned}$$
(112)

is used in Eq. (111).

On the one hand, we notice that the master equation (5) is recovered from Eq. (111) by summing over \(\Delta \mathbf{N}\). On the other hand, the probability distribution appearing in the fluctuation relation (67) is defined as

$$\begin{aligned} p_t(\Delta \mathbf{N}) \equiv \sum _{m_1\ldots m_l} P_t(m_1\ldots m_l,\Delta \mathbf{N}). \end{aligned}$$
(113)

In the long-time limit, the same factorization as in Eq. (35) is expected:

$$\begin{aligned} P_t(m_1\ldots m_{l-1}m_l,\Delta \mathbf{N})\simeq p_t(\Delta \mathbf{N}) \, \mu (m_1 \ldots m_{l-1} m_l). \end{aligned}$$
(114)

Therefore, the probability distribution (113) evolves in time according to

$$\begin{aligned} \frac{dp_t}{dt} = \hat{L} \, p_t \end{aligned}$$
(115)

with the linear operator

$$\begin{aligned} \hat{L}\equiv & {} \sum _{m_{l-k}\ldots m_l} \left[ w_{+m_l\vert m_{l-1}\ldots m_{l-k}} \, \mu (m_{l-k}\ldots m_{l-1}) \, \left( \hat{E}^{-}_{m_{l-k}\ldots m_l}-1\right) \right. \nonumber \\&\left. +\, w_{-m_l\vert m_{l-1}\ldots m_{l-k}} \, \mu (m_{l-k}\vert m_{l-k+1}\ldots m_l) \, \mu (m_{l-k+1}\ldots m_l) \, \left( \hat{E}^{+}_{m_{l-k}\ldots m_l}-1\right) \right] \nonumber \\ \end{aligned}$$
(116)

where

$$\begin{aligned} {\hat{E}}^{\pm }_{m_{l-k}\ldots m_l} f(\Delta \mathbf{N}) = f(\Delta \mathbf{N}\pm \mathbf{1}_{m_{l-k}\ldots m_l}). \end{aligned}$$
(117)

The cumulant generating function of the counters \(\Delta \mathbf{N}\) is defined as

$$\begin{aligned} Q(\pmb {\lambda }) \equiv \lim _{t\rightarrow \infty } -\frac{1}{t} \ln \left\langle \mathrm{e}^{-\pmb {\lambda }\cdot \Delta \mathbf{N}}\right\rangle _t \end{aligned}$$
(118)

in terms of the counting parameters \(\pmb {\lambda }=\{\lambda _{m_{l-k}\ldots m_l}\}\in {\mathbb R}^{M^{k+1}}\). The cumulant generating function can be obtained as the leading eigenvalue, \(\hat{L}_{\pmb {\lambda }}\,\varphi = -Q(\pmb {\lambda })\,\varphi \), for the modified operator

$$\begin{aligned} \hat{L}_{\pmb {\lambda }} \equiv \mathrm{e}^{-\pmb {\lambda }\cdot \Delta \mathbf{N}} \hat{L} \, \mathrm{e}^{+\pmb {\lambda }\cdot \Delta \mathbf{N}}, \end{aligned}$$
(119)

giving

$$\begin{aligned} Q(\pmb {\lambda })= & {} \sum _{m_{l-k}\ldots m_l} \Big [ w_{+m_l\vert m_{l-1}\ldots m_{l-k}} \, \mu (m_{l-k}\ldots m_{l-1}) \, \left( 1-\mathrm{e}^{-\lambda _{m_{l-k}\ldots m_l}}\right) \nonumber \\&+\, w_{-m_l\vert m_{l-1}\ldots m_{l-k}} \, \mu (m_{l-k}\vert m_{l-k+1}\ldots m_l) \, \mu (m_{l-k+1}\ldots m_l) \, \left( 1-\mathrm{e}^{+\lambda _{m_{l-k}\ldots m_l}}\right) \Big ].\nonumber \\ \end{aligned}$$
(120)

Since the sum extends over all the multiplets of length \(k+1\), the indices can be arbitrarily changed as \(m_{l-k}\ldots m_l \rightarrow m_j\ldots m_{j+k}\) for any integer j.

The net rates (52) or (54) are obtained by differentiating the cumulant generating function (120) with respect to the counting parameters:

$$\begin{aligned} \mathbf{J} = \frac{\partial Q}{\partial \pmb {\lambda }}\Big \vert _{\pmb {\lambda }=0} = \lim _{t\rightarrow \infty } \frac{1}{t} \langle \Delta \mathbf{N}\rangle _t. \end{aligned}$$
(121)

The remarkable property is that the cumulant generating function (120) obeys the symmetry relation

$$\begin{aligned} Q(\pmb {\lambda }) =Q(\mathbf{A}-\pmb {\lambda }), \end{aligned}$$
(122)

with respect to the affinities (53). Using large-deviation theory [46, 47], the fluctuation relation (67) is thus established.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gaspard, P. Growth and Dissolution of Macromolecular Markov Chains. J Stat Phys 164, 17–48 (2016). https://doi.org/10.1007/s10955-016-1532-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-016-1532-x

Keywords

Navigation