• Aucun résultat trouvé

Receptor-induced transient responses in cells with oscillatory actin dynamics

N/A
N/A
Protected

Academic year: 2021

Partager "Receptor-induced transient responses in cells with oscillatory actin dynamics"

Copied!
19
0
0

Texte intégral

(1)

HAL Id: hal-03010651

https://hal.archives-ouvertes.fr/hal-03010651

Submitted on 30 Nov 2020

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

oscillatory actin dynamics

Jose Negrete, Alain Pumir, Christian Westendorf, Marco Tarantola, Eberhard

Bodenschatz, Carsten Beta

To cite this version:

Jose Negrete, Alain Pumir, Christian Westendorf, Marco Tarantola, Eberhard Bodenschatz, et al..

Receptor-induced transient responses in cells with oscillatory actin dynamics. Physical Review

Re-search, American Physical Society, 2020, 2 (1), �10.1103/PhysRevResearch.2.013239�. �hal-03010651�

(2)

J Negrete Jr1,2,3, A Pumir3,4, C Westendorf3, M Tarantola3, E Bodenschatz3,5,6 and C Beta3,7∗

1Max Planck Institute for the Physics of Complex Systems (MPIPKS), D-01187 Dresden, Germany 2´

Ecole Polytechnique F´ed´erale de Lausanne (EPFL), CH-1015 Lausanne, Switzerland

3

Max Planck Institute for Dynamics and Self-Organization (MPIDS), D-37077 G¨ottingen, Germany

4Laboratoire de Physique, Ecole Normale Sup´erieure de Lyon,

Universit´e de Lyon 1 and Centre National de la Recherche Scientifique, F-69007 Lyon, France

5

Institute for Dynamics of Complex Systems, University of G¨ottingen, D-37073 G¨ottingen, Germany

6Laboratory of Atomic and Solid-State Physics and Sibley School of Mechanical and Aerospace Engineering,

Cornell University, Ithaca, New York 14853, USA and

7

Institute of Physics and Astronomy, University of Potsdam, D-14476 Potsdam, Germany Living cells adjust their sensing and migratory machinery in response to changes in their envi-ronment. In this work, we show that cells of the social amoeba Dictyostelium discoideum modulate the dynamical state of their actin cytoskeleton in response to an external pulse of the chemoattrac-tant cyclic adenosine monophosphate (cAMP). In particular, we focus on a population of cells that exhibit noisy oscillatory cycles of actin polymerization and systematically study receptor-induced transitions in their cytoskeletal dynamics. In response to a short external pulse of cAMP, these cells adopt a noisy quiescent state, before returning to their initial, oscillatory dynamics. The response exhibits a biphasic time profile, with a duration that shows strong variability between cells; it can extend as long as approximately twelve oscillation cycles. We propose a model that is based on a generic nonlinear noisy oscillator. Our theoretical analysis suggests that the transient termination of oscillations in response to a receptor stimulus occurs via a Hopf bifurcation.

I. INTRODUCTION

Amoeboid cells constantly probe the chemical compo-sition of their surroundings, in order to respond to chang-ing environmental conditions. For instance, individ-ual cells of the social amoeba Dictyostelium discoideum (D. discoideum) react to extracellular cAMP, emitted by neighboring D. discoideum cells, to aggregate and self-organize into a multicellular fruiting body. Once de-tected by the corresponding receptors, the cAMP signal is processed by a complex system of chemotactic signal-ing pathways, which finally leads to reorganization of the actin cytoskeleton and migration of the cells towards in-creasing concentrations of cAMP. The cytoskeletal acto-myosin system not only generates the forces required for amoeboid locomotion, but also determines the shape and mechanical stability of the cells [1]. In our previous work, we have probed the dynamical properties of actin poly-merization in response to external pulses of cAMP [2]. We found that the polymerization/depolymerization dy-namics at the cell cortex is reminiscent of a damped os-cillator, and the response to a cAMP pulse saturates above a critical cAMP level [3]. A subpopulation of the investigated cells exhibits self-sustained oscillations in the polymerization/depolymerization dynamics, lead-ing to the conjecture that the actin cytoskeleton operates close to an oscillatory instability. Self-sustained oscilla-tions are an example for the rich autonomous dynamics of the cytoskeleton and its governing chemotactic sig-naling system [4, 5]. Other examples include intracellu-lar traveling waves and spontaneous pseudopod

forma-∗beta@uni-potsdam.de

tion resulting in random motility and reorientation [6]. These autonomous processes can be directly influenced by receptor-mediated chemoattractant stimuli [7]. In this work, we focus on the subpopulation of cells, which show self-sustained oscillations (see figure 1 for an example) and probe their response to a cAMP stimulus. We re-port that a single short pulse of cAMP (∼ 1 sec) induces a transition of the actin cytoskeleton in D. discoideum from self-sustained oscillations to a non-oscillatory state. Af-ter a transient time of varying length, the actin cytoskele-ton returns back to its initial oscillatory state, showing that the D. discoideum amoeba may change their dynam-ical state as a response to an external chemoattractant stimulus.

We use the flow photolysis method to stimulate single cells with short-time pulses of cAMP. This method relies on the photo-activation of caged compounds in microflu-idic channels [8, 9]. The cytoskeletal actin response is recorded as changes in the subcellular LimE-mRFP dis-tribution [2]. This Lim-domain protein is a marker for filamentous actin [10, 11] and reflects changes in cortical actin polymerization by translocating between the cy-tosol and the cell cortex [12], see figure 1 (a). The tran-sient response of a D. discoideum amoeba to a sudden change in cAMP concentration can be divided into two phases (biphasic response). First, both the signalling sys-tem and the actin cytoskeleton are uniformly activated along the cell membrane with a characteristic duration of approximately 20 sec [13–15]. This is followed by a second more irregular phase, during which localized patches of activity appear and disappear at random lo-cations around the cell perimeter [16–19]. This response pattern is also observed for the oscillatory cells investi-gated here. Interestingly, during the second phase, which

(3)

FIG. 1. Experimental set-up and data analysis. (a) Exemplary image series. The D. discoideum cell is centered within a 48 × 48 µm imaging region. The cAMP stimulus is released upstream of the cell by a line scan of a 405 nm laser (uncaging region, purple) and advected by the flow towards the cell (mean velocity = 120 µm/sec). The imaged LimE-mRFP fluorescence is segmented into a cortical and a cytosolic signal (displayed in frame 2, blue/yellow). (b) Averaged fluorescence intensity of the cortical (blue) and the cytosolic region (yellow) over time. The respective frames in (a) are marked by blue/yellow dots. (c) Cytosolic signal after detrending. The cAMP stimulus was applied at t = 100 sec and is denoted by a gray bar in (b) and (c).

varies in duration between∼ 40 to ∼ 200 sec, the oscil-latory dynamics stop before the cells recover their initial oscillatory state (see figure 2 a & d).

The aim of this work is to characterize the dynami-cal transition behind this observation. A dynamidynami-cal sys-tem with oscillatory dynamics can transition to a non-oscillatory state in different ways [20, 21]. Here, we use as a starting point a generic model, which implicitly as-sumes that the oscillatory system operates close to a Hopf bifurcation and is subject to additive noise. In addition to the frequency of the underlying oscillations, the model introduces three parameters: λ characterizes the distance to the bifurcation (i.e. the transition point to the non-oscillatory state), D gives the amplitude of the noise, and g sets the strength of the nonlinearity. Based on the dimensionless ratio λ/√gD, we can quantify the relative impact of deterministic oscillations and noise on the over-all dynamics. This model provides a good description of oscillatory cells in the absence of any stimulation [22], and we now extend the analysis to cells responding to an external cAMP pulse.

II. EXPERIMENTAL AND ANALYSIS

METHODS

A. Experimental set-up and protocol

D. discoideum cell culture and preparation for an ex-periment were carried out as previously described [2, 22]. The starved cells were injected into the microfluidic chan-nel (500 µm × 30 mm × 26 µm) and allowed to set-tle for 15 min. Afterwards, the phosphate buffer (PB) was replaced with a solution of 10 µM BCMCM-caged cAMP in PB (Biolog, Bremen, Germany). A 250 µl sy-ringe (Hamilton, Bonaduz, Switzerland), mounted on a standard infuse syringe pump (Harvard Apparatus PHD 2000, Holliston, MA, USA) was used to insert the caged compound at a constant flow rate of 5 µl/h, corre-sponding to an average flow velocity of approximately 120 µm/sec.

In this work, we analyzed fluorescence images of LimE-mRFP [23], which labels polymerized actin, in combi-nation with either Coronin-GFP [24] or Aip1-GFP [25]. The fluorophores were imaged with an Olympus FV-1000 inverted confocal laser scanning microscope, using an UPLSAPO 60X oil immersion objective. The mRFP was excited by the 543 nm line of a multiline 1 mW HeNe laser (Melles Griot, Carlsbad, CA, USA). Each cell was imaged individually in a region of 48×48 µm (120×120 pixels, figure 1(a)), with the confocal plane set

(4)

approx-0 50 100 150 200 250 300 350 400 −400 −200 0 200 400 Intensity [au]

(a)

τa τφ LimE-mRFP H[LimE-mRFP] 0 50 100 150 200 250 300 350 400 −400 −200 0 200 400 Intensity [au]

(d)

τa τφ LimE-mRFP H[LimE-mRFP] 0 50 100 150 200 250 300 350 400 0 100 200 300 400 A(t)

(b)

τa µa(1)± σa(1) µa(2)± σa(2) 0 50 100 150 200 250 300 350 400 0 100 200 300 400 A(t)

(e)

τa µ(1) a ± σa(1) µ(2) a ± σa(2) 0 50 100 150 200 250 300 350 400 Time [s] −20 −10 0 10 φ (t )− ω t

(c)

τφ 0 50 100 150 200 250 300 350 400 Time [s] −40 −30 −20 −10 0 φ (t )− ω t

(f)

τφ

FIG. 2. Examples of single cell D. discoideum LimE-mRFP responses to a short pulse of cAMP. The two columns correspond to two different cells. Panels (a) and (d) show the detrended fluorescence time traces of LimE-mRFP (red) and their corresponding Hilbert transforms (blue). The cells are stimulated at τstim= 100 sec with a pulse of cAMP for one second (the stimulation time

is denoted by a gray bar). The evolution of the amplitudes of the time traces are shown in (b) and (e), and the corresponding phase drifts in (c) and (f). The response to cAMP is divided into two phases. The duration of the first phase, the amplitude transient time τa, is defined by the time it takes to return to 25% of the response amplitude maximum. The duration of the

second phase, which we call the phase transient time τφ, is defined as the time interval between the end of the first phase and

recovery of the phase φ(t) back to its main frequency (shown in (c) and (f)). The amplitude mean (black line) and standard deviation (dashed lines) are calculated over the first 100 sec and in the second phase of the transient response, shown in (b) and (e).

imately 2 µm above the cover slip surface. At the up-stream side of the imaging region, a line of 0.4×40 µm (1×120 pixels, figure 1(a), panel 1, pink) was scanned by a 405 nm laser diode to photo-chemically release the cAMP from its caged precursor (see flow photolysis, de-scribed in [8, 26]). The cAMP was transported by the fluid flow towards the cell, which was centered at a dis-tance of about 20 µm downstream of the uncaging region. During each experiment, the cell was first imaged for 100 sec in the absence of any cAMP stimulation to record the initial activity of the actin cytoskeleton (imag-ing frame rate: 1 Hz). These 100 sec of each data set were previously presented and analyzed in [22]. At τstim= 100 sec, we initiated the uncaging laser to

stim-ulate the cell with an external pulse of cAMP for one second and continued the recordings for a time interval of 100 to 300 sec.

B. Data analysis

The LimE-mRFP images were segmented into two re-gions, indicated in figure 1 (a, panel 2). The cortex is defined as a 1.2 µm thick region at the cell’s outer edge (blue), while the cytosol is defined as the remainder of the cell’s interior (yellow). The mean of each region yields the average fluorescence that is shown over time in fig-ure 1(b). Each cytosolic time trace was detrended by

(5)

subtracting a moving average (figure 1 (c), window size of 40 sec). The data sets were divided into oscillatory and non-oscillatory time traces by calculating the auto-correlation function of the first 100 sec prior to cAMP stimulation. Only time traces with an autocorrelation function that showed at least one oscillation cycle with a period below 33.33 sec (i.e. 1/3 of the time window before τstim) were categorized as oscillatory and selected for further analysis. We furthermore required that the largest peak that occurs within the first 33.33 sec of the autocorrelation reaches at least 10% of the initial value of the autocorrelation function. In addition, we excluded cells that did not show any response to the cAMP stim-ulus in the amplitude or phase signal.

Out of a total of 165 recorded cells, 26 passed this stringent selection protocol and were included in the sub-sequent analysis. The corresponding time traces and their respective autocorrelation functions are shown in figures 11–14 in the appendix. We also computed the Hilbert transform of the LimE-mRFP time trace for each case, an example is shown in blue in figure 2 (a) & (d). This allowed us to calculate the amplitude A(t) (figure 2 (b) & (e)) and phase φ(t) (figure 2 (c) & (f)) as a func-tion of time. The majority of cases (22 cells) show a deceleration of the phase after stimulus. Only four cases, displayed in the four last rows of figure 14, exhibit an accelerating phase after stimulation.

Based on the time traces of amplitude and phase records, we can then define the time scales of the two phases of the response (biphasic response profile). The duration of the first phase is defined as the interval be-tween the time point of stimulation and the time point when the amplitude has dropped back to 25% of its max-imal value. We call this time scale the amplitude tran-sient time τa. The time evolution of the phase φ(t) is well captured by

φ(t) = ωot + φo+ η(t) , (1) where ωo is the central frequency of the time series, φo is the initial phase value, and η(t) is the deviation from the linear evolution caused by noise. We fitted the first 100 sec of the phase with a simple functional form ωt+φ0

o, where ω and φ0

oare determined by minimizing the differ-ence between the signal φ(t) and the linear dependdiffer-ence ωt + φ0

o (L2 norm of φ− ωt − φ0o). Elementary algebra shows that ω = 12 t2 obs h φ(t)t−φ(t)t2obs i, (2) where we denote ω as the detected frequency and the overbar as the average over the observation time tobs. The detected frequency ω might differ from the value of the central frequency ωo, the error between them will depend on tobs.

To determine the duration of the second phase of the cellular cAMP response, we analyzed the phase drift

φ(t)− ωt. This quantity drifts almost linearly during the transient phase, followed by a recovery towards a constant value. The time interval between the end of τa and the end of the phase drift is defined as the phase transient time τφ. Note that in 9 out of the total num-ber of 26 cases, the phase transient time τφ exceeds the duration of the recording and cannot be determined. For this reason, the analysis of the phase transient time τφ is based on 17 records only, where as 26 data points are available in all other cases.

III. MODELLING THE ACTIN

CYTOSKELETON DYNAMICS: FAST OSCILLATIONS, SLOW RESPONSES AND

NOISE SCALING

In this section, we present the extension of the model introduced in [22], to capture the transient regime in-duced by a pulse of chemoattractant, see Fig. 2. As an initial test, we demonstrate that our model captures semi-quantitatively the statistical features of the signal shown in figure 4.

We briefly recall the structure of the original model [22]. Namely, the system is described by a complex variable, z = x + iy, satisfying the equation:

dz

dt = [λ + iωo]z− g|z| 2

z + ξ(t) , (3)

The real part Re[z] = x corresponds to the mean fluores-cence emitted by polymerization markers at time t. The imaginary part, Im[z] = y, can be viewed as a negative feedback acting on them. The parameter ωo is the cen-tral frequency of the oscillator. In the absence of noise (ξ = 0), the resting state, z = 0, is a solution, whose stability is given by the sign of λ. For λ > 0 the resting state is unstable with respect to small perturbations, and the system displays a stable limit cycle. The amplitude of the limit cycle oscillations is set by a balance between the amplification rate and the nonlinearity, g|z|2z, and scales as∼pλ/g. For λ < 0 the resting state is stable and os-cillations are damped. The noise, ξ(t) = ξR(t) + iξI(t), is taken to be Gaussian, with short time correlations that are white in time in the present analysis,

hξi(t)ξj(t0)i = 2Dδ(t − t0)δij (4) where i, j refer to the indices R and I. To capture the re-sponse of the cell, we introduce two additional variables: s(t) describing the direct effect of the receptor stimulus on the actin polymerization, and an inhibitor w(t), trig-gered by the stimulation. The corresponding interaction network is summarized in figure 3 (a). The model given by Eq. (3) is then extended as

dz dt = [˜λ(w(t)) + iωo]z− g|z| 2 z−s(t)τ r + ξ(t) . (5)

(6)

FIG. 3. Model for the response of the D. discoideum actin cytoskeleton to short time cAMP pulses. (a) Network topology of the proposed model. The signal s(t) is activated by an external cAMP stimulation. The variable x(t) describes the dynamics of LimE-mRFP (i.e. the actin polymerization), and y(t) is the negative feedback, leading to spontaneous oscillations in the absence of an external stimulus. The variable w(t) acts as an inhibitor which determines the duration of the transient response. (b) The response of w(t) elicited by the signal s(t) (inset). Several realizations with different values of τ (25 sec ≤ τ ≤ 200 sec) are shown. (c) A single numerical realization of the dynamics of x(t) (red) and its Hilbert transform (blue) with the corresponding amplitude A(t) (d) and phase drift φ − ωot (e).

The stimulus, s(t), has the following functional form,

s(t) = so

t− τstim τs

4

e−t−τstimτs Θ(t− τstim) , (6)

where Θ(t) corresponds to the Heaviside function (Θ(t) = 0 for t < 0, and Θ(t) = 1 for t > 0). Its time evolution is shown in the inset of Fig. 3 (b), which corresponds to the response profile of the upstream regulator Ras, which was measured in Ref. [27], see the appendix for details. The stimulus acts directly on the oscillatory variable x through the additive term −s(t)/τr; this interaction is represented by a vertical arrow in Fig. 3 (a). In addition, the stimulus s(t) triggers the inhibitory variable, with a characteristic response time τ according to

dw dt =

1

τ(s(t)− w) . (7)

The inhibitory action of w(t) on the dynamics of z occurs via a modulation of ˜λ(w(t)) given by

˜

λ(w(t)) = λ− λoΘ(w(t)− K) . (8) Effectively, the variable λ is reduced by a positive quan-tity λo as long as w(t) > K, and returns to the orig-inal value when w(t) < K. Therefore, in this model, the inhibitory variable w(t) controls the duration of the transient response. Figure 3 (b) shows the response of w(t) for different values of the relaxation time τ . As τ increases, w(t) becomes less sensitive to s(t) and takes longer to relax back to its initial state.

As an example, figure 3 (c) shows a single numerical re-alization of the oscillatory actin signal x(t) that receives a stimulus at t = 100 sec, showing that the system tran-siently oscillates with a reduced amplitude after stimula-tion. The corresponding amplitude A(t) and phase drift

(7)

φ(t)− ωot are shown in figure 3 (d) and (e), respectively. The behavior of the numerical solution is qualitatively similar to our experimental results, see figure 2. In the appendix, we show several examples of how individual ex-perimental time traces can be associated with numerical realizations of the model, demonstrating that the model is able to reproduce details of the experimental data, see figure 15.

In the model, the relaxation time τ of the inhibitory variable w(t) determines the total duration of the bipha-sic response T = τa+ τφ. The dependence of T on the relaxation time τ in the model is shown in figure 7 (b), given a fixed value of K. To model the variability in the duration T of the biphasic response observed in experi-ments, we assume, for simplicity, that the relaxation time τ is distributed exponentially,

ρ(τ ) = Θ(τ− τ1) τ2

e−(τ −τ1)/τ2, (9)

with two fitting parameters τ1 and τ2. However, from the experiments the distribution ρ(τ ) is not directly ac-cessible. Only the distribution ρ(T ) of the total duration T is known, see figure 7 (c). We fitted the experimental distribution ρ(T ) by tuning the fitting parameters τ1,2in the distribution ρ(τ ), which determines the correspond-ing distribution ρ(T ) via relation (11), see also the ap-pendix for a detailed presentation of the fitting proce-dure. In our numerical simulations, we then draw for each simulation a different value of τ from the distribu-tion ρ(τ ).

Previously we have shown that the oscillatory actin dynamics follows equation (3), where each cell is char-acterized by a different value of the dimensionless pa-rameter Λ = λ/√gD [22]. Therefore, in simulations we assigned to each cell a different value of Λ, drawn from the following distribution

ρ(Λ) = 1

p2πσ2 Λ

e−Λ2/2σ2Λ, (10)

where σ2

Λ= 2.45, see figure 7 (d). All other parameters in equations (5–6) are set to constant values that are given in the appendix.

IV. STATISTICAL CHARACTERIZATION OF

TRANSIENTS AND NOISE SCALING

The mean and standard deviation of the amplitude transient times are ai = 29 sec and στa = 13.6 sec

respectively, thus the coefficient of variation yields στa/hτai = 0.39. The phase transient times (τφ) show

a broader distribution (hτφi ± στφ = 66 ± 44.5 sec)

with a coefficient of variation of στφ/hτφi = 0.70. The

largest phase transient time (τφ) observed is more than two times larger than the largest amplitude transient

time, (max[τφ] = 2.11 max[τa]) (see in the appendix fig-ure 6 (a) and (b) for histograms of the distribution).

We determined the mean and standard deviation of the amplitude of the detrended LimE-mRFP fluorescence signal prior to stimulation (µ(1)a & σ

(1)

a ) and during the second phase of the cAMP response (µ(2)a & σ

(2) a ), see figure 2 (b) & (e). Figure 4 (a) shows the standard devi-ation as a function of the amplitude mean before stimu-lation (blue) and during the second phase of the response (red). The data indicate that the standard deviation σa increases mean µa. In the majority of recorded cell re-sponses, the mean and standard deviation of the F-actin fluorescence signal is decreased after stimulation, see the negative quadrant in figure 4 (b). We did not find any sig-nificant correlation between the duration of the responses and their corresponding amplitudes (see supplementary figure 10).

Finally, we quantified the coefficient of variation (CVi= σ

(i) a /µ

(i)

a ), which measures the level of amplitude noise, where i = 1, 2 denotes the coefficient of variation before stimulation and during the second phase of the response. Changes in CVi are displayed in figure 4 (c) as a function of the coefficient of variation before stim-ulus, CV1. Roughly, half of the cells exhibited a noisier regime after stimulation, (CV2− CV1)/CV1 > 0, while the amplitude noise is reduced for the remaining cells.

Paragraph about τa and τφ moved to the appendix

We also compared the experimental mean values µa and standard deviations σa to the results from numerical simulations of our model in figure 4. The distribution of σa as a function of the mean µa before stimulus (fig-ure 4 (d), blue points) can be attributed to cell to cell variability. This is taken into account in the simulations by including parameter distributions, such as, for exam-ple, the distribution of λ/√gD. The distribution after stimulus (red points) is more compact and centered at lower values compared to the distribution before stimu-lation (blue points). A similar behavior was observed for the experimental values, see figure 4 (a).

In figure 4 (e), the relative change in mean and stan-dard deviation before and after stimulation are shown. In agreement with the experiments, mean and standard deviation decrease and the simulation results lie mostly within the lower left quadrant. With respect to the am-plitude noise, the model correctly reproduces that for half of the cells the dynamics of the actin cytoskeleton become noisier (CV2−CV1

CV1 > 0), while for the other half

the dynamics become more coherent (CV2−CV1

CV1 < 0), see

figure 4 (f).

V. FREQUENCY ANALYSIS

We denote the frequency detected before stimulation as ω1 and the frequency detected during the second phase of the response as ω2. The experimentally de-rived distributions of ω1and ω2are shown in figure 5 (a)

(8)

0 50 100 150 200 250 µa 0 20 40 60 80 σa

(a)

n = 26

Experiment

Before stimulation Duringτφ 0.0 0.1 0.2 0.3 µa 0.00 0.02 0.04 0.06 0.08 0.10 σa

(d)

n = 4684

Simulation

Before stimulation Duringτφ −1.0 −0.5 0.0 0.5 1.0 (µa(2)− µa(1))/µa(1) −1.0 −0.5 0.0 0.5 1.0 (2 ) a − σ (1 ) a )/ σ (1 ) a

(b)

n = 26 −1.0 −0.5 0.0 0.5 1.0 (µa(2)− µa(1))/µa(1) −1.0 −0.5 0.0 0.5 1.0 (2 ) a − σ (1 ) a )/ σ (1 ) a

(e)

n = 4684 0.2 0.4 0.6 0.8 1.0 1.2 CV1 −1.0 −0.5 0.0 0.5 1.0 1.5 2.0 (C V2 − C V1 )/ C V1

(c)

n = 26 0.2 0.4 0.6 0.8 1.0 1.2 CV1 −1.0 −0.5 0.0 0.5 1.0 1.5 2.0 (C V2 − C V1 )/ C V1

(f)

n = 4684 low high Point cloud density

low high lowPoint cloudhigh

density

low high

Point cloud density

FIG. 4. Mean, standard deviation and coefficient of variation in the amplitude fluctuations. Experimental data: (a) Mean amplitude µa and standard deviation σa measured before stimulation (blue) and during τφ (red). (b) Relative change of the

mean amplitude versus relative change of the standard deviation. Superscript indices (1) and (2) denote the mean/standard deviation before stimulation and during τφ, respectively. (c) Changes in the amplitude coefficient of variation before stimulation

(CV1 = σ(1)a /µ(1)a ) and during τφ (CV2 = σ(2)a /µ(2)a ). Model predictions: (d) Mean amplitude µa and standard deviation σa

derived from numerically evaluated time traces. Blue and red dots correspond to the time intervals before stimulation and during τφ, respectively. (e) Relative change of the mean amplitude versus relative change of the standard deviation. (f) Changes

in the amplitude coefficient of variation before stimulation (CV1) and during τφ(CV2). For all point clouds in (d)-(e) a kernel

density estimate of the distribution was calculated and the color gradient indicates the density of points. Additionally, contours at different density levels were drawn into (d), as both point clouds are partially overlapping.

in blue and orange, respectively. The histogram of ω1 is peaked at ωo ∼ 2π/13 sec−1 with a mean value of ω1

= 2π/15.1 sec−1 and a standard deviation of σω1 = 0.07 sec

−1, while the histogram of ω

2 is irregular shaped (ω2 ± σω2 = 2π/28.0 ± 0.06 sec

−1). Addition-ally, in figure 5 (b) each cell is represented as a point in the (ω1, ω2)-plane to show that for the majority of cells the frequency of cortical oscillations decreases after stim-ulation.

In the numerical simulations of the model, we set the frequency ωoin equation (5) to ωo= 2π/13 sec−1. Before stimulation, the frequency ω1determined from numerical simulations of the model fluctuates, with a mean value of hω1i ± σω1 = 2π/14.0± 0.07 sec

−1, see figure 5 (c). The

good agreement between the distribution of frequencies in the experiment and in the model suggests that the dispersion in the recorded frequency is due to phase noise, as discussed in the appendix.

After stimulation, the frequency ω2 predicted by the model is also fluctuating (hω2i ± σω2 = 2π/35.4 ±

0.15 sec−1), see figure 5 (c). We note that for a noisy oscillator (Eq. 1) with a small value of λ/√Dg as it oc-curs in the noise-dominated or weakly oscillatory case, the amplitude of z can occasionally become zero, result-ing in a phase jump (phase slip; see appendix for details). This makes the definition of the phase ambiguous, so that the observation of ω2 should be interpreted as an effec-tive frequency. When (λ− λo) is large and negative, the

(9)

0.2 0.0 0.2 0.4 0.6 0.8 1.0 w[s 1] 0 1 2 3 4 5 6 7 8 count s

(a)

E

xpe

ri

m

ent

w1 w2 0.2 0.0 0.2 0.4 0.6 0.8 w[s 1] 0 1 2 3 4 5 6 7 8 r (w )

(c)

Si

m

ul

at

ion

w1 w2 0.0 0.2 0.4 0.6 0.8 1.0 w1[s 1] 0.2 0.0 0.2 0.4 0.6 0.8 1.0 w2 [s 1 ]

(b)

n = 26 0.0 0.2 0.4 0.6 0.8 1.0 w1[s 1] 0.2 0.0 0.2 0.4 0.6 0.8 1.0 w2 [s 1 ]

(d)

n = 4684 low high Poi nt cl oud de ns ity

FIG. 5. Distributions of the detected frequencies. (a) Histograms of ω1 (before stimulation) and ω2 (during τφ) obtained from

the experimental data and (b) scatter plot of ω1 versus ω2. The dashed line corresponds to ω2 = ω1. (c) Histograms of ω1

and ω2 resulting from the numerical simulations. (d) Scatter plot of ω1 versus ω2 for the numerical data. The color gradient

indicates the density of the point cloud, calculated by a kernel density estimate of the distribution. Cases in (b) and (d) for which ω1> ω2 (ω1< ω2) are shown in blue (black).

solution of the model is z(t) ≈ −ξ(t)/[(λ − λo) + iωo]. Therefore, the effective frequency is entirely determined by the noise ξ(t), which leads to a small effective fre-quency ω2, corresponding to the peak observed at ω2= 0 in figure 5 (c).

The density of points in the (ω1, ω2)-plane, shown in figure 5 (d), demonstrates that in most cases ω2 < ω1 (blue), with a small fraction of the cells having ω2> ω1 (black). The qualitative resemblance between the mea-surements in figure 5 (a) & (b) and simulations in fig-ure 5 (c) & (d) validates our model. In particular, it justifies our assumption that actin oscillations are slowed down by changing λ→ (λ − λo), close to a Hopf bifurca-tion. Finally, based on the comparison between experi-ments and simulations, our analysis suggests that 1) the coupling strength between actin polymerization and its negative regulators (x(t) and y(t), respectively) does not change during the transient response, and 2) the central frequency ωo does not vary between cells. The

varia-tions in the detected frequencies ω1,2 can be attributed to phase noise, which we discuss in detail in the appendix.

Note: the section ”Long transient and noise scaling: model predictions” is removed and merged with section ”Statistical characterization of transients and noise scal-ing”

VI. CONCLUSIONS

The cAMP-induced dynamic reorganization of the actin cytoskeleton in D. discoideum has been intensively investigated over the past decades [28–30]. Its tempo-ral evolution shows a biphasic profile with a first pro-nounced maximum in the cortical F-actin concentration 5-10 sec after stimulation, followed by a second, less pro-nounced and broader peak of variable duration that may last up to several minutes [13, 14, 31]. Also upstream sig-naling components such as phosphatidylinositol

(10)

(3,4,5)-trisphosphate (PIP3) and the GTP binding protein RacB exhibit a biphasic time profile [15, 16] that has also been addressed by theoretical models [18, 19].

In this work, we extended the investigation of the cAMP-induced actin response, focusing on D. discoideum cells that showed self-sustained polymerization cycles prior to stimulation. Also in the oscillatory cells, a bipha-sic response profile is observed. After stimulation, the actin system transiently leaves the limit cycle and shifts to another dynamical state. This transient can be di-vided into two regimes (biphasic profile) that are charac-terized by the amplitude and phase transient times τaand τφ. Similar to the biphasic profiles reported in the previ-ous literature, we found that the duration of the second regime varies strongly, resulting in a broad distribution of the total transient time between 40 and 200 sec, which corresponds to approximately 3 to 14 oscillation cycles in the pre-stimulation state. The dynamics during the sec-ond regime of the response is characterized by a lower amplitude and frequency.

Previously, we showed that oscillatory actin dynamics in D. discoideum can be described by a noisy Stuart-Landau oscillator [22]. Here we extended this model to account for the experimentally observed cAMP-induced biphasic transient. The extended model suggests that external cAMP stimulation transiently reduces the value of the bifurcation parameter, λ → (λ − λo), such that for most oscillatory cells the transient corresponds to a regime where periodic activity is disrupted or solely

maintained by noise. Note that previous experiments on the cAMP-induced actin response did not specifically fo-cus on oscillatory cells but relied on native D. discoideum populations that are dominated by a majority of non-oscillatory cells. For the non-non-oscillatory cells the bifur-cation parameter is negative, λ < 0. In these cells, a further reduction of λ in response to the cAMP stimulus will be difficult to detect, as no qualitative change in the dynamical state is associated with it.

While the detailed molecular mechanism of the bipha-sic response remains elusive, our results support the hy-pothesis that the cAMP-induced transients observed here primarily reflect the time scale of the signal transduction machinery. This is suggested by observations of intracel-lular signaling patterns in cells that have been treated with Latrunculin A to disrupt the actin cytoskeleton and exhibited time scales that were of the same order as the cAMP-induced transients observed here [32, 33]. Similar time scales also emerged during gradient-induced Ras re-organization or mTORC2 activation [34, 35] suggesting that the transition to a lower amplitude state may be re-lated to the organization of the chemotactic machinery. Also recently proposed models that rely on an oscillatory actin machinery guided by an upstream wave-generating signaling systems support this view [27, 36].

ACKNOWLEDGEMENTS

We acknowledge funding for this work by the DFG SFB 937 ”Collective behavior of soft and biological matter”, project A09.

[1] Nichols J M, Veltman D and Kay R R 2015 Current Opin-ion in Cell Biology 36 7–12

[2] Westendorf C, Negrete J, Bae A J, Sandmann R, Boden-schatz E and Beta C 2013 Proceedings of the National Academy of Sciences of the United States of America 110 3853–3858

[3] Hsu H F, Bodenschatz E, Westendorf C, Gholami A, Pumir A, Tarantola M and Beta C 2017 Physical Review Letters 119 148101

[4] Beta C and Kruse K 2017 Annual Review of Condensed Matter Physics 8 239–264

[5] Antebi Y E, Nandagopal N and Elowitz M B 2017 Cur-rent Opinion in Systems Biology 1 16–24

[6] Bosgraaf L and Van Haastert P J M 2009 PLoS ONE 4 e5253

[7] Bosgraaf L and Van Haastert P J M 2009 PLoS ONE 4 e6842

[8] Beta C, Wyatt D, Rappel W J and Bodenschatz E 2007 Analytical Chemistry 79 3940–3944

[9] Beta C and Bodenschatz E 2011 European Journal of Cell Biology 90 811–816

[10] Prassler J, Murr A, Stocker S, Faix J, Murphy J and Marriott G 1998 Mol Biol Cell 9 545–559

[11] Schneider N, Weber I, Faix J, Prassler J, M¨ uller-Taubenberger A, K¨ohler J, Burghardt E, Gerisch G and Marriott G 2003 Cell Motility 56 130–139

[12] Etzrodt M, Ishikawa H, Dalous J, Annette M, Bretschnei-der T and Gerisch G 2006 Febs Lett 580 6707–6713 [13] Hall A, Schlein A and Condeelis J 1988 Journal of

Cel-lular Biochemistry 37 285–299

[14] Chen L, Janetopoulos C, Huang Y, Iijima M, Borleis J and Devreotes P 2003 Molecular Biology of the Cell 14 5028–5037

[15] Park K C, Rivero F, Meili R, Lee S, Apone F and Firtel R A 2004 EMBO Journal 23 4177–4189

[16] Postma M, Roelofs J, Goedhart J, Gadella T W J, Visser A J W G and Van Haastert P J M 2003 Molecular Biology of the Cell 14 5019–5027

[17] Postma M, Roelofs J, Goedhart J, Loovers H M, Visser A J W G and Van Haastert P J M 2004 Journal of Cell Science 117 2925–2935

[18] Xiong Y, Huang C H, Iglesias P A and Devreotes P N 2010 Proceedings of the National Academy of Sciences 107 17079–17086

[19] Hecht I, Kessler D A and Levine H 2010 Physical Review Letters 104 158301

[20] Guckenheimer J and Holmes P, Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields (Springer New York, 2002).

[21] Strogatz S, Nonlinear Dynamics and Chaos: with Appli-cations to Physics, Biology, Chemistry and Engineering (Avalon Publishing, 2014).

(11)

[22] Negrete J, Pumir A, Hsu H F, Westendorf C, Tarantola M, Beta C and Bodenschatz E 2016 Physical Review Let-ters 117

[23] Diez S, Gerisch G, Anderson K, Annette M and Bretschneider T 2005 Proceedings of the National Acadamy of Sciences of the United States of America 102 7601–7606

[24] Gerisch G, Albrecht R, Heizer C, Hodgkinson S and Ma-niak M 1995 Curr Biology Cb 5 1280–5

[25] Konzok A, Weber I, Simmeth E, Hacker U, Maniak M and Annette M 1999 J Cell Biology 146 453–464 [26] Bae A, Beta C and Bodenschatz E 2009 Lab on a Chip

9 3059–3065

[27] Huang C H, Tang M, Shi C, Iglesias P A and Devreotes P N 2013 Nature Cell Biology 15 1307–1316

[28] McRobbie S J and Newell P C 1983 Biochemical and Biophysical Research Communications 115 351–359 [29] McRobbie S J and Newell P C 1984 Journal of Cell

Sci-ence 68 139–151

[30] Yumura S 1994 Cell Structure and Function 19 143–151 [31] Condeelis J, Hall A, Bresnick A, Warren V, Hock R,

Ben-nett H and Ogihara S 1988 Cell Motility 10 77–90 [32] Gerisch G, Bretschneider T, M¨uller-Taubenberger A,

Simmeth E, Ecke M, Diez S and Anderson K 2004 Bio-physical Journal 87 3493–3503

[33] Arai Y, Shibata T, Matsuoka S, Sato M J, Yanagida T and Ueda M 2010 Proceedings of the National Academy of Sciences of the United States of America 107 12399– 12404

[34] Kortholt A, Keizer-Gunnink I, Kataria R and Haastert P J M V 2013 Journal of Cell Science 126 4502–4513 [35] Senoo H, Kamimura Y, Kimura R, Nakajima A, Sawai

S, Sesaki H, and Iijima M 2019 Nature Cell Biology 21 867–878

[36] Ecke M and Gerisch G 2019 Small GTPases 10 72–80

APPENDIX A. Time distributions

Also for the statistics of transient time scales and amplitude fluctuations the model predictions agree well with our experimental findings. Figures 6 (a) & (b) show a comparison between the distributions of τa and τφ obtained from numerical simulations (orange lines) and from experiments (blue bars). As explained in Section III, the distribution of the total transient time T = τa+ τφin our numerical studies is prescribed by the experiment. The resulting distributions in figure 6 (a) and (b) show that our model correctly divides the total transient time T into τaand τφ. In our model, τais deter-mined by the shape of the input s(t) and the parameters in equations (5–6). Thus, τa reflects the response prop-erties and time scales of the cytoskeleton in response to a pulse of cAMP. On the other hand, the phase transient time τφ is determined by the time interval during which w(t) exceeds the threshold K in response to a pulse of cAMP.

Fitting procedure

In this section we describe the detailed procedure to constrain the parameters of equations (5-6). The shape of s(t), given by equation (6), was obtained by fitting the response of RasC to short-time stimuli of cAMP. The fitted data originated from an experiment described in [27]. Figure 7 (a) shows the response of the fluores-cently labeled Ras binding domain (RBD-GFP) to two consecutive 2 sec pulses of cAMP. The fit of these data (black line) resulted in the parameters so = 1/4.69 and τs= 1.75.

The distribution of relaxation time τ in equation (7) was assumed to decay exponentially, see equation (9), ρ(τ ) = Θ(τ −τ1)

τ2 e

−(τ −τ1)/τ2. It is related to the

distribu-tion ρ(T ) of the total transient time T by

ρ(T ) = ρ(τ (T )) dτ (T ) dT . (11)

The dependence of the total transient time T (τ ) on the relaxation time τ , shown in figure 7 (b), was obtained by fixing K to a value of 0.035, see equation (8). The fitting parameters τ1= 7.8 sec and τ2= 20.75 sec in the distribution ρ(τ ) are chosen such that the distribution ρ(T ) resulting form equation (11) fits the experimentally obtained distribution of the total transient time T = τa+ τφ, which is shown in figure 7 (c).

For completeness, we recall that the distribution of the dimensionless parameter Λ = λ/√gD has been simplified here as a normal distribution, see equation (10).

The oscillator frequency ωoof equation (5) is set to the average observed frequency ω = ωo. The parameter

(12)

0 20 40 60 80 100 τa[s] 0 2 4 6 8 10 12 14 Counts

(a)

Experiment Simulation 0 40 80 120 160 200 τφ[s] 0 2 4 6 8 Counts

(b)

Experiment Simulation 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 ρ a ) 0.000 0.005 0.010 0.015 0.020 ρ (τφ )

FIG. 6. Comparison of the experimentally and numerically derived amplitude and phase transient times. The experimental results are displayed as a histogram (blue) and the numerical results as a probability distribution function (orange) for τa(a)

and τφ(b).

λo(equation (8)) was estimated by plotting µ (1)

a against µ(2)a , see figure 8 (a), and comparing it with the theoreti-cally derived curves (figure 8 (b–e)). The transition from a steep to a smaller slope, as µ(1)a is increased, is most pronounced in figure 8 (e) with λo = 6√gD. Therefore, we chose the value λo= 6√gD as a good approximation. Finally, the remaining free parameter in equation (5) was set to τr= 1/2 sec.

Phase slips

The rotation number in the numerical simulations, de-fined as the number of oscillator cycles, depends on the sampling acquisition rate. The amplitude A(t) and phase φ(t), plotted in figure 9 (a) and (b), were calculated with a high sampling rate (∆t = 0.01 sec). The phase shows several phase slips (i.e. discontinuities), which are marked by black arrows in figure 9 (b). These re-sult from amplitude fluctuations close to A(t) = 0. The detected number of oscillator cycles in this case yields 80/2π = 12.73.

On the other hand, the number of values close to A(t) = 0 is reduced if the numerically derived time series are downsampled to the same rate as in experi-ment (∆t = 1.0 sec). The number of phase slips dimin-ishes dramatically in this case and as a consequence the number of oscillator cycles is reduced to 30/2π = 2.77. Therefore, the detected frequency ω, obtained by fitting a straight line to the phase φ(t) is lower for the coarse sampled signal (compare figure 9 (c) and (d), where the red dashed line indicates the linear fit). Furthermore, the fit accuracy is increased by the coarse sampling. There-fore, we used the experimental sampling rate to analyze the detected frequency distribution in both the experi-mentally and the numerically obtained time series.

Correlation of amplitude mean and transient timescales

The dependence of the mean fluorescence intensity (µ(1)a ) prior to stimulation to the different response timescales (τa and τφ) are shown in figure 10. No sig-nificant correlations were found.

Time series

All experimental time series, which were identified as oscillatory before stimulation, along with their respec-tive autocorrelation functions (AC(t)) are shown in fig-ures 11– 14. The AC(t) are given only for the initial 100 sec of the respective time series.

(13)

FIG. 7. Fitting procedure and derivation of parameters, based thereon. (a) Response of the chemotactic signaling system, measured by the translocation of the fluorescently labeled Ras binding domain (RBD-GFP) to two differently spaced pulses of cAMP. The signal input s(t) was obtained by fitting (black line) the first responses. Figure reprinted and modified with permission from [27]. (b) Functional relation between the total transient time and the relaxation time τ , given a fixed value of K. (c) Fit of the total transient time distribution resulting from ρ(τ ) =Θ(τ −τ1)

τ2 exp

−(τ −τ1)/τ2 and equation (11). (d) Fit of a normal distribution with zero mean and a variance of 2.45 to the distribution of λ/√gD. Reprinted with permission from [22].

0 50 100 150 200 250 µa(1) 0 20 40 60 80 100 µ (2 ) a (a) n = 26 Experiment 0.0 0.1 0.2 µa(λ) 0.0 0.1 0.2 0.3 µa − 3 √ gD ) (b) n = 1501 Simulation 0.0 0.1 0.2 µa(λ) 0.0 0.1 0.2 0.3 µa − 4 √ gD ) (c) n = 1401 0.0 0.1 0.2 µa(λ) 0.0 0.1 0.2 0.3 µa − 5 √ gD ) (d) n = 1301 0.0 0.1 0.2 µa(λ) 0.0 0.1 0.2 0.3 µa − 6 √ gD ) (e) n = 1201

FIG. 8. Comparison of experimentally and numerically derived amplitude means. (a) Relation of the measured amplitude means before stimulation (µ(1)a ) and during τφ (µ(2)a ). (b-e) Amplitude means obtained by evaluating equation (3) with the

following parameter set: (b) λo= 3

√ gD, (c) λo= 4 √ gD, (d) λo= 5 √ gD and (e) λo= 6 √ gD.

(14)

0 20 40 60 80 100 Time [s] −0.1 0.0 0.1 0.2 0.3 0.4 A(t) (a) Fine sampling 0 20 40 60 80 100 Time [s] 0 20 40 60 80 100 φ (t ) (b) 0 20 40 60 80 100 Time [s] −0.1 0.0 0.1 0.2 0.3 0.4 A(t) (c) Coarse sampling 0 20 40 60 80 100 Time [s] 0 20 40 60 80 100 φ (t ) (d)

FIG. 9. Comparison of fine and course sampling. Amplitude A(t) (a) and phase φ(t) (b) resulting from a numerical simulation with fine sampling (∆t = 0.01 sec). Black arrows mark phase slips. (c,d) Amplitude A(t) and phase φ(t) obtained from a numerical simulation with course sampling (∆t = 1.0 sec). Red dashed lines in (b) and (d) indicate a linear fit to the phase φ(t). 0 50 100 150 200 250 µa(1) 20 40 60 80 100 τa

(a)

n = 26 0 50 100 150 200 250 µa(1) 0 25 50 75 100 125 150 175 200 τφ

(b)

n = 17 0 50 100 150 200 250 µa(1) 25 50 75 100 125 150 175 200 225 τa + τφ

(c)

n = 17

FIG. 10. Dependence of the mean fluorescence intensity (µ(1)a ) prior to stimulation to the different response timescales. (a)

Comparison of µ(1)a to the amplitude transient time τa, (b) to the phase transient time τφ, and (c) to the sum of both transient

times.

Comparison between simulations and experimental data

We have simulated 4684 realizations of the model given by equations (5)–(6). To check if the model reproduced the main features of the experimental data, we defined

the following error function

E(i,j)= |A (i) exp− A (j) sim| |A(i)exp| |ψ(i)exp− ψ (j) sim| |ψexp(i)| , (12)

(15)

−400 −200 0 200 Intensity [au] −400 −200 0 200 400 Intensity [au] −600 −400 −200 0 200 Intensity [au] −300 −200 −100 0 100 Intensity [au] −300 −200 −100 0 100 200 Intensity [au] −500 −250 0 250 500 Intensity [au] 0 100 200 300 400 Time [s] −300 −200 −100 0 100 200 Intensity [au] −1 0 1 2 C ) ×106 −2 0 2 4 C ) ×106 −2 0 2 4 C ) ×105 −0.5 0.0 0.5 1.0 1.5 C ) ×105 −1 0 1 2 C ) ×105 −2 0 2 4 C ) ×105 0 20 40 60 80 100 τ[s] −0.5 0.0 0.5 1.0 C ) ×105 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=28s τφ=60s −100 −50 0 50 φ (t )− ω0 t− φ0 τa=42s τφ=134s −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=22s τφ=41s −20 −10 0 10 φ (t )− ω0 t− φ0 τa=18s τφ=60s −15 −10 −5 0 5 φ (t )− ω0 t− φ0 τa=26s τφ=26s −20 −10 0 10 20 φ (t )− ω0 t− φ0 τa=27s τφ=28s 0 100 200 300 400 Time [s] −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=24s τφ=58s

FIG. 11. Individual fluorescent time traces used for data analysis. Left column shows the individual time signals after detrending and the middle column the corresponding autocorrelation functions (C(τ )), calculated over the first 100 sec of the time signal. The right column gives the evolution of the phase. All cases shown in this figure are decelerating the phase. The time of stimulation (t = 100 sec) is marked by a gray bar in the time signals.

where A(i)expand ψ(i)expcorrespond to the experimental time traces of the amplitude and phase drifts, respectively, and the index i corresponds to the experiment number (1 ≤ i ≤ 22). The amplitude A(i)exp was normalized by dividing it with its maximum value. At the same time,

A(j)sim and ψsim(j) correspond to the simulated time traces of the amplitude and phase drifts, respectively, and the index j corresponds to the simulation number (1≤ j ≤ 4684). We applied the same criteria to the simulated realizations concerning the decay of the autocorrelation function and applied the same sampling rate (see main

(16)

−400 −200 0 200 Intensity [au] −300 −200 −100 0 100 200 Intensity [au] −400 −200 0 200 400 Intensity [au] −300 −200 −100 0 100 200 Intensity [au] −300 −200 −100 0 100 200 Intensity [au] −400 −200 0 200 Intensity [au] 0 100 200 300 400 Time [s] −500 −250 0 250 500 Intensity [au] −1 0 1 2 3 C ) ×105 −1 0 1 2 C ) ×105 −1 0 1 2 C ) ×106 −0.5 0.0 0.5 1.0 C ) ×106 −1 0 1 2 3 C ) ×105 −2 0 2 4 C ) ×105 0 20 40 60 80 100 τ[s] −1 0 1 2 C ) ×105 −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=26s τφ=n.a. −10 0 10 20 30 φ (t )− ω0 t− φ0 τa=31s τφ=68s −15 −10 −5 0 5 φ (t )− ω0 t− φ0 τa=27s τφ=56s −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=48s τφ=65s −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=34s τφ=151s −40 −30 −20 −10 0 10 20 φ (t )− ω0 t− φ0 τa=25s τφ=188s 0 100 200 300 400 Time [s] −5 0 5 10 15 φ (t )− ω0 t−

φ0 New base line

Phase slip

τa=24s

τφ=17s

FIG. 12. Individual fluorescent time traces used for data analysis. (Continued from figure 11)

text for details).

We have assigned to each experimental time trace a numerical time trace by finding min[E(i,j)]. In figure 15, we show three examples of experimental time trace along with the numerical simulations that were chosen in this

way. The numerical time traces reproduce also details of the experimental traces. In particular, we note that in a small set of numerical simulations, the frequency increased after stimulation, i.e. ω2 > ω1, similar to the experimental results. This demonstrates that both cases observed in experiment, ω2< ω1and ω2> ω1, emerge as part of the same oscillatory mechanism.

(17)

−400 −200 0 200 400 Intensity [au] −400 −200 0 200 Intensity [au] −400 −200 0 200 Intensity [au] −300 −200 −100 0 100 200 Intensity [au] −400 −200 0 200 Intensity [au] −200 −100 0 100 200 Intensity [au] 0 100 200 300 400 Time [s] −200 −100 0 100 200 Intensity [au] −2 −1 0 1 2 C ) ×106 −0.5 0.0 0.5 1.0 C ) ×105 −1 0 1 2 C ) ×105 −0.5 0.0 0.5 1.0 1.5 C ) ×105 −2 0 2 4 C ) ×105 −0.5 0.0 0.5 1.0 C ) ×105 0 20 40 60 80 100 τ[s] −0.5 0.0 0.5 1.0 C ) ×105 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=27s τφ=n.a. −15 −10 −5 0 5 φ (t )− ω0 t− φ0 τa=25s τφ=16s −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=25s τφ=n.a. −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=19s τφ=48s −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=24s τφ=n.a. −20 0 20 40 φ (t )− ω0 t− φ0

New base line Phase slip τa=23s τφ=21s 0 100 200 300 400 Time [s] −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=22s τφ=72s

(18)

0 100 200 300 400 Time [s] −200 −100 0 100 200 Intensity [au] −500 −250 0 250 500 Intensity [au] −100 −50 0 50 Intensity [au] −400 −200 0 200 Intensity [au] 0 100 200 300 400 Time [s] −200 −100 0 100 200 Intensity [au] 0 20 40 60 80 100 τ[s] −1 0 1 2 3 C ) ×104 −2 0 2 4 6 C ) ×105 −2 0 2 4 C ) ×104 −1 0 1 2 3 C ) ×105 0 20 40 60 80 100 τ[s] −1 0 1 2 C ) ×105 0 100 200 300 400 Time [s] −30 −20 −10 0 10 φ (t )− ω0 t− φ0 τa=21s τφ=n.a. −50 0 50 100 φ (t )− ω0 t− φ0 τa=26s τφ=n.a. −50 0 50 100 150 φ (t )− ω0 t− φ0 τa=89s τφ=n.a. −20 0 20 40 φ (t )− ω0 t− φ0 τa=16s τφ=n.a. 0 100 200 300 400 Time [s] −5 0 5 10 15 φ (t )− ω0 t− φ0 τa=27s τφ=n.a.

FIG. 14. Individual fluorescent time traces used for data analysis. (Continued from figure 11). The lower 4 rows show the cases of accelerating phase after stimulation.

(19)

0 100 200 300 400 Time (s) −1.0 −0.5 0.0 0.5 Intensity [au] (a) 0 100 200 300 400 Time [s] −0.6 −0.4 −0.2 0.0 0.2 x(t) 0 100 200 300 400 Time (s) −1.0 −0.5 0.0 0.5 Intensity [au] (b) 0 100 200 300 400 Time [s] −0.6 −0.4 −0.2 0.0 0.2 x(t) 0 100 200 300 400 Time (s) −1.0 −0.5 0.0 0.5 Intensity [au] (c) 0 100 200 300 400 Time [s] −0.6 −0.4 −0.2 0.0 0.2 x(t) 0 100 200 300 400 Time [s] −20 −10 0 10 20 φ(t) −ω0t − φ0 Experiment Theory 0 100 200 300 400 Time [s] −40 −20 0 20 φ(t) −ω0t − φ0 Experiment Theory 0 100 200 300 400 Time [s] −30 −20 −10 0 10 φ(t) −ω0t − φ0 Experiment Theory 0 100 200 300 400 Time [s] 0.00 0.25 0.50 0.75 1.00 A(t) 0 100 200 300 400 Time [s] 0.00 0.25 0.50 0.75 1.00 A(t) 0 100 200 300 400 Time [s] 0.00 0.25 0.50 0.75 1.00 A(t)

FIG. 15. Comparison of experimental and numerical results, selected according to minimization of the error function (12) as explained in the text. In (a), (b), and (c), three examples are displayed. On the left, blue and orange curves display the experimental and numerical time traces, respectively. In the middle and on the right, phase and amplitude drifts are shown, respectively, using the same color coding.

Figure

FIG. 1. Experimental set-up and data analysis. (a) Exemplary image series. The D. discoideum cell is centered within a 48 × 48 µm imaging region
FIG. 2. Examples of single cell D. discoideum LimE-mRFP responses to a short pulse of cAMP
FIG. 3. Model for the response of the D. discoideum actin cytoskeleton to short time cAMP pulses
FIG. 4. Mean, standard deviation and coefficient of variation in the amplitude fluctuations
+7

Références

Documents relatifs

As shown above, the actin cytoskeleton is a complex system, whose dynamics depends on multiple factors: i) nucleotide state of actin monomers and filament protomers; ii) age

Actin-rich membrane protrusions induced by MIM-B in serum-starved Swiss 3T3 cells look like ruffles and lamellipodia, structures generally observed following the activation of the

Similarly to micropost based adhesion force measurements, the direction central adhesion forces are not necessarily directed inward (left image). Right) View of the model:

mutations on actin filaments, we studied SA levels, callose accumulation and gene expression 344. responses to latB in

(B) To assess identity of the factor detected in the nuclear extracts only after incubation with PRL (A and lanes 8 and 9), a supershift experiment using monoclonal antibodies

Expression of EGFP β -actin in mammalian cells In order to visualize the actin cytoskeleton in stationary and migrating mammalian cells, the enhanced green fluorescent protein

In neuroendocrine PC-12 cells, evanescent-field fluorescence microscopy was used to track motions of green fluorescent protein (GFP)-labeled actin or GFP-labeled secretory granules in

discoideum cells and as in early endosomes (Figure 2), tubular structures that extend from late endosomes containing ag- gregated gold particles can also be documented by EM (Figure