Andreas Hilfinger, Thomas M. Norman, Johan Paulsson  Cell Systems 

Slides:



Advertisements
Similar presentations
Signal Processing in Single Cells Tony 03/30/2005.
Advertisements

By Rosalind Allen Regulatory networks Biochemical noise.
Xiaoshu Chen, Jianzhi Zhang  Cell Systems 
Pressure and Temperature Dependence of Growth and Morphology of Escherichia coli: Experiments and Stochastic Model  Pradeep Kumar, Albert Libchaber  Biophysical.
Jingkui Wang, Marc Lefranc, Quentin Thommen  Biophysical Journal 
Volume 19, Issue 4, Pages (April 2012)
Low Dimensionality in Gene Expression Data Enables the Accurate Extraction of Transcriptional Programs from Shallow Sequencing  Graham Heimberg, Rajat.
Masahiro Ueda, Tatsuo Shibata  Biophysical Journal 
Volume 4, Issue 4, Pages e5 (April 2017)
Dual Modes of Cdc42 Recycling Fine-Tune Polarized Morphogenesis
Computational Re-design of Synthetic Genetic Oscillators for Independent Amplitude and Frequency Modulation  Marios Tomazou, Mauricio Barahona, Karen.
Whole-Embryo Modeling of Early Segmentation in Drosophila Identifies Robust and Fragile Expression Domains  Jonathan Bieler, Christian Pozzorini, Felix.
Volume 82, Issue 1, Pages (April 2014)
Volume 19, Issue 4, Pages (April 2012)
Passive Noise Filtering by Cellular Compartmentalization
Transcription Stochasticity of Complex Gene Regulation Models
Choice Certainty Is Informed by Both Evidence and Decision Time
Transcriptional Memory in the Drosophila Embryo
A Role for Codon Order in Translation Dynamics
Codon Clarity or Conundrum?
Braden A. Purcell, Roozbeh Kiani  Neuron 
Volume 1, Issue 4, Pages (October 2015)
Single-Cell Analysis of Growth in Budding Yeast and Bacteria Reveals a Common Size Regulation Strategy  Ilya Soifer, Lydia Robert, Ariel Amir  Current.
Real-Time Kinetics of Gene Activity in Individual Bacteria
Noise Induces Hopping between NF-κB Entrainment Modes
Volume 52, Issue 1, Pages (October 2013)
The Mechanism of Rate Remapping in the Dentate Gyrus
Volume 111, Issue 2, Pages (July 2016)
Fuqing Wu, David J. Menn, Xiao Wang  Chemistry & Biology 
Hedging Your Bets by Learning Reward Correlations in the Human Brain
Cell-Size Control and Homeostasis in Bacteria
Intrinsic and Task-Evoked Network Architectures of the Human Brain
Corentin Briat, Ankit Gupta, Mustafa Khammash  Cell Systems 
Volume 49, Issue 1, Pages (January 2013)
Motor Learning Is Optimally Tuned to the Properties of Motor Noise
Cortical Mechanisms of Smooth Eye Movements Revealed by Dynamic Covariations of Neural and Behavioral Responses  David Schoppik, Katherine I. Nagel, Stephen.
Passive Noise Filtering by Cellular Compartmentalization
Volume 66, Issue 4, Pages (May 2010)
Volume 23, Issue 10, Pages (October 2016)
Andrew D. Bosson, Jesse R. Zamudio, Phillip A. Sharp  Molecular Cell 
Michał Komorowski, Jacek Miękisz, Michael P.H. Stumpf 
Lea Goentoro, Marc W. Kirschner  Molecular Cell 
Nicholas J. Priebe, David Ferster  Neuron 
Volume 88, Issue 3, Pages (November 2015)
Volume 5, Issue 4, Pages e4 (October 2017)
Volume 4, Issue 5, Pages e5 (May 2017)
Integration Trumps Selection in Object Recognition
BOLD fMRI Correlation Reflects Frequency-Specific Neuronal Correlation
Volume 3, Issue 5, Pages e13 (November 2016)
Neuronal Selectivity and Local Map Structure in Visual Cortex
Medial Axis Shape Coding in Macaque Inferotemporal Cortex
Dual Modes of Cdc42 Recycling Fine-Tune Polarized Morphogenesis
Quantitative Imaging of Transcription in Living Drosophila Embryos Links Polymerase Activity to Patterning  Hernan G. Garcia, Mikhail Tikhonov, Albert.
Volume 37, Issue 6, Pages (March 2010)
Regulation by Small RNAs in Bacteria: Expanding Frontiers
Neuronal Origins of Choice Variability in Economic Decisions
Søren Vedel, Harry Nunns, Andrej Košmrlj, Szabolcs Semsey, Ala Trusina 
Stephen V. David, Benjamin Y. Hayden, James A. Mazer, Jack L. Gallant 
Arjun Raj, Alexander van Oudenaarden  Cell 
Supervised Calibration Relies on the Multisensory Percept
Semblance of Heterogeneity in Collective Cell Migration
Theodore R. Rieger, Richard I. Morimoto, Vassily Hatzimanikatis 
Brandon Ho, Anastasia Baryshnikova, Grant W. Brown  Cell Systems 
Volume 66, Issue 4, Pages (May 2010)
Xiaoshu Chen, Jianzhi Zhang  Cell Systems 
Stable Stochastic Dynamics in Yeast Cell Cycle
Joana Pinto Vieira, Julien Racle, Vassily Hatzimanikatis 
Volume 28, Issue 19, Pages e8 (October 2018)
Iain G. Johnston, Nick S. Jones  The American Journal of Human Genetics 
George D. Dickinson, Ian Parker  Biophysical Journal 
Presentation transcript:

Exploiting Natural Fluctuations to Identify Kinetic Mechanisms in Sparsely Characterized Systems  Andreas Hilfinger, Thomas M. Norman, Johan Paulsson  Cell Systems  Volume 2, Issue 4, Pages 251-259 (April 2016) DOI: 10.1016/j.cels.2016.04.002 Copyright © 2016 Elsevier Inc. Terms and Conditions

Cell Systems 2016 2, 251-259DOI: (10.1016/j.cels.2016.04.002) Copyright © 2016 Elsevier Inc. Terms and Conditions

Figure 1 Invariant Relations that Characterize Entire Classes of Stochastic Networks (A) Consider all possible stochastic processes that have in common that some component X2 within a large network is made at a rate proportional to X1 levels. Stochastic processes in this class do not just involve the two components X1 and X2 but can involve an arbitrarily large number of other interacting components (X3, X4, X5,…, as indicated by the cloud). Such systems can then, for example, realize arbitrary control networks or extrinsic noise influences, which typical modeling approaches must consider individually as specific realizations. (B) The reactions that are conserved across this class of systems are defined. (C) Simulation data for specific realization of such systems with different parameters and global network topologies (navy dots, systems with negative feedback; red, positive feedback; green, extrinsic noise). Depending on the chosen details, systems exhibit vastly different averages and variances. (D) Regardless of the details, Equation 2 immediately shows that all such systems must satisfy the simple invariant relation CV22=1/〈x2〉+Cov(x1,x2)/(〈x1〉〈x2〉), illustrating how a small set of specified molecular interactions produce testable predictions between variances and covariances that are rigorously independent of global network influences. The data correspond to the same simulations depicted in (C). (E) As an additional example, we consider a different class of systems in which the specified X2 production rate is proportional to the cube of X1 levels and X2 molecules degrade through an irreversible dimerization reaction. All other details remain arbitrary. (F) Depending on the details, averaged abundances and noise levels vary enormously. (G) Nevertheless, Equation 2 immediately shows that all possible processes of this type must satisfy Cov(x2,x22−x2)/[〈x2〉(〈x22〉−〈x2〉)]=1.5/〈x2〉+Cov(x2,x13)/(〈x2〉〈x13〉), as illustrated by simulation data for various systems of this type. (Because each degradation reaction removes two X2 molecules, we have 〈s22〉=(1+2)/2=1.5 for all systems of this type.) Note that systems in this class do not satisfy the invariant relation of (D) (see the Supplemental Results). Cell Systems 2016 2, 251-259DOI: (10.1016/j.cels.2016.04.002) Copyright © 2016 Elsevier Inc. Terms and Conditions

Figure 2 Comparing Published Gene Expression Data against a Broad Class of Models Comparing gene expression data (Taniguchi et al., 2010) from E. coli to the invariant relation derived for the class of systems in which translation is proportional to mRNA levels and proteins undergo first-order degradation. Regardless of molecular lifetimes, feedback control, and upstream randomization effects, all such systems must fall on the red line defined by Equation 4 (for protein abundances as high as those of the data). Only very few genes are consistent with this prediction. Bars indicate reported error estimates (Taniguchi et al., 2010). The blue part of the line identifies the range predicted by the specific case of the model considered by Taniguchi et al. (2010), with an effective protein lifetime of 180 min and mRNA lifetimes between 2 and 10 min, as reported in Taniguchi et al. (2010). The distance of individual genes from the prediction is quantified in the Supplemental Results. Cell Systems 2016 2, 251-259DOI: (10.1016/j.cels.2016.04.002) Copyright © 2016 Elsevier Inc. Terms and Conditions

Figure 3 Independent Translational Noise Cannot Explain the Data Consider all models in which protein levels are randomized by spontaneous mRNA fluctuations and independently fluctuating translation rates per mRNA molecule. All such systems must satisfy Equation 6 and thus fall onto the red line. Direct comparison with experimental data (Taniguchi et al., 2010) shows that most genes are inconsistent with such models. The data here is the subset of genes for which the mRNA lifetime τ1 was explicitly reported. To plot the data, we assumed an effective protein lifetime τ2 of 180 min set by the reported cell-generation times. Active degradation of proteins on a faster timescale would move the data to the right. Cell Systems 2016 2, 251-259DOI: (10.1016/j.cels.2016.04.002) Copyright © 2016 Elsevier Inc. Terms and Conditions

Figure 4 Accounting for Fluorescent Maturation Does Not Make Simple Mechanisms Consistent with the Data (A) Accounting for the maturation of fluorescent proteins turns the invariant relation of Equation 4 into the inequalities of Equation 7. The colored area corresponds to the accessible region with a protein lifetime τ2 = 180 min (set by the cell-cycle time) and the estimated fluorescent protein maturation time τmat = 7.5 min (Xiao et al., 2008). Many of the genes remain inconsistent with the theoretical prediction. Only genes that fall within the colored region can be consistent with the simple assumption that translation rates are proportional to mRNA levels, even when accounting for a finite fluorescent maturation time. The tested class of systems is very broad (see cloud), allowing for extreme feedback scenarios in which the protein levels effectively control mRNA levels rather than the other way around. In contrast, without allowing for feedback, the blue line indicates the accessible range of the original model subject to maturation with the above parameters and mRNA lifetimes between 2 and 10 min. (B) Explicitly accounting for delayed fluorophore maturation turns the invariant relation of Equation 6 into the prediction Equation 8. Comparing the data against this invariant relation (for τmat = 7.5 min, τ2 = 180 min, and τ1 as reported individually for each gene) shows that many of the genes remained inconsistent with the class of models defined by Equation 5, even when accounting for a finite fluorescent maturation time. Cell Systems 2016 2, 251-259DOI: (10.1016/j.cels.2016.04.002) Copyright © 2016 Elsevier Inc. Terms and Conditions

Figure 5 Models with Noise Cancellation between Translational and Transcriptional Fluctuations Can Explain the Data Systems in which mRNA levels and the fluctuations in the protein synthesis rate per mRNA are correlated are not constrained by the invariants Equations 4 or 6. Indeed, numerical simulation data (blue dots) indicate that systems in which the effects of transcription and translational noise partly cancel out can exhibit fluctuations similar to that of the experimental data. Representative histograms of protein distributions for such systems show a near perfect fit to negative binomials (black lines) despite the significant noise cancellation. Systems with noise cancellation thus exhibit fluctuations that are consistent with both experimental observations reported in Taniguchi et al. (2010). Cell Systems 2016 2, 251-259DOI: (10.1016/j.cels.2016.04.002) Copyright © 2016 Elsevier Inc. Terms and Conditions

Cell Systems 2016 2, 251-259DOI: (10.1016/j.cels.2016.04.002) Copyright © 2016 Elsevier Inc. Terms and Conditions