Skip to main content
Log in

A Model of Oscillatory Protein Dynamics in Bacteria

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

Abstract

Spatial oscillations of proteins in bacteria have recently attracted much attention. The cellular mechanism underlying these oscillations can be studied at molecular as well as at more macroscopic levels. We construct a minimal mathematical model with two proteins that is able to produce self-sustained regular pole-to-pole oscillations without having to take into account molecular details of the proteins and their interactions. The dynamics of the model is based solely on diffusion across the cell body and protein reactions at the poles, and is independent of stimuli coming from the environment. We solve the associated system of reaction–diffusion equations and perform a parameter scan to demonstrate robustness of the model for two possible sets of the reaction functions.

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

Similar content being viewed by others

References

  • Adams, D. W., & Errington, J. (2009). Bacterial cell division: assembly, maintenance and disassembly of the Z ring. Nat. Rev., Microbiol., 7, 642–653.

    Article  Google Scholar 

  • Arnold, V. I. (1980). Ordinary differential equations (1st ed.). Cambridge: MIT Press. Transl. and ed. by R.A. Silverman.

    MATH  Google Scholar 

  • Bulyha, I., Schmidt, C., Lenz, P., Jakovljevic, V., Höne, A., Maier, B., Hopper, M., & Søgaard-Andersen, L. (2009). Regulation of the type IV pili molecular machine by dynamic localization of two motor proteins. Mol. Microbiol., 74, 691–705.

    Article  Google Scholar 

  • Bulyha, I., Hot, E., Huntley, S., & Søgaard-Andersen, L. (2011). GTPases in bacterial cell polarity and signaling. Curr. Opin. Microbiol., 14, 726–733.

    Article  Google Scholar 

  • Cytrynbaum, E. N., & Marshall, B. D. L. (2007). A multistranded polymer model explains MinDE dynamics in E. coli cell division. Biophys. J., 93, 1134–1150.

    Article  Google Scholar 

  • Edelstein-Keshet, L. (2005). Mathematical models in biology. Philadelphia: SIAM.

    Book  MATH  Google Scholar 

  • Elowitz, M. B., Surette, M. G., Wolf, P.-E., Stock, J. B., & Leibler, S. (1999). Protein mobility in the cytoplasm of Escherichia coli. J. Bacteriol., 181, 197–203.

    Google Scholar 

  • Fange, D., & Elf, J. (2006). Noise induced min phenotypes in E. coli. PLoS Comput. Biol., 2, e80.

    Article  Google Scholar 

  • Gerdes, K., Howard, M., & Szardenings, F. (2010). Pushing and pulling in prokaryotic DNA segregation. Cell, 141, 927–942.

    Article  Google Scholar 

  • Gitai, Z., Dye, N. A., Reisenauer, A., Wachi, M., & Shapiro, L. (2005). MreB actin-mediated segregation of a specific region of a bacterial chromosome. Cell, 120, 329–341.

    Article  Google Scholar 

  • Guckenheimer, J., & Holmes, P. (1983). Nonlinear oscillations, dynamical systems, and bifurcations of vector fields. Berlin: Springer.

    MATH  Google Scholar 

  • Hastings, S. P., & Murray, J. D. (1975). The existence of oscillatory solutions in the Field–Noyes model for the Belousov–Zhabotinskii reaction. SIAM J. Appl. Math., 28, 678–688.

    Article  MathSciNet  Google Scholar 

  • Howard, M., Rutenberg, A. D., & de Vet, S. (2001). Dynamic compartmentalization of bacteria: accurate division in E. coli. Phys. Rev. Lett., 87(21), 278102.

    Article  Google Scholar 

  • Huang, K. C., Meir, Y., & Wingreen, N. S. (2003). Dynamic structures in Escherichia coli: spontaneous formation of MinE rings and MinD polar zones. Proc. Natl. Acad. Sci. USA, 100(22), 12724–12728.

    Article  Google Scholar 

  • Kruse, K. (2002). A dynamic model for determining the middle of Escherichia coli. Biophys. J., 82, 618–627.

    Article  Google Scholar 

  • Laub, M. T., Shapiro, L., & McAdams, H. H. (2007). Systems biology of caulobacter. Annu. Rev. Genet., 41, 429–441.

    Article  Google Scholar 

  • Lenz, P., & Søgaard-Andersen, L. (2011). Temporal and spatial oscillations in bacteria. Nat. Rev., Microbiol., 9, 565–577.

    Article  Google Scholar 

  • Leonardy, S., Bulyha, I., & Søgaard-Andersen, L. (2008). Reversing cells and oscillating proteins. Mol. BioSyst., 4, 1009–1014.

    Article  Google Scholar 

  • Leonardy, S., Miertzschke, M., Bulyha, I., Sperling, E., Wittinghofer, A., & Søgaard-Andersen, L. (2010). Regulation of dynamic polarity switching in bacteria by a Ras-like G-protein and its cognate GAP. EMBO J., 29, 2276–2289.

    Article  Google Scholar 

  • Lichtenberg, A. J., & Lieberman, M. A. (1992). Regular and chaotic dynamics (2nd ed.). Berlin: Springer.

    MATH  Google Scholar 

  • Loose, M., Fischer-Friedrich, E., Ries, J., Kruse, K., & Schwille, P. (2008). Spatial regulators for bacterial cell division self-organize into surface waves in vitro. Science, 320, 789–792.

    Article  Google Scholar 

  • Loose, M., Kruse, K., & Schwille, P. (2011). Protein self-organization: lessons from the min system. Annu. Rev. Biophys., 40, 315–336.

    Article  Google Scholar 

  • López, D., & Kolter, R. (2010). Functional microdomains in bacterial membranes. Genes Dev., 24, 1893–1902.

    Article  Google Scholar 

  • Lutkenhaus, J. (2007). Assembly dynamics of the bacterial MinCDE system and spatial regulation of the Z ring. Annu. Rev. Biochem., 76, 539–562.

    Article  Google Scholar 

  • Matsumoto, K., Kusaka, J., Nishibori, A., & Hara, H. (2006). Lipid domains in bacterial membranes. Mol. Microbiol., 61, 1110–1117.

    Article  Google Scholar 

  • Meacci, G., Ries, J., Fischer-Friedrich, E., Kahya, N., Schwille, P., & Kruse, K. (2006). Mobility of min-proteins in Escherichia coli measured by fluorescence correlation spectroscopy. Phys. Biol., 3, 255–263.

    Article  Google Scholar 

  • Meinhardt, H., & de Boer, P. A. J. (2001). Pattern formation in Escherichia coli: a model for the pole-to-pole oscillations of min proteins and the localization of the division site. Proc. Natl. Acad. Sci. USA, 98, 14202–14207.

    Article  Google Scholar 

  • Montero Llopis, P., Jackson, A. F., Sliusarenko, O., Surovtsev, I., Heinritz, J., Emonet, T., & Jacobs-Wagner, C. (2010). Spatial organization of the flow of genetic information in bacteria. Nature, 466, 77–81.

    Article  Google Scholar 

  • Nan, B., Mauriello, E. M. F., Wong, A., Sun, I. H., & Zusman, D. R. (2010). A multi-protein complex from Myxococcus xanthus required for bacterial gliding motility. Mol. Microbiol., 76, 1539–1554.

    Article  Google Scholar 

  • Nelson, W. J. (2003). Adaptation of core mechanisms to generate cell polarity. Nature, 422, 766–774.

    Article  Google Scholar 

  • Nevo-Dinur, K., Nussbaum-Shochat, A., Ben-Yehuda, S., & Amster-Choder, O. (2011). Translation-independent localization of mRNA in E. coli. Science, 331, 1081–1084.

    Article  Google Scholar 

  • Prigogine, I., & Nicolis, G. (1967). On symmetry-breaking instabilities in dissipative systems. J. Chem. Phys., 46, 3542–3551.

    Article  Google Scholar 

  • Prigogine, I., & Lefever, R. (1968). On symmetry-breaking instabilities in dissipative systems (II). J. Chem. Phys., 46, 1695–1700.

    Article  Google Scholar 

  • Ringgaard, S., van Zon, J., Howard, M., & Gerdes, K. (2009). Movement and equipositioning of plasmids by ParA filament disassembly. Proc. Natl. Acad. Sci. USA, 106, 19369–19374.

    Article  Google Scholar 

  • Rinzel, J., & Troy, W. C. (1982). Bursting phenomena in a simplified Oregonator flow system model. J. Chem. Phys., 76, 1775–1789.

    Article  MathSciNet  Google Scholar 

  • Seydel, R. (2010). Practical bifurcation and stability analysis (3rd ed.). New York: Springer.

    Book  MATH  Google Scholar 

  • Shapiro, L., McAdams, H. H., & Losick, R. (2009). Why and how bacteria localize proteins. Science, 326, 1225–1228.

    Article  Google Scholar 

  • Teleman, A. A., Graumann, P. L., Lin, D. C.-H., Grossman, A. D., & Losick, R. (1998). Chromosome arrangement within a bacterium. Curr. Biol., 8, 1102–1109.

    Article  Google Scholar 

  • Viollier, P. H., Thanbichler, M., McGrath, P. T., West, L., Meewan, M., McAdams, H. H., & Shapiro, L. (2004). Rapid and sequential movement of individual chromosomal loci to specific subcellular locations during bacterial DNA replication. Proc. Natl. Acad. Sci. USA, 101, 9257–9262.

    Article  Google Scholar 

  • Verhulst, F. (1990). Nonlinear differential equations and dynamical systems. Berlin: Springer.

    Book  MATH  Google Scholar 

  • Wang, X., Liu, X., Possoz, C., & Sherratt, D. J. (2006). The two Escherichia coli chromosome arms locate to separate cell halves. Genes Dev., 20, 1727–1731.

    Article  Google Scholar 

  • Weiner, R., Schmitt, B. A., & Podhaisky, H. (1997). ROWMAP—a ROW-code with Krylov techniques for large stiff ODEs. Appl. Numer. Math., 25(2–3), 303–319.

    Article  MathSciNet  MATH  Google Scholar 

  • Zhang, Y., Franco, M., Ducret, A., & Mignot, T. (2010). A bacterial Ras-like small GTP-binding protein and its cognate GAP establish a dynamic spatial polarity axis to control directed motility. PLoS Biol., 8, e1000430.

    Article  Google Scholar 

Download references

Acknowledgements

We would like to thank Thorsten Raasch for work done at the early stages of the project, Steffen Beck for advice on the MATLAB code, and Bruno Eckhardt for helpful discussions. This work has been supported by the Center for Synthetic Microbiology (SYNMIKRO) in Marburg, promoted by the LOEWE Excellence Program of the state of Hessen, Germany.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Stephan Dahlke.

Appendix

Appendix

This section describes the numerical algorithm used in the simulations and the parameter scan.

Algorithm

The ultimate goal is to check for robustness of the model, in other words, to determine the range of the values of the parameters where the dynamical systems with reaction functions (35) and (36) display qualitatively the desired behavior (oscillations of the protein concentrations at a fixed position in space). To automatize computation, the parameters are input from an array p. A parameter search involves scanning the admissible values of the array p and selecting those leading to oscillatory solutions. In our simulations the array has 3 entries.

Corollary 5 gives us a criterion for existence of an unstable equilibrium (9), which means that the chosen equilibrium c in (50) for the approximating ODE system given by (37), (46)–(45) is unstable, also. It is therefore enough to check whether the solution of the discrete problem is periodic whenever the matrix V given by (13) has an eigenvalue with positive real part. Due to the nonlinear terms the orbits starting from a perturbation of c may bend and not tend to ∞. Because of mass conservation (48) and the positivity of the variables, guaranteed by the form of on- and off-rates, whenever the eigenvalues of V are positive, the orbit starting near c may either circle around the equilibrium or finish at a different stable equilibrium c′. However, if a limit cycle around c exists, the orbit starting from a perturbation of c may never join it, but continue moving around it (Verhulst 1990). A periodic solution of the ODE system is thus asymptotically stable, with the orbit approaching the limit cycle arbitrarily close.

The period of oscillation of the solution is approximated numerically using a Poincaré section. The Poincaré section Ω is a (n−1)-dimensional manifold transversal to the limit cycle. To determine the length of the cycle we search for the time points t where an orbit starting from the given initial value passes through Ω, always entering from the same direction (Seydel 2010; Verhulst 1990). Since in the laboratory experiment, the motion of the bacterium is observed on a specific time scale (say, 45 min to 1 h), it is enough to check that the numerically computed solution sol oscillates within a certain time interval [0,T max] for a fixed T max. To find the period of oscillation, we need at least one period occurring in [0,T max], since periods longer than the period of observation are not registered in the experiment. Then we take the average over the computed periods. We have chosen an upper limit of T max=500 time units.

To minimize errors arising from transient irregular oscillations at the beginning of the orbit, we begin our search from the endpoint at T max, that is, we inspect backward the numerically computed solution sol(j)↦sol(T maxj).

We search for the range of the values of p leading to oscillations, using the following pseudocode:

  1. P1.

    set p.

  2. P2.

    compute eigenvalues of V.

  3. P3.

    if a positive eigenvalue exists, goto P4, else goto P1.

  4. P4.

    solve ODE.

  5. P5.

    if the ODE solution sol is periodic, goto P6 else goto P1.

  6. P6.

    check for period of solution (optional).

Since the Jacobian of the ODE system (37), (46)–(45) linearized at c has eigenvalues whose absolute value grows with increasing the grid size n, the problem is stiff.

A linearly-implicit solver scheme must be used to achieve numerical precision. In particular, we have used a Krylov-ROW method implemented in ROWMAP, a solver developed for solving large stiff initial value problems in (Weiner et al. 1997). Its advantage lies in the “matrix-free” method, which does not require explicit computation of the full Jacobian, and solves the equations by a multiple Arnoldi process. Furthermore, the time-step used is variable because it is controlled by local error estimates for increased efficiency. Because the ultimate goal is to perform a parameter scan, these features are an advantageous way to speed up the computations. The current version of ROWMAP can be found at http://numerik.mathematik.uni-halle.de/forschung/software/.

Checking for Ghost Solutions

To verify the robustness of our model, we have to exclude ghost solutions arising from discretization errors. In our model, these could appear whenever the Jacobian of the discrete ODE system (37), (46)–(45) has an eigenvalue with positive real part, but the matrix A −1 V from Corollary 5 does not have an eigenvalue with positive real part.

It has been numerically verified that for reaction rates (35) and (36) inside the computed parameter ranges the eigenvalues λ with \(\operatorname{Re}\lambda>0\) of the Jacobian of the right-hand side of the discrete ODE system (37), (46)–(45) never correspond to eigenvalues ρ of A −1 V with \(\operatorname{Re}\rho<0\). Thus, “ghost” oscillations arising from discretization errors do not appear. In fact, for the Jacobian matrix, \(\operatorname{Re}\lambda\) crosses into the positive only after \(\operatorname{Re}\rho\) when one parameter a i is varied and the other two are fixed; see, for example, Fig. 9.

Fig. 9
figure 9

Plot of \(\operatorname{Re}\rho\) and smallest non zero eigenvalue λ of the Jacobian vs. a 3 for different grid size n and Choice A with a 1=0.4,a 2=0.8

Rights and permissions

Reprints and permissions

About this article

Cite this article

Rashkov, P., Schmitt, B.A., Søgaard-Andersen, L. et al. A Model of Oscillatory Protein Dynamics in Bacteria. Bull Math Biol 74, 2183–2203 (2012). https://doi.org/10.1007/s11538-012-9752-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-012-9752-y

Keywords

Navigation