Skip to main content
Log in

Non-equilibrium processes in an unconserved network model with limited resources

  • Regular Article
  • Published:
The European Physical Journal Plus Aims and scope Submit manuscript

Abstract

We present a study of a peculiar form of network topology comprising of lanes connected via a junction, competing for particles in a reservoir of limited capacity with non-conserving dynamics. We exploit mean-field approximation to thoroughly analyze stationary dynamic properties such density profiles, phase boundaries, and phase diagrams. The steady-state properties have been studied by taking into account the time evolution of particle density. It is found that the ratio of the number of incoming and outgoing lanes from the network junction as well as that of the kinetic rates substantially influences the number of stationary phases and the complexity of the phase diagram concerning the increasing number of particles in the system. The maximal current phase can persist in a fragment or the complete lanes, when the number of incoming and outgoing lanes is equal, as opposed to unequal counterparts. For lower values of the total number of particles, only low density phase is achieved in the left subsystem. All the theoretical findings are supported by extensive Monte Carlo simulations and explained using simple physical arguments.

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

Similar content being viewed by others

Data Availability

The authors declare that the data supporting the findings of this study are available within the article.

References

  1. T. Chou, K. Mallick, R.K. Zia, Non-equilibrium statistical mechanics: from a paradigmatic model to biological transport. Rep. Prog. Phys. 74(11), 116601 (2011)

    Article  ADS  Google Scholar 

  2. C. Domb, Phase Transitions and Critical Phenomena (Elsevier, Amsterdam, The Netherlands, 2000)

    MATH  Google Scholar 

  3. D. Chowdhury, L. Santen, A. Schadschneider, Statistical physics of vehicular traffic and some related systems. Phys. Rep. 329(4–6), 199–329 (2000)

    Article  ADS  Google Scholar 

  4. R.K. Zia, J.J. Dong, B. Schmittmann, Modeling translation in protein synthesis with TASEP: a tutorial and recent developments. J. Stat. Phys. 144(2), 405–28 (2011)

    Article  ADS  MATH  Google Scholar 

  5. T. Chou, G. Lakatos, Clustered bottlenecks in mRNA translation and protein synthesis. Phys. Rev. Lett. 93(19), 198101 (2004)

    Article  ADS  Google Scholar 

  6. R. Lipowsky, S. Klumpp, T.M. Nieuwenhuizen, Random walks of cytoskeletal motors in open and closed compartments. Phys. Rev. Lett. 87(10), 108101 (2001)

    Article  ADS  Google Scholar 

  7. B. Hölldobler, E.O. Wilson, The Ants (Harvard University Press, Cambridge, MA, 1990)

    Book  Google Scholar 

  8. H.J. Hilhorst, C. Appert-Rolland, A multi-lane TASEP model for crossing pedestrian traffic flows. J. Stat. Mech. Theory Exp. 2012(06), P06009 (2012)

    Article  Google Scholar 

  9. C.T. MacDonald, J.H. Gibbs, A.C. Pipkin, Kinetics of biopolymerization on nucleic acid templates. Biopolym. Orig. Res. Biomol. 6(1), 1–25 (1968)

    Article  Google Scholar 

  10. C.T. MacDonald, J.H. Gibbs, Concerning the kinetics of polypeptide synthesis on polyribosomes. Biopolym. Orig. Res. Biomol. 7(5), 707–25 (1969)

    Article  Google Scholar 

  11. A. Schadschneider, D. Chowdhury, K. Nishinari, Stochastic Transport in Complex Systems: From Molecules to Vehicles (Elsevier, Amsterdam, The Netherlands, 2010)

    MATH  Google Scholar 

  12. T. Antal, A.G. Schütz, Asymmetric exclusion process with next-nearest-neighbor interaction: some comments on traffic flow and a nonequilibrium reentrance transition. Phys. Rev. E 62(1), 83 (2000)

    Article  ADS  Google Scholar 

  13. A.B. Kolomeisky, M.E. Fisher, Molecular motors: a theorist’s perspective. Annu. Rev. Phys. Chem. 5(58), 675–95 (2007)

    Article  ADS  Google Scholar 

  14. T. Midha, A.B. Kolomeisky, A.K. Gupta, Effect of interactions for one-dimensional asymmetric exclusion processes under periodic and bath-adapted coupling environment. J. Stat. Mech. Theory Exp. 2018(4), 043205 (2018)

    Article  MATH  Google Scholar 

  15. B. Derrida, An exactly soluble non-equilibrium system: the asymmetric simple exclusion process. Phys. Rep. 301(1–3), 65–83 (1998)

    Article  ADS  Google Scholar 

  16. J. Krug, Boundary-induced phase transitions in driven diffusive systems. Phys. Rev. Lett. 67(14), 1882 (1991)

    Article  ADS  Google Scholar 

  17. G. Schütz, E. Domany, Phase transitions in an exactly soluble one-dimensional exclusion process. J. Stat. Phys. 72(1), 277–96 (1993)

    Article  ADS  MATH  Google Scholar 

  18. S. Muhuri, L. Shagolsem, M. Rao, Bidirectional transport in a multispecies totally asymmetric exclusion-process model. Phys. Rev. E 84(3), 031921 (2011)

    Article  ADS  Google Scholar 

  19. A. Parmeggiani, T. Franosch, E. Frey, Phase coexistence in driven one-dimensional transport. Phys. Rev. Lett. 90(8), 086601 (2003)

    Article  ADS  Google Scholar 

  20. A. Parmeggiani, T. Franosch, E. Frey, Totally asymmetric simple exclusion process with Langmuir kinetics. Phys. Rev. E 70(4), 046101 (2004)

    Article  ADS  Google Scholar 

  21. M.R. Evans, R. Juhász, L. Santen, Shock formation in an exclusion process with creation and annihilation. Phys. Rev. E 68(2), 026117 (2003)

    Article  ADS  Google Scholar 

  22. V. Popkov, A. Rákos, R.D. Willmann, A.B. Kolomeisky, G.M. Schütz, Localization of shocks in driven diffusive systems without particle number conservation. Phys. Rev. E 67(6), 066117 (2003)

    Article  ADS  Google Scholar 

  23. A.K. Gupta, I. Dhiman, Asymmetric coupling in two-lane simple exclusion processes with Langmuir kinetics: phase diagrams and boundary layers. Phys. Rev. E 89(2), 022131 (2014)

    Article  ADS  Google Scholar 

  24. R. Jiang, R. Wang, Q.S. Wu, Two-lane totally asymmetric exclusion processes with particle creation and annihilation. Phys. A Stat. Mech. Appl. 375(1), 247–56 (2007)

    Article  Google Scholar 

  25. A. Gupta, A.K. Gupta, Particle creation and annihilation in an exclusion process on networks. J. Phys. A Math. Theor. 24(55), 105001 (2022)

    Article  ADS  Google Scholar 

  26. J. Brankov, N. Pesheva, N. Bunzarova, Totally asymmetric exclusion process on chains with a double-chain section in the middle: computer simulations and a simple theory. Phys. Rev. E 69(6), 066128 (2004)

    Article  ADS  Google Scholar 

  27. R. Wang, M. Liu, R. Jiang, Theoretical investigation of synchronous totally asymmetric exclusion processes on lattices with multiple-input-single-output junctions. Phys. Rev. E 77(5), 051108 (2008)

    Article  ADS  Google Scholar 

  28. E. Pronina, A.B. Kolomeisky, Theoretical investigation of totally asymmetric exclusion processes on lattices with junctions. J. Stat. Mech. Theory Exp. 2005(07), P07010 (2005)

    Article  MATH  Google Scholar 

  29. K. Fitzpatrick, M.D. Wooldridge, J.D. Blaschke, Urban Intersection Design Guide: Volume 1-Guidelines (Texas A &M Transportation Institute, Bryan, TX, 2005)

    Google Scholar 

  30. D. Chretien, F. Metoz, F. Verde, E. Karsenti, R.H. Wade, Lattice defects in microtubules: protofilament numbers vary within individual microtubules. J. Cell Biol. 117(5), 1031–40 (1992)

    Article  Google Scholar 

  31. E. Vasileva, S. Citi, The role of microtubules in the regulation of epithelial junctions. Tissue Barriers 6(3), 1539596 (2018)

    Article  Google Scholar 

  32. L.S. Goldstein, Kinesin molecular motors: transport pathways, receptors, and human disease. Proc. Natl. Acad. Sci. 98(13), 6999–7003 (2001)

    Article  ADS  Google Scholar 

  33. D.D. Hurd, W.M. Saxton, Kinesin mutations cause motor neuron disease phenotypes by disrupting fast axonal transport in drosophila. Genetics 144(3), 1075–85 (1996)

    Article  Google Scholar 

  34. A. Raguin, A. Parmeggiani, N. Kern, Role of network junctions for the totally asymmetric simple exclusion process. Phys. Rev. E 88(4), 042104 (2013)

    Article  ADS  Google Scholar 

  35. I. Neri, N. Kern, A. Parmeggiani, Modeling cytoskeletal traffic: an interplay between passive diffusion and active transport. Phys. Rev. Lett. 110(9), 098102 (2013)

    Article  ADS  Google Scholar 

  36. I. Neri, N. Kern, A. Parmeggiani, Exclusion processes on networks as models for cytoskeletal transport. New J. Phys. 15(8), 085005 (2013)

    Article  ADS  Google Scholar 

  37. Z.P. Cai, Y.M. Yuan, R. Jiang, K. Nishinari, Q.S. Wu, The effect of attachment and detachment on totally asymmetric exclusion processes with junctions. J. Stat. Mech. Theory Exp. 2009(02), P02050 (2009)

    Article  Google Scholar 

  38. I. Neri, N. Kern, A. Parmeggiani, Totally asymmetric simple exclusion process on networks. Phys. Rev. Lett. 107(6), 068702 (2011)

    Article  ADS  Google Scholar 

  39. E. Pronina, A.B. Kolomeisky, Two-channel totally asymmetric simple exclusion processes. J. Phys. A Math. Gen. 37(42), 9907 (2004)

    Article  ADS  MATH  Google Scholar 

  40. P. Greulich, L. Ciandrini, R.J. Allen, M.C. Romano, Mixed population of competing totally asymmetric simple exclusion processes with a shared reservoir of particles. Phys. Rev. E 85(1), 011142 (2012)

    Article  ADS  Google Scholar 

  41. D.A. Adams, B. Schmittmann, R.K. Zia, Far-from-equilibrium transport with constrained resources. J. Stat. Mech. Theory Exp. 2008(06), P06009 (2008)

    Article  Google Scholar 

  42. M. Ha, M. Den Nijs, Macroscopic car condensation in a parking garage. Phys. Rev. E 66(3), 036118 (2002)

    Article  ADS  Google Scholar 

  43. T.L. Blasius, N. Reed, B.M. Slepchenko, K.J. Verhey, Recycling of kinesin-1 motors by diffusion after transport. PloS ONE 8(9), e76081 (2013)

    Article  ADS  Google Scholar 

  44. L. Ciandrini, I. Neri, J.C. Walter, O. Dauloudet, A. Parmeggiani, Motor protein traffic regulation by supply-demand balance of resources. Phys. Biol. 11(5), 056006 (2014)

    Article  ADS  Google Scholar 

  45. L.J. Cook, R.K. Zia, B. Schmittmann, Competition between multiple totally asymmetric simple exclusion processes for a finite pool of resources. Phys. Rev. E 80(3), 031142 (2009)

    Article  ADS  Google Scholar 

  46. A.K. Verma, A.K. Gupta, Limited resources in multi-lane stochastic transport system. J. Phys. Commun. 2(4), 045020 (2018)

    Article  Google Scholar 

  47. A. Jindal, A.K. Gupta, Exclusion process on two intersecting lanes with constrained resources: symmetry breaking and shock dynamics. Phys. Rev. E 104(1), 014138 (2021)

    Article  ADS  Google Scholar 

  48. R.M. Corless, G.H. Gonnet, D.E. Hare, D.J. Jeffrey, D.E. Knuth, On the LambertW function. Adv. Comput. Math. 5(1), 329–59 (1996)

    Article  MATH  Google Scholar 

Download references

Acknowledgements

The first author thanks Council of Scientific & Industrial Research (CSIR), India, for financial support under File No: 09/1005(0023)/2018-EMR-I. A.K.G. acknowledges support from DST-SERB, Government of India (Grant No. CRG/2019/004669 & MTR/2019/000312).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Arvind Kumar Gupta.

Ethics declarations

Conflict of interest

The authors have declared that no competing interests exist.

Appendices

Appendix A

Here, we investigate the transient solution, i.e., the evolution in particle density by analyzing the master equation as stated in Eq. (10) and further utilize it to acquire the corresponding steady-state solution. To completely understand the first-order differential equation in Eq. (10), one needs to study its characteristics which are given by

$$\begin{aligned} \frac{d\rho }{dt}= & {} \Omega _D[Kz-(1+Kz)\rho ] , \end{aligned}$$
(38)
$$\begin{aligned} \frac{d\rho }{dx}= & {} \frac{\Omega _D[Kz-(1+Kz)\rho ]}{1-2\rho }, \end{aligned}$$
(39)

where \(Kz=\Omega _A^*/\Omega _D\). We consider the density profile \(\rho (x,t)\) that develops for an initial density step

$$\begin{aligned} \rho _0(x,0)={\left\{ \begin{array}{ll} \alpha ^*, &\quad 0\le x<0.5\\ 1-\beta , &\quad 0.5< x \le 1 \end{array}\right. } \end{aligned}$$
(40)

where the left (\(x=0\)) and the right \((x=1)\) boundaries are fixed at \(\alpha ^*\) and \(1-\beta\), respectively. Integrating Eq. (38) and utilizing the initial density profile given by Eq. (40), we obtain the solution \(\rho _I(x,t)\) given by

$$\begin{aligned} \displaystyle \rho _I(x,t)=\frac{Kz-\bigg ( \Big (Kz-(1+Kz)\rho _0(x,0)\Big )e^{-\Omega _D(1+Kz)t}\bigg )}{1+Kz}. \end{aligned}$$
(41)

For the study of Eq. (39), we introduce a rescaled density of the form

$$\begin{aligned} \sigma =\frac{(Kz+1)(2\rho -1)}{Kz-1}-1, \end{aligned}$$
(42)

where Langmuir isotherm \(\rho _l=Kz/(Kz+1)\) eventuate for \(\sigma =0\) and is similar to that in reference [20] . It is evident that the above equation is not defined for \(Kz=1\). Depending upon the values of Kz, different scenarios are possible. Therefore, we categorize our analysis into two cases, i) \(Kz=1\), and ii) \(Kz\ne 1\).

1.1 A1.\(\ Kz=1\)

Let us first discuss the case when \(Kz=1\), for which Eq. (38) and (39) simplify considerably and lead to

$$\begin{aligned} \frac{d\rho }{dt}=\Omega _D(1-2\rho )\quad \textrm{and} \quad \frac{d\rho }{dx}=\Omega _D. \end{aligned}$$
(43)

Solving the above equations, we obtain the general solution of Eq. (10) as

$$\begin{aligned} (1-2\rho )e^{2\Omega _D t}=f(\rho -\Omega _D x), \end{aligned}$$
(44)

where f is an arbitrary function to be calculated. Utilizing the initial and boundary conditions, we obtain three solutions:

$$\begin{aligned} \rho _{\alpha }(x)&= \Omega _D x +\alpha ^*, \end{aligned}$$
(45)
$$\begin{aligned} \rho _I(x,t)&= \frac{1-(1-2\rho _0)e^{-2\Omega _D t}}{2}, \end{aligned}$$
(46)
$$\begin{aligned} \rho _{\beta }(x)&= \Omega _D (x-1)+1-\beta , \end{aligned}$$
(47)

where \(\rho _{\alpha }(x)\) and \(\rho _\beta (x)\) denote the solutions satisfying the left and the right boundary conditions, respectively. Depending upon how \(\rho _{\alpha }(x)\), \(\rho _{I}(x,t)\) and \(\rho _{\beta }(x)\) can be matched, different scenarios for the density profile occur [20]. The density profile \(\rho _{\alpha }(x)\) is separated from \(\rho _{I}(x,t)\) at \(x_{\alpha }\) and \(\rho _{I}(x,t)\) is separated from \(\rho _{\beta }(x)\) at \(x_{\beta }\) whose values are given by:

$$\begin{aligned} x_{\alpha }&=\frac{1-(1-2\rho _0)e^{-2\Omega _D t}-2\alpha ^*}{2\Omega _D}, \end{aligned}$$
(48)
$$\begin{aligned} x_{\beta }&=\frac{1-(1-2\rho _0)e^{-2\Omega _D t}-2(1-\beta )}{2\Omega _D}+1. \end{aligned}$$
(49)

According to the relative positions of \(x_{\alpha }\) and \(x_{\beta }\), the density profile is obtained as:

  1. 1.

    When \(x_{\alpha }\le x_{\beta }\), the density profile is given by

    $$\begin{aligned} \rho (x,t)= {\left\{ \begin{array}{ll} \Omega _D x +\alpha ^*, &\quad 0\le x \le x_{\alpha }\\ \frac{1-(1-2\rho _0)e^{-2\Omega _D t}}{2}, &\quad x_{\alpha }\le x \le x_{\beta }\\ \Omega _D (x-1)+1-\beta , &\quad x_{\beta }\le x \le 1. \end{array}\right. } \end{aligned}$$
    (50)
  2. 2.

    When \(x_{\alpha }> x_{\beta }\), there is necessarily a density discontinued at the point \(x_w\) where the currents corresponding to \(\rho _{\alpha }(x)\) and \(\rho _{\beta }(x)\) matches. The density profile is expressed as

    $$\begin{aligned} \rho (x,t)= {\left\{ \begin{array}{ll} \Omega _D x +\alpha ^*, &\quad 0\le x \le x_{w}\\ \Omega _D (x-1)+1-\beta , &\quad x_{w} \le x \le 1 \end{array}\right. } \end{aligned}$$
    (51)

    where \(x_{w}=(\Omega _D-\gamma +\beta )/(2\Omega _D).\)

1.2 A2. \(Kz\ne 1\)

In this case, the transformation given by Eq. (42) is well defined and Eq. (39) in the rescaled form reduces to

$$\begin{aligned} \left( \frac{\sigma +1}{\sigma }\right) \frac{\partial \sigma }{\partial x}=\frac{\Omega _D (Kz+1)^2}{Kz-1}. \end{aligned}$$
(52)

Integrating the above equation yields

$$\begin{aligned} |\sigma (x)|\text {exp}(\sigma (x))=Y(x), \end{aligned}$$
(53)

where Y(x) is

$$\begin{aligned} Y(x)=|\sigma (x_0)|\text { exp}\left\{ \Omega _D \frac{(Kz+1)^2}{Kz-1}(x-x_0)+\sigma (x_0)\right\} , \end{aligned}$$
(54)

and \(x_0\) is the reference point. In particular, at the boundaries \(x_0\) is equal to 0 or 1, which gives

$$\begin{aligned} Y_{\alpha }(x)&=|\sigma (0)|\text { exp}\left\{ \Omega _D \frac{(Kz+1)^2}{Kz-1}x+\sigma (0)\right\} , \\ Y_{\beta }(x)&=|\sigma (1)|\text { exp}\left\{ \Omega _D \frac{(Kz+1)^2}{Kz-1}(x-1)+\sigma (1)\right\} . \end{aligned}$$

The rescaled Eq. (53) is known to have an explicit solution in the form of a special function known as Lambert W function [19, 20, 48] and can be written as

$$\begin{aligned} \sigma (x)={\left\{ \begin{array}{ll} W_{-1}(-Y(x)), &\quad \sigma (x)<-1 \\ W_0(-Y(x)), &\quad -1\le \sigma (x)<0\\ W_0(Y(x)), &\quad 0< \sigma (x). \end{array}\right. } \end{aligned}$$
(55)

Using the properties of the Lambert W function and the values of \(\alpha ^*\) and \(\beta\), a specific branch of the Lambert W function can be chosen to obtain the solution to Eq. (53). The rescaled solution satisfying the left boundary condition is the left rescaled solution denoted by \(\sigma _{\alpha }(x)\), while the one obeying the right boundary condition is represented by \(\sigma _{\beta }(x)\) called as the right rescaled solution. Therefore, the particle densities \(\rho _{\alpha }(x)\) and \(\rho _{\beta }(x)\) can be obtained by substituting back \(\sigma _{\alpha }(x)\) and \(\sigma _{\beta }(x)\) in Eq. (42). Employing the suitable solution to \(\sigma _\alpha (x)\) and \(\sigma _\beta (x)\), the complete solution is constructed from the possible combination of the three solutions, i.e.,

$$\begin{aligned} \rho (x,t)= {\left\{ \begin{array}{ll} \rho _{\alpha }(x), &{} 0\le x \le x_{\alpha }\\ \rho _I(x,t), &{} x_{\alpha }\le x \le x_{\beta }\\ \rho _{\beta }(x), &{} x_{\beta }\le x \le 1 \end{array}\right. } \end{aligned}$$
(56)

where \(\rho _I(x,t)\) is given by Eq. (41). The two solutions \(\rho _{\alpha }(x)\) and \(\rho _I(x,t)\) are matched at the position \(x_{\alpha }\) for which the current for both the solutions is equal. Similarly, the currents for \(\rho _{I}(x,t)\) and \(\rho _{\beta }(x)\) are equated to calculate the value of \(x_\beta\).

1.3 A2.1 Explicit solution

Analogous to the homogeneous single-lattice TASEP model with LK coupled to an infinite reservoir [19, 20], the solution \(\rho _\alpha (x)\) corresponding to the left boundary condition is stable only if \(\alpha ^*\le 0.5\). The entry rate \(0\le \alpha ^*\le 0.5\) implies that \(0\le \rho _\alpha (x)\le 0.5\). Utilizing rescaled density Eq. (42), one has \(\sigma _\alpha \in [-2Kz/(Kz-1),-1]\) for \(Kz>1\) and \(\sigma _\alpha \in [-1,-2Kz/(Kz-1)]\) for \(Kz<1\). Hence, the left rescaled solution is given as follows:

  1. (a)

    If \(Kz>1\), then

    $$\begin{aligned} \sigma _\alpha (x)=W_{-1}(-Y_\alpha (x)). \end{aligned}$$
    (57)
  2. (b)

    If \(Kz<1\), then

    $$\begin{aligned} \sigma _\alpha (x)={\left\{ \begin{array}{ll} W_0(Y_\alpha (x)),&\quad 0\le \alpha \le \rho _l\\ 0, &\quad \alpha =\rho _l\\ W_0{(-Y_\alpha (x))},&\quad \rho _l\le \alpha \le 0.5.\\ \end{array}\right. } \end{aligned}$$
    (58)

Similarly, the solution \(\rho _\beta (x)\) matching the right boundary is stable only for \(\beta \le 0.5\) and is always in high density regime (\(\rho _\beta (x)\ge 0.5\)). For the right rescaled solution \(\sigma _\beta (x)\), employing \(\beta \le 0.5\) and \(\rho _\beta (x)\ge 0.5\) transforms to \(\sigma _\beta (x) \in [-1, 2/(Kz-1)]\) if \(Kz>1\) and \(\sigma _\beta (x) \in [ 2/(Kz-1),-1]\) for \(Kz<1\). Hence, by choosing the suitable branch of the Lambert W function, one obtains

  1. (a)

    If \(Kz>1\), then

    $$\begin{aligned} \sigma _\beta (x)={\left\{ \begin{array}{ll} W_0(Y_\beta (x)), &\quad 0\le \beta \le 1-\rho _l\\ 0, &\quad \beta = 1-\rho _l\\ W_0(-Y_\beta (x)), &\quad 1-\rho _l\le \beta \le 0.5\\ \end{array}\right. } \end{aligned}$$
    (59)

    where \(\rho _l\) is the Langmuir isotherm.

  2. (b)

    If \(Kz<1\), then

    $$\begin{aligned} \sigma _\beta (x)=W_{-1}(-Y_\beta (x)). \end{aligned}$$
    (60)

For thorough exploration, it is essential to perceive the phase boundaries separating the different phases in the phase diagram, in addition to the particle densities. The following section supplies a different method to calculate the density profiles as well as the phase boundaries.

1.4 A2.2 Implicit solution

An alternative method for obtaining the particle densities \(\rho _{\alpha }(x)\) and \(\rho _\beta (x)\) is through direct integration of Eq. (39). This equation can be rewritten as

$$\begin{aligned} {dx}=\frac{1-2\rho }{\Omega _D[Kz-(1+Kz)\rho ]} d\rho . \end{aligned}$$
(61)

Upon integration, we have

$$\begin{aligned} \frac{1}{\Omega _D(1+Kz)}\left( 2\rho +\frac{Kz-1}{Kz+1} \ln \left| Kz-(1+Kz)\rho \right| \right) =x+const. \end{aligned}$$
(62)

In the low density situation, the density of the left end \(\alpha ^*\) is employed, to find the profile \(\rho _\alpha (x)\) through

$$\begin{aligned} \frac{1}{\Omega _D(1+Kz)}\Bigg (2(\rho _\alpha -\overline{\alpha })+\frac{Kz-1}{Kz+1} ln\left| \frac{ Kz-(1+Kz)\rho _\alpha }{ Kz-(1+Kz)\overline{\alpha } } \right| \Bigg )= x \end{aligned}$$
(63)

where \(\overline{\alpha }=\text {min}(\alpha ^*,0.5)\). The high density solution \(\rho _\beta (x)\) corresponding to the right boundary condition is obtained from

$$\begin{aligned} \frac{1}{\Omega _D(1+Kz)}\Bigg (2(1-\bar{\beta }-\rho _\beta )+\frac{Kz-1}{Kz+1}\ln \left| \frac{ Kz-(1+Kz)(1-\bar{\beta })}{ Kz-(1+Kz)\rho _\beta } \right| \Bigg )=1-x \end{aligned}$$
(64)

for \(\bar{\beta }=\text {min}(\beta ,0.5)\).

This implicit solution also retrieves the density solutions \(\rho _{\alpha }(x)\) and \(\rho _\beta (x)\) obtained for \(Kz=1\). Now, the solutions procured for the densities \(\rho _{\alpha }(x)\) and \(\rho _\beta (x)\) in the subsection A2 can be deployed in Eq. (56) to finally compute the overall density of the lattice. Both the implicit and the explicit solutions are equivalent and can be used interchangeably. For the sake of completeness, we will be utilizing the explicit solutions to calculate the density profiles while the implicit solution will provide the phase boundaries.

Appendix B

Density profiles corresponding to different phases of the phase diagrams are depicted in Fig. 5.

Fig. 5
figure 5

Typical density profiles for \(m=n,\ K=1,\ \Omega _D=0.1\) (a-c with \(\mu =2.5\) and d-j with \(\mu =200\)), \(m=2,\ n=1,\ K=1,\ \Omega _D=0.1\) (k-m with \(\mu =1\)), and \(m=2,\ n=1,\ K=3,\ \Omega _D=0.01\) (n-o with \(\mu =1\)). Solid blue curves and red symbols denote mean-field and Monte Carlo simulations, respectively

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gupta, A., Gupta, A.K. Non-equilibrium processes in an unconserved network model with limited resources. Eur. Phys. J. Plus 138, 108 (2023). https://doi.org/10.1140/epjp/s13360-023-03722-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjp/s13360-023-03722-x

Navigation