Skip to main content

Transient Memory in Gene Regulation

  • Conference paper
  • First Online:
Book cover Computational Methods in Systems Biology (CMSB 2019)

Part of the book series: Lecture Notes in Computer Science ((LNBI,volume 11773))

Included in the following conference series:

Abstract

The expression of a gene is characterised by its transcription factors and the function processing them. If the transcription factors are not affected by gene products, the regulating function is often represented as a combinational logic circuit, where the outputs (product) are determined by current input values (transcription factors) only, and are hence independent on their relative arrival times. However, the simultaneous arrival of transcription factors (TFs) in genetic circuits is a strong assumption, given that the processes of transcription and translation of a gene into a protein introduce intrinsic time delays and that there is no global synchronisation among the arrival times of different molecular species at molecular targets.

In this paper, we construct an experimentally implementable genetic circuit with two inputs and a single output, such that, in presence of small delays in input arrival, the circuit exhibits qualitatively distinct observable phenotypes. In particular, these phenotypes are long lived transients: they all converge to a single value, but so slowly, that they seem stable for an extended time period, longer than typical experiment duration. We used rule-based language to prototype our circuit, and we implemented a search for finding the parameter combinations raising the phenotypes of interest.

The behaviour of our prototype circuit has wide implications. First, it suggests that GRNs can exploit event timing to create phenotypes. Second, it opens the possibility that GRNs are using event timing to react to stimuli and memorise events, without explicit feedback in regulation. From the modelling perspective, our prototype circuit demonstrates the critical importance of analysing the transient dynamics at the promoter binding sites of the \({\mathsf {DNA}}\), before applying rapid equilibrium assumptions.

Tatjana Petrov’s research was supported by SNSF Advanced Postdoc. Mobility Fellowship grant number \(P300P2\_161067\), the Ministry of Science, Research and the Arts of the state of Baden-Württemberg, and the DFG Centre of Excellence 2117 ‘Centre for the Advanced Study of Collective Behaviour’ (ID: 422037984). Claudia Igler is the recipient of a DOC Fellowship of the Austrian Academy of Sciences. Thomas A. Henzinger’s research was supported in part by the Austrian Science Fund (FWF) under grant Z211-N23 (Wittgenstein Award).

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Models used in this paper will count 23 and 6 distinct \({\mathsf {DNA}}\) binding configurations.

  2. 2.

    The ratio between the binding and unbinding rate.

  3. 3.

    Notice that these six parameter combinations are different than those used for the model with looping.

References

  1. Bintu, L., et al.: Transcriptional regulation by the numbers: applications. Curr. Opin. Genet. Devel. 15(2), 125–135 (2005)

    Article  Google Scholar 

  2. Bolouri, H., Davidson, E.H.: Modeling transcriptional regulatory networks. Bioessays: News Rev. Mol. Cell. Devel. Biol. 24(12), 1118–1129 (2002)

    Article  Google Scholar 

  3. Boutillier, P., et al.: The Kappa platform for rule-based modeling. Bioinformatics 34(13), i583–i592 (2018)

    Article  Google Scholar 

  4. Brim, L., Češka, M., Dražan, S., Šafránek, D.: Exploring parameter space of stochastic biochemical systems using quantitative model checking. In: Sharygina, N., Veith, H. (eds.) CAV 2013. LNCS, vol. 8044, pp. 107–123. Springer, Heidelberg (2013). https://doi.org/10.1007/978-3-642-39799-8_7

    Chapter  Google Scholar 

  5. Brunner, M., Bujard, H.: Promoter recognition and promoter strength in the Escherichia coli system. EMBO J. 6(10), 3139 (1987)

    Article  Google Scholar 

  6. Buchler, N.E., Gerland, U., Hwa, T.: On schemes of combinatorial transcription logic. Proc. Nat. Acad. Sci. 100(9), 5136–5141 (2003)

    Article  Google Scholar 

  7. Byers, R., Hansell, R., Madras, N.: Stability-like properties of population models. Theor. Popul. Biol. 42(1), 10–34 (1992)

    Article  MathSciNet  Google Scholar 

  8. Casadesús, J., D’Ari, R.: Memory in bacteria and phage. BioEssays 24(6), 512–518 (2002)

    Article  Google Scholar 

  9. Češka, M., Dannenberg, F., Paoletti, N., Kwiatkowska, M., Brim, L.: Precise parameter synthesis for stochastic biochemical systems. Acta Informatica 54(6), 589–623 (2017)

    Article  MathSciNet  Google Scholar 

  10. Česka, M., Šafránek, D., Dražan, S., Brim, L.: Robustness analysis of stochastic biochemical systems. PLoS ONE 9(4), e94553 (2014)

    Article  Google Scholar 

  11. Danos, V., Feret, J., Fontana, W., Harmer, R., Krivine, J.: Rule-based modelling of cellular signalling. In: Caires, L., Vasconcelos, V.T. (eds.) CONCUR 2007. LNCS, vol. 4703, pp. 17–41. Springer, Heidelberg (2007). https://doi.org/10.1007/978-3-540-74407-8_3

    Chapter  Google Scholar 

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

    Article  Google Scholar 

  13. Guet, C.C., Elowitz, M.B., Hsing, W., Leibler, S.: Combinatorial synthesis of genetic networks. Science 296(5572), 1466–1470 (2002)

    Article  Google Scholar 

  14. Harada, Y., Funatsu, T., Murakami, K., Nonoyama, Y., Ishihama, A., Yanagida, T.: Single-molecule imaging of RNA polymerase-dna interactions in real time. Biophys. J. 76(2), 709–715 (1999)

    Article  Google Scholar 

  15. Hermsen, R., Tans, S., Ten Wolde, P.R.: Transcriptional regulation by competing transcription factor modules. PLoS Comput. Biol. 2(12), e164 (2006)

    Article  Google Scholar 

  16. Jacob, F.: Evolution and tinkering. Science 196(4295), 1161–1166 (1977)

    Article  Google Scholar 

  17. Jacob, F., Monod, J.: Genetic regulatory mechanisms in the synthesis of proteins. J. Mol. Biol. 3(3), 318–356 (1961)

    Article  Google Scholar 

  18. Kurtz, T.G.: Solutions of ordinary differential equations as limits of pure jump Markov processes. J. Appl. Probab. 7(1), 49–58 (1970)

    Article  MathSciNet  Google Scholar 

  19. Kurtz, T.G.: Limit theorems for sequences of jump Markov processes approximating ordinary differential processes. J. Appl. Probab. 8(2), 344–356 (1971)

    Article  MathSciNet  Google Scholar 

  20. Kwok, R.: Five hard truths for synthetic biology. Nature 463(7279), 288–290 (2010)

    Article  Google Scholar 

  21. Levin, D.A., Peres, Y., Wilmer, E.L.: Markov Chains and Mixing Times. American Mathematical Society (2009)

    Google Scholar 

  22. Levine, J.H., Lin, Y., Elowitz, M.B.: Functional roles of pulsing in genetic circuits. Science 342(6163), 1193–1200 (2013)

    Article  Google Scholar 

  23. Lin, Y., Sohn, C.H., Dalal, C.K., Cai, L., Elowitz, M.B.: Combinatorial gene regulation by modulation of relative pulse timing. Nature 527(7576), 54–58 (2015)

    Article  Google Scholar 

  24. Marchisio, M.A., Stelling, J.: Automatic design of digital synthetic gene circuits. PLoS Comput. Biol. 7(2), e1001083 (2011)

    Article  Google Scholar 

  25. McAdams, H.H., Arkin, A.: It‘sa noisy business! genetic regulation at the nanomolar scale. Trends Genet. 15(2), 65–69 (1999)

    Article  Google Scholar 

  26. Milo, R., Jorgensen, P., Moran, U., Weber, G., Springer, M.: Bionumbers–the database of key numbers in molecular and cell biology. Nucleic Acids Res. 38(suppl 1), D750–D753 (2010)

    Article  Google Scholar 

  27. Müller-hill, B.: Lac Operon. Wiley Online Library (1996)

    Google Scholar 

  28. Myers, C.J.: Engineering Genetic Circuits. CRC Press (2009)

    Google Scholar 

  29. Nashun, B., Hill, P.W., Hajkova, P.: Reprogramming of cell fate: epigenetic memory and the erasure of memories past. EMBO J. 34(10), 1296–1308 (2015)

    Article  Google Scholar 

  30. Ptashne, M.: A Genetic Switch: Phage Lambda Revisited, vol. 3. Cold Spring Harbor Laboratory Press Cold Spring Harbor, New York (2004)

    Google Scholar 

  31. Radulescu, O., Swarup Samal, S., Naldi, A., Grigoriev, D., Weber, A.: Symbolic dynamics of biochemical pathways as finite states machines. In: Roux, O., Bourdon, J. (eds.) CMSB 2015. LNCS, vol. 9308, pp. 104–120. Springer, Cham (2015). https://doi.org/10.1007/978-3-319-23401-4_10

    Chapter  Google Scholar 

  32. Saiz, L., Rubi, J.M., Vilar, J.M.G.: Inferring the in vivo looping properties of DNA. Proc. Nat. Acad. Sci. U.S.A. 102(49), 17642–17645 (2005)

    Article  Google Scholar 

  33. Saiz, L., Vilar, J.M.: DNA looping: the consequences and its control. Curr. Opin. Struct. Biol. 16(3), 344–350 (2006). Nucleic acids/Sequences and topology Anna Marie Pyle and Jonathan Widom/Nick V Grishin and Sarah A Teichmann

    Google Scholar 

  34. Saiz, L., Vilar, J.M.: Stochastic dynamics of macromolecular-assembly networks. Mol. Syst. Biol. 2(1) (2006)

    Google Scholar 

  35. Santillán, M., Mackey, M.C.: Why the lysogenic state of phage \(\lambda \) is so stable: a mathematical modeling approach. Biophys. J. 86(1), 75–84 (2004)

    Article  Google Scholar 

  36. Schnoerr, D., Sanguinetti, G., Grima, R.: Approximation and inference methods for stochastic biochemical kinetics–a tutorial review. J. Phys. A: Math. Theor. 50(9), 093001 (2017)

    Article  MathSciNet  Google Scholar 

  37. Segal, E., Widom, J.: From dna sequence to transcriptional behaviour: a quantitative approach. Nat. Rev. Genet. 10(7), 443–456 (2009)

    Article  Google Scholar 

  38. Shea, M.A., Ackers, G.K.: The OR control system of bacteriophage lambda: a physical-chemical model for gene regulation. J. Mol. Biol. 181(2), 211–230 (1985)

    Article  Google Scholar 

  39. Vilar, J.M., Leibler, S.: DNA looping and physical constraints on transcription regulation. J. Mol. Biol. 331(5), 981–989 (2003)

    Article  Google Scholar 

  40. Vilar, J.M., Saiz, L.: Dna looping in gene regulation: from the assembly of macromolecular complexes to the control of transcriptional noise. Curr. Opin. Genet. Devel. 15(2), 136–144 (2005)

    Article  Google Scholar 

  41. Vilar, J.M., Saiz, L.: Suppression and enhancement of transcriptional noise by DNA looping. Phys. Rev. E 89(6), 062703 (2014)

    Article  Google Scholar 

  42. Vivek-Ananth, R., Samal, A.: Advances in the integration of transcriptional regulatory information into genome-scale metabolic models. Biosystems 147, 1–10 (2016)

    Article  Google Scholar 

  43. Yuh, C.H., Bolouri, H., Davidson, E.H.: Cis-regulatory logic in the endo16 gene: switching from a specification to a differentiation mode of control. Devel. (Cambridge, England) 128(5), 617–629 (2001)

    Google Scholar 

  44. Zeller, R.W., Griffith, J.D., Moore, J.G., Kirchhamer, C.V., Britten, R.J., Davidson, E.H.: A multimerizing transcription factor of sea urchin embryos capable of looping DNA. Proc. Nat. Acad. Sci. 92(7), 2989–2993 (1995)

    Article  Google Scholar 

  45. Zhu, R., Salahub, D.: Delay stochastic simulation of single-gene expression reveals a detailed relationship between protein noise and mean abundance. FEBS Lett. 582(19), 2905–2910 (2008)

    Article  Google Scholar 

Download references

Acknowledgements

We are very grateful to Moritz Lang, Tiago Paixao and Jakob Ruess, for their feedback during the manuscript preparation.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tatjana Petrov .

Editor information

Editors and Affiliations

Appendices

Appendix 1.A Parameter Values

Table 1 lists the parameter ranges used for our case study example. We next explain the choice of each of the parameters with respect to their biological context.

Table 1. Parameter combinations tested in the model with looping.

1.A.1   Stochastic Scaling Constant

The stochastic scaling of rates and concentrations is done with a standard scaling rate for E. coli  cell \(N= 10^9\) [26].

1.A.2   Protein Production and Degradation

The protein production is taken 0.5 molec.s\(^{-1}\) ([41], caption of Fig. 2) and the degradation rate is taken 0.001 s\(^{-1}\) (corresponding to the halflife of \(12\min \), consistent with [26]).

1.A.3   RNAP Rates

Table 2. Parameter combinations tested for the model without looping.
Table 3. Simulation parameters: all models were run for three different delays and in six different regimes.

On rate, off rate and number of RNAP molecules are consistent with the orders of values reported in [5, 14, 35].

1.A.4   Activator

The activation mechanism is inspired by the activation of the \(\mathsf {PRM}\) promoter in the lysogenic state by protein \({\mathsf {CI}}\) in the regulation of \(\lambda \)-phage: \({\mathsf {CI}}\) competes with \({\mathsf {Cro}}\) to bind to the promoter sites, and, when bound, it recruits \({\mathsf {RNAP}}\) (increases \(\mathsf {PRM}\) activity). The mechanism with looping, explained at mechanistic detail level in [35], contains three left and three right operators, leading to 1200 different DNA binding states. We model a mechanism with two states for the activator without looping (‘bound’ or ‘not bound’) and with four binding states for the activator with looping (see Fig. 5). The on-rate, off-rate well as the number of activators is taken from [35] (page 82) When activator is bound, the recruitment of \({\mathsf {RNAP}}\) is increased by factor 10 or 50 ([34] and [27] respectivelly) (Table 3).

1.A.5   Repressor

The repression mechanism is inspired by the well-studied transcriptional regulation, there is a word missing after transcription of the \({\mathsf {lac}}\) operon, the repressor \({\mathsf {LacI}}\). We take the binding and unbinding rates for the repressor from ([39], Fig. 4).

1.A.6   Looping Rates

The stability of the looped state is incorporated in the model by scaling down the unlooping rate. We choose the scaling factors of 100 and 1000 based on the computation of the ratio of dissociation rates for the models with and without looping ([40], Table 1; parameter a in [39]). The mechnism proposed in, eg. [39] suggests that the looping increases the binding rate (due to increased local concentration of TFs), while leaving the unbinding rate unchanged. As the scaled on-rates may exceed theoretical limit for diffusion-limited reactions, in our model, we incorporate the same effect by leaving the binding rate identical, and scaling down the unlooping rate.

Appendix 1.B Kappa Models

figure a
figure b
figure c
figure d

Appendix 1.C Supplementary Theory and Proofs

1.C.1    Deterministic Limit

In the continuous, deterministic model of a chemical reaction network, the state \(\mathbf {z}(t) = (z_1,\ldots ,z_n)(t)\in {\mathbb R}^n\) is represented by listing the concentrations of each species. The dynamics is given by a set of differential equations in form

$$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}t}z_i = \nu _{ij}\sum _{j=1}^rk_j \prod _{i=1}^nz_i(t)^{a_{ij}}, \end{aligned}$$
(4)

where \(k_j\) is a deterministic rate constant, computed from the stochastic one and the volume \(N\) according to \(k_j := c_jN^{|{\mathbf {a}}_{j}|-1}\) (\(|\mathbf {x}|\) denotes the 1-norm of the vector \(\mathbf {x}\)). The deterministic model is a limit of the stochastic model when all species in a reaction network are highly abundant [19]: by scaling the species multiplicities with the volume: \(Z_i(t) = X_i(t)/N\), adjusting the propensities accordingly, in the limit of infinite volume \(N\rightarrow \infty \), the scaled process \({\mathbf Z}(t)\) follows an ordinary differential Eq. (4).

1.C.2    Expected Output in the Transient

The CME implies that the expectation of the marginal distribution of \(\{X_t\}\) satisfies the equations

$$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}t} \mathsf {E}(\mathbf {X}_t) = \sum _{j=1}^{r} \varvec{\nu }_{j}\mathsf {E}(\lambda _{j}(\mathbf {X}_t)). \end{aligned}$$
(5)

To check (5), observe a transition from \(\mathbf {x}\) to \(\mathbf {x}+\varvec{\nu }_j\). The term \(\lambda _{j}(\mathbf {x})\mathsf {P}(\mathbf {X}_t=\mathbf {x})\) appears exactly once when summing up for the state \(\hat{\mathbf {x}} = \mathbf {x}\) as the outflow probability, and exactly once when summing up for the state \(\hat{\mathbf {x}} = \mathbf {x}+\varvec{\nu }_j\), as the inflow probability. This gives the term \((\mathbf {x}+\varvec{\nu }_j)-\mathbf {x}=\varvec{\nu }_j\cdot \lambda _{j}(\mathbf {x})p^{(t)}(\mathbf {x})\). It is worth noting that, upon scaling the rate constants, the equations for \(\mathsf {E}(\mathbf {X}_t)\) are equivalent to (4) only if all rate functions are linear, that is, when all reactions are unimolecular.

1.C.3    Proof for Lemma 1

We first notice that the process \(\mathbf {X}(t)|_{\mathsf {D}}\) is indeed Markovian, because all states of \(\mathbf {X}(t)\) projected to the same state in \(\mathbf {X}(t)|_{\mathsf {D}}\) are behaviourally indistinguishable (bisimulation equivalent), due to rates between lumped states not depending on protein count. From (5), it follows that

where \(\langle x_{\mathsf {D}1j}(t)\rangle \) denotes the expected value of being in one of the active promoter configurations. The latter equals (2), since in every reachable state \(\mathbf {x}\in (D_{\texttt {0}}\uplus D_{\texttt {1}})\times \{0,1,\ldots \}\), exactly one \({\mathsf {DNA}}\) configuration takes value 1.

Appendix 1.D Supporting Figures

See Fig. 13.

Fig. 2.
figure 2

Transitions of the CTMC underlying basal gene expression. The state space \(S\cong \{\texttt {0},\texttt {1}\}\times \{0,1,2,\ldots \}\), such that \(s_{\texttt {1},x}\) denotes an active configuration (where the \({\mathsf {RNAP}}\) is bound to the \({\mathsf {DNA}}\)) and \(x\in \mathbf N\) protein molecules.

Fig. 3.
figure 3

Each binding configuration of the \({\mathsf {DNA}}\) can be active (green, polymerase bound) or inactive (gray, polymerase not bound). Protein count can increase only when the \({\mathsf {DNA}}\) configuration is active. (Color figure online)

Fig. 4.
figure 4

Searching for long lived transients in gene regulation without feedback: three modelled scenarios. We demonstrate that small delays \(\varDelta \) can raise qualitatively different phenotypes, which are stable for cell lifetime.

Fig. 5.
figure 5

Two prototype GRNs and their promoter logic: (a) Model without looping: regulatory architecture (promoter logic), (b) Model with looping: regulatory architecture (promoter logic). Mechanistic models are listed in (Appendix 1.B).

Fig. 6.
figure 6

Model without looping: the CTMC regulating six different \({\mathsf {DNA}}\) configurations. Thicker blue line denotes that the recruitment of \({\mathsf {RNAP}}\) is faster when the activator is bound. (Color figure online)

Fig. 7.
figure 7

Model with looping: CTMC regulating the \({\mathsf {DNA}}\) configurations has 23 different states. It is naturally represented as a composition of two sub-models: (left) the switching among configurations with respect to activator binding to its main and auxiliary binding sites (\(\mathsf {O_{Am}}\) and \(\mathsf {O_{Ax}}\) respectively, and (right) the switching among configurations with respect to repressor binding to its main and auxiliary binding sites (\(\mathsf {O_{Am}}\) and \(\mathsf {O_{Ax}}\) respectively. The unlooping rates (thicker blue lines) are typically much weaker than the TF unbinding. Any combination of the states in the two sub-models can be observed (reachable), except the state where both repressor and activator are looped. (Color figure online)

Fig. 8.
figure 8

The effect of kinetic parameters on the shape and duration of transients for a model with looping (for six parameter values listed in Table 1 and time delays of \(1\min \), \(5\min \) and \(15\min \) respectively).

Fig. 9.
figure 9

Long lived transients in protein expression follow the long lived transients (mixing times) in promoter activity. We plot the statistically inferred probability of promoter logic being in the active state for six different parameter combinations, listed in Table 1 and time delays of \(1\min \), \(5\min \) and \(15\min \) respectively.

Fig. 10.
figure 10

Long lived transients are not observed in the model without looping (for six parameter values listed in Table 1 and time delays of \(1\min \), \(5\min \) and \(15\min \) respectively).

Fig. 11.
figure 11

In the reference parameter set, when there is no delay (\(\varDelta = 0\)), the phenotype in scenario \(\mathcal{X}_{R\rightarrow A}(\varDelta )\) is equal to the one in scenario \(\mathcal{X}_{A || R}\), visibly different than the phenotype \(\mathcal{X}_R\) (290 protein molecules). The difference of from the scenario \(\mathcal{X}_{A || R}\) (the characteristic we formally termed amplitude – see Sect. 5) exponentially grows as the delay increases, that is, it quickly approaches the phenotype of scenario \(\mathcal{X}_R\). The difference of \(\mathcal{X}_{R\rightarrow A}(\varDelta )\) from \(\mathcal{X}_R\) becomes observably negligible already for delays larger than \(\varDelta \) = 5 min = 300 s (difference of 10 molecules, \(0.035\%\) of the initial difference). We obtained the dependency by fitting the data obtained by simulating the system for \(\varDelta \in \{1,20,40,60,120,180,240,300,600,900\}\).

Fig. 12.
figure 12

For chosen parameter sets (Tables 1 and 2) and for a delay \(\varDelta = 5\min \), we plot the amplitude and the half-life (defined in Sect. 2).

Fig. 13.
figure 13

We define a more global measure of the effect of delay in arrival times of TFs: first, the range of values that can be observed with a delay \(\varDelta = 5\min \) by the measuring the maximum distance between phenotypes the scenarios \(\mathcal{X}_{A\rightarrow R}\) and \(\mathcal{X}_{R\rightarrow A}\) \(\beta : = \max _{t \ge t_0} |x_p(t|\mathcal{X}_{A\rightarrow R})-x_p(t|\mathcal{X}_{R\rightarrow A})|\), and secondly, the halflife of this range \(t_{1/2}: = \arg \min _{t\ge t_{\beta }}\{t\mid |x_p(t|\mathcal{X}_{A\rightarrow R})-x_p(t\mid \mathcal{X}_{R\rightarrow A})|<\frac{1}{2}\beta \}\).

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Switzerland AG

About this paper

Check for updates. Verify currency and authenticity via CrossMark

Cite this paper

Guet, C., Henzinger, T.A., Igler, C., Petrov, T., Sezgin, A. (2019). Transient Memory in Gene Regulation. In: Bortolussi, L., Sanguinetti, G. (eds) Computational Methods in Systems Biology. CMSB 2019. Lecture Notes in Computer Science(), vol 11773. Springer, Cham. https://doi.org/10.1007/978-3-030-31304-3_9

Download citation

  • DOI: https://doi.org/10.1007/978-3-030-31304-3_9

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-030-31303-6

  • Online ISBN: 978-3-030-31304-3

  • eBook Packages: Computer ScienceComputer Science (R0)

Publish with us

Policies and ethics