43
Utilization of extracellular information before ligand-receptor binding reaches equilibrium expands and shifts the input dynamic range Alejandra C. Ventura a,b , Alan Bush a,b , Gustavo Vasen a,b , Matías A. Goldín a,b , Brianne Burkinshaw a,b , Nirveek Bhattacharjee c , Albert Folch c , Roger Brent d , Ariel Chernomoretz e,f,g , and Alejandro Colman-Lerner a,b,1 a Institute of Physiology, Molecular Biology, and Neuroscience (IFIBYNE), University of Buenos Aires (UBA)-National Scientific and Technical Research Council (CONICET), b Department of Physiology, Molecular, and Cell Biology, School of Exact and Natural Sciences (FCEN), e Physics Institute of Buenos Aires (IFIBA), CONICET, and f Department of Physics, FCEN, UBA, C1428EGA Buenos Aires, Argentina; g Fundación Instituto Leloir, C1405BWE Buenos Aires, Argentina; c Department of Bioengineering, University of Washington, Seattle, WA 98195; and d Division of Basic Sciences, Fred Hutchinson Cancer Research Center, Seattle, WA 98109 Edited by Peter N. Devreotes, Johns Hopkins University School of Medicine, Baltimore, MD, and approved July 30, 2014 (received for review December 9, 2013) Cell signaling systems sense and respond to ligands that bind cell surface receptors. These systems often respond to changes in the concentration of extracellular ligand more rapidly than the ligand equilibrates with its receptor. We demonstrate, by modeling and experiment, a general systems levelmechanism cells use to take advantage of the information present in the early signal, before receptor binding reaches a new steady state. This mechanism, pre- equilibrium sensing and signaling (PRESS), operates in signaling sys- tems in which the kinetics of ligand-receptor binding are slower than the downstream signaling steps, and it typically involves transient activation of a downstream step. In the systems where it operates, PRESS expands and shifts the input dynamic range, allowing cells to make different responses to ligand concentrations so high as to be otherwise indistinguishable. Specifically, we show that PRESS applies to the yeast directional polarization in response to pheromone gra- dients. Consideration of preexisting kinetic data for ligand-receptor interactions suggests that PRESS operates in many cell signaling sys- tems throughout biology. The same mechanism may also operate at other levels in signaling systems in which a slow activation step cou- ples to a faster downstream step. cellular signaling | binding kinetics | doseresponse D etecting and responding to a chemical gradient is a central feature of a multitude of biological processes (1). For this behavior, organisms use signaling systems that sense information about the extracellular world, transmit this information into the cell, and orchestrate a response. Measurements of the direction and proximity of the extracellular stimuli usually rely on the binding of diffusing chemical particles (ligands) to specific cell surface receptors. Different organisms have evolved different strategies to make use of this information. Small motile organ- isms, including certain bacteria, use a temporal sensing strategy, measuring and comparing concentration signals over time along their swimming tracks (2). In contrast, some eukaryotic cells, including Saccharomyces cerevisiae, are sufficiently large to im- plement a spatial sensing mechanism, measuring concentration differences across their cell bodies (3). The observation that some eukaryotes that use spatial sensing exhibit remarkable precision in response to shallow gradients (12% differences in ligand concentration between front and rear) (4, 5) has led to several proposed models in which large amplification is achieved by positive feedback loops in the signaling pathways triggered by the ligand-receptor binding (6, 7). Here, we describe a different mechanism, dependent on ligand-receptor binding dynamics, which improves gradient sensing when the concentra- tion of external ligand is close to saturation. We use the budding yeast S. cerevisiae to study the efficiency of this mechanism. Haploid yeast cells exist in two mating types, MATa and MATα (also referred to as a and α cells). Mating occurs when a and α cells sense each others secreted mating pheromones: a- factor and α-factor (αF) (8). The pheromone secreted by the nearby mating partner diffuses, forming a gradient surrounding the sensing cell. Pheromone binds a membrane receptor, Ste2, in MATa yeast (9) that activates a pheromone response system (PRS), which cells use to decide whether to fuse with a mating partner or not. At high enough αF concentrations, cells develop a polarized chemotropic growth toward the pheromone source (4). To do that, the nonmotile yeast determines the direction of the potential mating partner measuring on which side there are more bound pheromone receptors, which are initially distributed homogeneously on the cell surface (10). However, this sensing modality can only work when external pheromone is non- saturating: If all receptors are bound, cells should not be able to determine the direction of the gradient. Surprisingly, even at high pheromone concentrations, yeast tend to polarize in the correct direction (4, 11). Different amplification mechanisms have been proposed to account for the conversion of small differences in ligand concentration across the yeast cell, as is the case for dense mating mixtures, into chemotropic growth (6). We previously studied induction of reporter gene output by the PRS after step increases in the concentration of αF. We found large cell-to-cell variability, the bulk of which was due to large differences in the ability of individual cells to send signal Significance Many cell decisions depend on precise measurements of ex- ternal ligands reversibly bound to receptors. Yeast cells orient in gradients of sex pheromone detecting differences in the amount of ligand-receptor complex. However, yeast can orient in gradients with nearly all receptors occupied. We describe a general systems-level mechanism, pre-equilibrium sensing and signaling (PRESS), which overcomes this saturation limit by shifting and expanding the input dynamic range to which cells can respond. PRESS requires that events downstream of the receptor be transient and faster than the time required for the receptor to reach equilibrium binding. Experiments and simu- lations show that PRESS operates in yeast and may help cells orient in gradients. Many ligand-receptor interactions are slow, suggesting that PRESS is widespread throughout eukaryotes. Author contributions: A.C.V., A.B., and A.C.-L. designed research; A.C.V., A.B., G.V., M.A.G., A.C., and A.C.-L. performed research; B.B., N.B., A.F., and R.B. contributed new reagents/ analytic tools; A.C.V., A.B., G.V., A.C., and A.C.-L. analyzed data; and A.C.V., R.B., and A.C.-L. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. 1 To whom correspondence should be addressed. Email: [email protected]. This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. 1073/pnas.1322761111/-/DCSupplemental. www.pnas.org/cgi/doi/10.1073/pnas.1322761111 PNAS Early Edition | 1 of 10 BIOPHYSICS AND COMPUTATIONAL BIOLOGY PNAS PLUS

Utilization of extracellular information before ligand-receptor binding reaches equilibrium expands and shifts the input dynamic range

  • Upload
    uba

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Utilization of extracellular information beforeligand-receptor binding reaches equilibrium expandsand shifts the input dynamic rangeAlejandra C. Venturaa,b, Alan Busha,b, Gustavo Vasena,b, Matías A. Goldína,b, Brianne Burkinshawa,b,Nirveek Bhattacharjeec, Albert Folchc, Roger Brentd, Ariel Chernomoretze,f,g, and Alejandro Colman-Lernera,b,1

aInstitute of Physiology, Molecular Biology, and Neuroscience (IFIBYNE), University of Buenos Aires (UBA)-National Scientific and Technical Research Council(CONICET), bDepartment of Physiology, Molecular, and Cell Biology, School of Exact and Natural Sciences (FCEN), ePhysics Institute of Buenos Aires (IFIBA),CONICET, and fDepartment of Physics, FCEN, UBA, C1428EGA Buenos Aires, Argentina; gFundación Instituto Leloir, C1405BWE Buenos Aires, Argentina;cDepartment of Bioengineering, University of Washington, Seattle, WA 98195; and dDivision of Basic Sciences, Fred Hutchinson Cancer Research Center,Seattle, WA 98109

Edited by Peter N. Devreotes, Johns Hopkins University School of Medicine, Baltimore, MD, and approved July 30, 2014 (received for review December 9, 2013)

Cell signaling systems sense and respond to ligands that bind cellsurface receptors. These systems often respond to changes in theconcentration of extracellular ligand more rapidly than the ligandequilibrates with its receptor. We demonstrate, by modeling andexperiment, a general “systems level” mechanism cells use to takeadvantage of the information present in the early signal, beforereceptor binding reaches a new steady state. This mechanism, pre-equilibrium sensing and signaling (PRESS), operates in signaling sys-tems inwhich the kinetics of ligand-receptor binding are slower thanthe downstream signaling steps, and it typically involves transientactivation of a downstream step. In the systems where it operates,PRESS expands and shifts the input dynamic range, allowing cells tomake different responses to ligand concentrations so high as to beotherwise indistinguishable. Specifically,we showthat PRESS appliesto the yeast directional polarization in response to pheromone gra-dients. Consideration of preexisting kinetic data for ligand-receptorinteractions suggests that PRESS operates in many cell signaling sys-tems throughout biology. The same mechanismmay also operate atother levels in signaling systems in which a slow activation step cou-ples to a faster downstream step.

cellular signaling | binding kinetics | dose–response

Detecting and responding to a chemical gradient is a centralfeature of a multitude of biological processes (1). For this

behavior, organisms use signaling systems that sense informationabout the extracellular world, transmit this information into thecell, and orchestrate a response. Measurements of the directionand proximity of the extracellular stimuli usually rely on thebinding of diffusing chemical particles (ligands) to specific cellsurface receptors. Different organisms have evolved differentstrategies to make use of this information. Small motile organ-isms, including certain bacteria, use a temporal sensing strategy,measuring and comparing concentration signals over time alongtheir swimming tracks (2). In contrast, some eukaryotic cells,including Saccharomyces cerevisiae, are sufficiently large to im-plement a spatial sensing mechanism, measuring concentrationdifferences across their cell bodies (3).The observation that some eukaryotes that use spatial sensing

exhibit remarkable precision in response to shallow gradients (1–2%differences in ligand concentration between front and rear) (4, 5)has led to several proposed models in which large amplification isachieved by positive feedback loops in the signaling pathwaystriggered by the ligand-receptor binding (6, 7). Here, we describea different mechanism, dependent on ligand-receptor bindingdynamics, which improves gradient sensing when the concentra-tion of external ligand is close to saturation. We use the buddingyeast S. cerevisiae to study the efficiency of this mechanism.Haploid yeast cells exist in two mating types, MATa and

MATα (also referred to as a and α cells). Mating occurs when

a and α cells sense each other’s secreted mating pheromones: a-factor and α-factor (αF) (8). The pheromone secreted by thenearby mating partner diffuses, forming a gradient surroundingthe sensing cell. Pheromone binds a membrane receptor, Ste2, inMATa yeast (9) that activates a pheromone response system(PRS), which cells use to decide whether to fuse with a matingpartner or not. At high enough αF concentrations, cells developa polarized chemotropic growth toward the pheromone source(4). To do that, the nonmotile yeast determines the direction ofthe potential mating partner measuring on which side there aremore bound pheromone receptors, which are initially distributedhomogeneously on the cell surface (10). However, this sensingmodality can only work when external pheromone is non-saturating: If all receptors are bound, cells should not be able todetermine the direction of the gradient. Surprisingly, even at highpheromone concentrations, yeast tend to polarize in the correctdirection (4, 11). Different amplification mechanisms have beenproposed to account for the conversion of small differences inligand concentration across the yeast cell, as is the case for densemating mixtures, into chemotropic growth (6).We previously studied induction of reporter gene output by

the PRS after step increases in the concentration of αF. Wefound large cell-to-cell variability, the bulk of which was due tolarge differences in the ability of individual cells to send signal

Significance

Many cell decisions depend on precise measurements of ex-ternal ligands reversibly bound to receptors. Yeast cells orientin gradients of sex pheromone detecting differences in theamount of ligand-receptor complex. However, yeast can orientin gradients with nearly all receptors occupied. We describea general systems-level mechanism, pre-equilibrium sensingand signaling (PRESS), which overcomes this saturation limit byshifting and expanding the input dynamic range to which cellscan respond. PRESS requires that events downstream of thereceptor be transient and faster than the time required for thereceptor to reach equilibrium binding. Experiments and simu-lations show that PRESS operates in yeast and may help cellsorient in gradients. Many ligand-receptor interactions are slow,suggesting that PRESS is widespread throughout eukaryotes.

Author contributions: A.C.V., A.B., and A.C.-L. designed research; A.C.V., A.B., G.V., M.A.G.,A.C., and A.C.-L. performed research; B.B., N.B., A.F., and R.B. contributed new reagents/analytic tools; A.C.V., A.B., G.V., A.C., and A.C.-L. analyzed data; and A.C.V., R.B., and A.C.-L.wrote the paper.

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.1To whom correspondence should be addressed. Email: [email protected].

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1322761111/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1322761111 PNAS Early Edition | 1 of 10

BIOPH

YSICSAND

COMPU

TATIONALBIOLO

GY

PNASPL

US

through the system and in their general capacity to expressproteins (12). The level of induced gene expression matches wellthe equilibrium binding curve of αF to receptor (13, 14), a phe-nomenon known as dose–response alignment (DoRA), commonto many other signaling systems (14). In the PRS, DoRA persistsfor several hours of stimulation.During these studies, we realized that the binding dynamics of

αF to its receptor is remarkably slow: At concentrations near thedissociation constant (Kd), binding takes about 20 min to reach90% of the equilibrium level (15, 16). This dynamics is slowrelative not only to the 90-min cell division cycle but also to thepheromone-dependent activation of the mitogen-activated pro-tein kinase (MAPK) Fus3, which takes 2 to 5 min to reachsteady-state levels (14). An unavoidable conclusion is that themachinery downstream of the αF receptor must be using pre-equilibrium binding information for its operation.This observation led us to study the consequences of fast and slow

ligand-receptor dynamics on the ability of cells to sense extracellularcues. In biology, the rates of ligand binding and unbinding to mem-brane receptors span a large range, including many cases with dy-namics similar to, or even slower than, that of mating pheromone(e.g., rates for EGF, insulin, glucagon, IFN-α1a, and IL-2 in Table 1).Our study revealed amode of sensing that can greatly increase the

ability of cells to discriminate doses at high ligand concentrations.

ResultsLigand-Receptor Binding Dose–Response Curve Changes Over Time.We consider the time evolution of occupied receptor at differentdoses of ligand for the simple case of one-step binding described by

L+R →k1

←k−

C;

where L is the ligand, R is the receptor, C is the ligand-receptorcomplex, and k+ and k− are the binding and unbinding rates,

respectively. Assuming free L is not significantly affected bythe reaction, binding over time may be described by

Cðt;LÞ =CeqðLÞ p�1− eð−t=τðLÞÞ�;

with

Ceq =RtotL

L+Kd;

and

τ=1

k− +L p k+:

Ceq is the equilibrium value of C, Rtot is the total of amount ofreceptor, and τ is the exponential time constant (time at whichC reaches ∼63.2% of the steady-state value). Thus, the time evo-lution of C depends on ligand concentration: The higher the con-centration of ligand, the faster binding reaches equilibrium (Fig.1A). Similarly, plotting C vs. L at different times shows that theEC50 (concentration of the ligand that occupies 50% of the recep-tors) of the binding curve is high at early times but becomes pro-gressively lower as time passes (Fig. 1B). As a result, beforebinding reaches equilibrium (t∞), the receptor is sensitive in a re-gion of ligand concentrations that will later saturate the receptor(SI Appendix, section 1). For example, in the case of αF, two highconcentrations, L1 and L2 (55 Kd and 80 Kd), result in a 12.7%difference in occupied receptors at 10 s, and in only a 0.56%difference at equilibrium (Fig. 1B).To verify experimentally our theoretical analysis suggesting

a slow reduction over time in the EC50 of the dose–response curvefor receptor binding, we measured the binding dynamics of fluo-rescently tagged αF (37) to live yeast by fluorescence cytometry(Fig. 1C). To minimize ligand depletion at low concentrations, we

Table 1. Sample receptor/ligand binding parameters

Receptor Ligand Cell type k− (1/s) Kd (M) τ (at L = Kd), s Ref.

Fce IgE Human basophils 2.50E-05 4.80E-10 20,000.00 (17)Fcγ 2.4G2 monoclonal Fab Mouse macrophage 3.80E-05 7.70E-10 13,157.89 (18)Canabinoid receptor CP55,940 Rat brain 1.32E-04 2.10E-08 3,787.88 (19)IL-2 receptor IL-2 T cells 2.00E-04 7.40E-12 2,500.00 (20)α1-Adrenergic Prazosin BC3H1 3.00E-04 7.50E-11 1,666.67 (21)Glucagon receptor Glucagon Rat hepatocytes 4.30E-04 3.06E-10 1,162.79 (22)Formyl peptide receptor (FPR) fMLP Rat neutrophils 5.50E-04 3.45E-08 909.09 (23)Ste2 (αF receptor) αF S. cerevisiae 1.00E-03 5.50E-09 500.00 (15, 16)IFN Human IFN-α1a A549 1.20E-03 3.30E-10 416.67 (24)Transferrin Transferrin HepG2 1.70E-03 3.30E-08 294.12 (25)EGF receptor EGF Fetal rat lung 2.00E-03 6.70E-10 250.00 (26)TNF TNF A549 2.30E-03 1.50E-10 217.39 (24)Insulin receptor Insulin Rat fat cells 3.30E-03 2.10E-08 151.52 (27)FPR FNLLP Rabbit neutrophils 6.70E-03 2.00E-08 74.63 (28)Total fibronectin receptors Fibronectin Fibroblasts 1.00E-02 8.60E-07 50.00 (29)T-cell receptor Class II MHC-peptide 2B4 T-cells 5.70E-02 6.00E-05 8.77 (30)FPR N-formyl peptides Human neutrophils 1.70E-01 1.20E-07 2.94 (31)cAMP receptor cAMP D. discoideum 1.00E+00 3.30E-09 0.50 (32)IL-5 receptor IL-5 COS 1.47E+00 5.00E-09 0.34 (33)NMDA receptor Glutamate Hippocampal neurons 5.00E+00 1.00E-06 0.10 (34)Adenosine A2A Adenosine HEK 293 (human) 1.75E+01 5.20E-08 0.03 (35)AMPA receptor Glutamate HEK 293 (human) 2.00E+03 5.00E-04 2.50E-04 (36)

A549, human lung alveolar carcinoma; BC3H1, smooth muscle-like cell line; COS, fibroblast-like cell line derived from monkey kidney tissue; 2.4G2 Fab, Fabportion of 2.4G2 antibody against receptor; fMLP, N-formyl-methionyl-leucyl-phenylalanine; FNLLP, N-formylnorleucylleucylphenylalanine; HepG2, humanhepatoma cell line; τ, time it takes the binding reaction to reach 63% of its final (equilibrium) value. The value of τ depends on the concentration of the ligand(Fig. 1). Thus, we show the data for a concentration of ligand equal to the Kd of each reaction. Prazosin is an antagonist to the receptor.

2 of 10 | www.pnas.org/cgi/doi/10.1073/pnas.1322761111 Ventura et al.

used low numbers of yeast in a rather large volume. To blockreceptor turnover, we performed the experiment in the presenceof the translation inhibitor cycloheximide and used a strain ex-pressing a truncated receptor that is not endocytosed (T305) (38).As predicted, the dose–response curve for receptor binding exhibi-ted a highEC50 at early times (184±60 nM) that slowly decreased toits low equilibrium value (23 ± 3 nM).

Utilization of Pre-equilibrium Information Modulates the Input DynamicRange. This slow decrease in the EC50 of the response curve forreceptor binding suggested that cells that could activate signal-ing rapidly, before ligand-receptor binding reaches equilibrium,might be able to discriminate ligand doses that would be other-wise indistinguishable. This idea has not been previously exploredand itmay potentially be very important.We refer to such signaling

0.01 0.1 1 10 100 10000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

m.

Bou

nd R

ecep

tor

(C)

0.1s

100s

10s1s

L1/Kd

L2/Kd

C1-10s

C2-10s

C2-eq

C1-eqB

0 5 10 15 200

0.2

0.4

0.6

0.8

1

Time (min)

Nor

m. B

ound

Rec

epto

r

0.1

L/Kd

1

10100

A

0

0.25

0.50

0.75

1

0.01 0.1 1 10 100Fluorescent αF/Kd

Mem

bran

e flu

ores

cenc

e(N

orm

. mea

n FL

/cel

l)

Time (min)153365180300

Time

C

Fig. 1. Time-dependent shift of the binding dose–response curve. (A) Time course of receptor/ligand complex formation for different ligand concentrations

(relative to the affinity dissociation constant, Kd), in the range 0.1–100 Kd. We computed values for the case of one-step binding using the reaction L+R →k1

←k−

C,

with the following αF binding reaction rates: k+ = 1.9 105 M−1·s−1 and k− = 0.001 s−1 (15, 16). We assumed that the concentration of free L over time isconstant. Norm., normalized. (B) Dose–response curves for the receptor/ligand reaction computed at different times, as indicated over the curves (and witha red arrow). Two concentrations, L1 and L2, result in well-separated levels of occupied receptor C1 and C2 at 10 s (0.4211 and 0.5483, respectively), but not atthe equilibrium values C1-eq and C2-eq (0.9821 and 0.9877, respectively). (C) Yeast cells of strain YAB3725 [Δbar1, PSTE2 STE2(T305)-CFP] were grown andstimulated with the indicated amounts of a fluorescent αF derivative (37), and the binding was determined by fluorescence microscopy as explained inMaterials and Methods. The mean and 95% confidence interval for the mean are shown for each concentration and time. A simple binding model wasglobally fitted to the data, resulting in Kd = 23 ± 3 nM and k− = 1.0 ± 0.2 10−4·s−1 (solid lines).

0 1 2 3 4 50

0.2

0.4

0.6

Time (min)

Dow

nstr

eam

Res

pons

e (X

*) 100

10

10.1

L+R C

X X*

X ^

A

0.01 1 100 100000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

mal

ized

Res

pons

e

slow fast

fast fast

81

114343

Ceq

0.75%

8.5%

B

1000100.1

C(t)X*(peak)C(tmax)

C

0.01 0.1 1 10 100 10000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

mal

ized

Res

pons

es

t = 3

3.5

sec

t = 4

.3 se

c

t = 1

15.4

sec

t = 1

1.3

sect =

359

.1 se

c

Ceq

Fig. 2. PRESS. Coupling a slow binding reaction to a fast and transient response can expand the input dynamic range beyond equilibrium saturation. (A) Insetshows a toy model with a downstream response activated by the ligand-receptor complex computed in Fig. 1A. Occupied receptor activates effector X; then,X* converts into X^, which slowly converts back to X, closing the cycle (details are provided in SI Appendix, section 2). The plot shows X* vs. time for all of theconcentrations of L included in Fig. 1A. (B) Shift and expansion of the input dynamic range are due to PRESS. L-receptor complex at equilibrium (Ceq, solidblack line), as well as peak X* resulting from using slow (red dotted line) or fast (blue dashed line) binding/unbinding rates of L to R, vs. input L, are shown.The resulting input dynamic ranges (the fold change required in input to elicit a change from 10 to 90% of maximum output) are indicated. The two dosesindicated, L1 = 55 Kd and L2 = 80 Kd, result in an 8.5% difference in peak X* for slow-fast coupling, and only a 0.75% difference for fast-fast coupling. (C)Graphical description of the expansion of the input dynamic range in the toy model. The plot shows L-receptor complex at equilibrium (Ceq, solid black line),or at the indicated times before equilibrium (C(t), solid gray line), as well as peak X* (solid red line), all as a function of input L. Note, as shown in A, that peakX* occurs at different times for different concentrations of input L. Therefore, the values of C at the time when X* peaks (Ctmax, solid green line) correspondto different C(t) gray curves for each dose of L (green ○). The Ctmax curve is itself shifted to higher doses than Ceq, and it has a larger input dynamic range (lesssteep) than either Ceq or any C(t) curves [EC50 = 17.7 and sensitivity (nH) = 0.9]. In B and C, the x axis corresponds to L normalized by the Kd of the ligand-receptor reaction. All data correspond to simulations, done using the following parameters for the X cycle: r1 = 0.1, r2 = 0.08, r3 = 0.001, and Xtot = 10 (Xtot

being the total amount of X). Binding/unbinding rates were k− = 0.001 1/s, k+ = 0.00019 (nM · s)−1 for slow binding (A, red line in B and C) and 100-fold thosevalues for fast binding (blue line in B) (SI Appendix, Fig. S1).

Ventura et al. PNAS Early Edition | 3 of 10

BIOPH

YSICSAND

COMPU

TATIONALBIOLO

GY

PNASPL

US

as pre-equilibrium sensing and signaling (PRESS) and to the oper-ation of signaling systems before reaching equilibriumas operation inPRESS mode. In PRESS mode, such systems could determinedownstream responses using quantities different from equilibriumoccupancy levels, such as absolute receptor occupancy at a givenmoment or the time derivative of receptor occupancy, which, atshort times, is proportional to ligand concentration. We will showbelow that operation of signaling systems in PRESS mode can havetwo consequences. Operation in PRESS mode can shift the inputdynamic range (the range of input concentrations that elicit distin-guishable outputs, usually quantified by the EC90 and EC10) toa region of higher dose concentrations. Operation in PRESS modecan also expand the input dynamic range, permitting better dis-crimination at high concentrations, although still maintaining a goodresponse at low doses. Response curves with enlarged input dynamicranges are also called “subsensitive” (39) (SI Appendix, section 3).We reasoned that if the downstream signaling is transient, as well

as fast, the system should have the extra advantage of avoiding thetransmission of occupancy levels during equilibrium, which conveyno information useful to discriminate high doses. To introducePRESS and describe its effects, we considered the behavior ofa “toy model” system (Fig. 2A). Here, occupied receptor activateseffector X, then active X (X*) converts into an inactive refractorystate (X̂ ), which slowly converts back to the inactive form (X),closing the cycle. In this toy model, when X activation (X → X*)and inactivation (X* → X̂ ) reaction rates are fast relative to thereset reaction (X̂ → X), the value of X* peaks and declines; thatis, activation of X is transient [reactions that generate transientactivation are sometimes referred to as “pulse-generators” (40)].In addition, when activation and inactivation of X are fast rela-tive to the speed of the binding reaction of L, the output of thissystem (peak X*) depends on receptor occupancy before equi-librium. With the rates we used, the overall system exhibited alarge shift in the output EC50 to high doses relative to the EC50of the receptor binding reaction at equilibrium. In this case,there was also a large expansion of the input dynamic rangerelative to binding equilibrium (65-fold shift and 4.4-fold ex-pansion, with the rates used in Fig. 2B). Note that, as expected,the shift of the EC50 and the expansion of the input dynamicrange result from an increased ability of the system to discrimi-nate high doses (Fig. 2B).In this toy model, PRESS shifts the input dynamic range be-

cause the peak of X* occurs at a time when the dose–responsebinding curve has not yet reached the receptor binding equilib-rium curve; thus, it has a higher EC50 (Fig. 1B). To understandhow it is that PRESS may also expand the input dynamic range,consider that the peak of X* occurs at different times for dif-ferent concentrations of L (Fig. 2 A and C). In particular, whenL is large, the peak of X* occurs earlier than when L is small.Now, at the same time that the X cycle is taking place, the EC50 ofthe binding dose–response curve is decreasing over time (Fig. 1and Fig. 2D, gray curves). Thus, the effective ligand-receptorcurve at the time of the peak of X* is obtained by linking thevalues of the ligand-receptor complex at the time of the peak ofX* (Ctmax). This curve across times (green line in Fig. 2C) hasa larger input dynamic range than that of C at any given time. Itis for this reason that the overall system with PRESS has anexpanded dynamic range (SI Appendix, section 3).To test whether the observed expansion and displacement of

the input dynamic range required PRESS (i.e., to test if ligandbinding needed to be slower than downstream signaling), weincreased the binding reaction rates. This change largely elimi-nated the expansion of the input dynamic range and the shift ofthe EC50 (Fig. 2B, compare red and blue curves). We also testedthe requirement of PRESS by making the downstream responseless transient. We decreased the X* inactivation rate, making thetransient X* response peak later, and we obtained a smaller shiftof the output EC50 (SI Appendix, Fig. S1A).

A number of general signaling architectures can generatetransient responses, and thus would be expected to bring aboutPRESS when coupled to a relatively slow binding receptor (SIAppendix, section 2). In SI Appendix, Fig. S1 B and C, we testedtwo additional architectures, systems with incoherent feed-forward or negative feedback control. In models based on thesearchitectures, for the rates used, PRESS resulted in a large (up totwo orders) displacement in the output EC50, which was lost whenwe increased the ligand binding rates. Thus, we conclude that pre-equilibrium signaling can operate in at least the three signalingarchitectures we presented as toy models (Fig. 2 and SI Appendix,Fig. S1). These signaling topologies are found throughout pro-karyotes and eukaryotes (41–43).We wondered if the transient response from signaling systems

that operated in PRESS mode might amplify upstream noisedue, for example, to stochastic differences in the occupation ofthe receptor by ligand. A large effect of noise on the transientresponse might reduce the precision by which systems usingPRESS can distinguish between different input concentrations.To determine the effect of upstream noise on signaling in PRESSmode, we ran stochastic simulations of the toy model in Fig. 2A,using different number of receptors (SI Appendix, Fig. S2), andthen compared the coefficient of variation (CV; the SD divided bythe mean) of peak X* and of receptor occupancy at the time ofthat peak. Surprisingly, even in the case of a very small number ofreceptors (30 per cell), which introduced a significant noise at thelevel of the occupied receptor, peak X* had a smaller CV thanreceptor occupancy (SI Appendix, Fig. S2G). This result suggestedthat in the context of PRESS, a transient response downstream ofthe receptor does not amplify noise, and therefore the benefits ofenlarged and shifted input dynamic range are not lost or otherwisemasked (SI Appendix, section 4).

PRESS Can Operate During Yeast Polarization in a Chemical Gradient.We began this investigation with a system that causes yeast po-larization toward a mating pheromone source, whose architec-ture suggested that it could function in PRESS mode. We testedthis idea by modeling using experimentally based parameters. Todo so, we modeled yeast cells as impermeable spheres exposed toa steady-state αF gradient generated by a point source (Fig. 3Aand SI Appendix, section 5.1). We considered two cells located ingradients of different strengths at the same distance from thesource, where cell 1 received an average input of αF ∼10 Kdand cell 2 received an average input of αF ∼1 Kd (Fig. 3 A and B).Cells receive more αF on the side proximal to the source (front)than on the distal side (back). The difference in αF between thefront and back is the information cells have in order to determinegradient direction (8). Cells convert this information into a dif-ference in occupied receptor via the binding reaction. The maxi-mum difference in each cell occurs between points f and b, locatedat opposite sides of each cell. We call this difference, normalizedto total receptor, Delta. In principle, larger values of Delta shouldimprove the ability of the downstream machinery to detect thegradient. At equilibrium, cell 2 was in a location (∼1 Kd) with aDelta of ∼0.23. Cell 1 was closer to the saturating region of thebinding dose–response curve (∼10 Kd); therefore, Delta wassmaller, about 0.07. If gradient orientation were dependent on thevalue of Delta at equilibrium only, cell 1 would have just aboutno information to differentiate front from back (SI Appendix, Fig.S3A and section 5). However, analysis of the αF binding dynamicsrevealed an altogether different situation. Cell 1 reached 90% ofthe value of equilibrium binding in about 3.5 min, whereas cell 2reached it in about 19 min (Fig. 3C). Notably, in both cells, Deltaovershot: It peaked (Deltamax) and then declined to the equilib-rium value (Deltaeq) (Fig. 3D). For cell 1, Deltamax occurred atabout 1.7 min and was around 4.3-fold higher than Deltaeq. Cell 2had a similar but much smaller overshoot (Deltamax occurred atabout 17 min and was only 1.1-fold greater than Deltaeq).

4 of 10 | www.pnas.org/cgi/doi/10.1073/pnas.1322761111 Ventura et al.

−1 0 10

5

10

15

20

θ (in π units)

αF

/Kd

αF

d

a

f

b

θ

A

Ext

erna

l αF

con

cent

ratio

n

αFf1

αFf2

αFb1

αFb2

B αFf1

αFf2

αFb1

αFb2

0.01 0.1 1 10 1000

0.2

0.4

0.6

0.8

αF/Kd

Nor

mal

ized

Bou

nd

Rec

epto

r at

equ

ilibr

ium

C

0 10 20 300

0.2

0.4

0.6

0.8

1

Time (min)

Nor

m. B

ound

Rec

epto

r

Deltamax1

Deltamax2

αFf1

αFf2αFb1

αFb2

Deltaeq2

Overshoot magnitude

D

Deltaeq1

0 20 40 600

0.1

0.2

0.3

Time (min)

Del

ta(C

fron

t - C

back

)

Overshoot duration

0 5 10 15 200

1

2

Time (min)

10 15 200

0.02

0.04

tmax1

tmax2

E

Tim

e de

rivat

ive

of r

ecep

tor

bind

ing

(Cn/

sec

x 10

-2)

kon*αFf1

FSte5-YFP

BF

-7 1 4 9 16 27 41 60 73 94

G

Bem1-mNG

BF

Time (min)

-17 21 24 27 31 35 38 42 45 59 69 79 89 101 131 160 191

Time (min)

Deltaeq-2

Deltaeq-11

Fig. 3. Mathematical framework to study the detection of a stationary spatial gradient. (A, Left) Cell modeled as an impermeable sphere. a, radius; b, back (θ = π); d,distance to the point source of the ligand (αF); f, front (θ = 0); θ, angle away from the line connecting the center of the sphere to the point source. (A, Right) αF profile asa function of θ for two cells located at the same d from a strong (cell 1, black) or weak (cell 2, red) source, modeledwith the following parameters: q/(4πDd)= 10Kd and1Kd, respectively, and a/d= 0.3 (for details and definitions forq andD, see SI Appendix, section 5.1). Themaximumandminimumconcentrations of αF on the cell surfaceare 16.7Kd and 6.6Kd for cell 1, and 1.7Kd and 0.7Kd for cell 2. (B) Normalized equilibriumbound receptor (C) vs. αF concentration, in units ofKd, for cell 1 (○; f1, b1) andcell 2 (□; f2, b2). Dotted lines indicate the difference in normalized bound receptor between the front and back for both cells (Deltaeq-1 and Deltaeq-2) at equilibriumbinding. (C) Normalized bound receptor vs. time at the front (solid lines) and back (dashed lines) for cell 1 (black, 10 Kd) and cell 2 (red, 1 Kd). Maximum Delta(Deltamax) is indicated by dotted lines. (D)Delta vs. time for cell 1 (black) and cell 2 (red). The amplitude and duration of the overshoot for cell 1 are indicatedwithdotted lines. (E) Time derivative (the rate of receptor occupation) vs. time of the data presented in C, for cell 1 (black) and cell 2 (red) at the front (solid line) and back(dashed line). Close to t = 0, the rate is directly proportional to the αF concentration (kon * αF, ○). Deltamax (the peak of the overshoot) occurs at tmax (x), when thecurves corresponding to the rates at the front and back cross (become equal). (Inset) Zoomed-in view of the same plot showing tmax for cell 2. (F) Fast polarization ofthe αF pathwaymachinery in living yeast. We stimulated yeast expressing theMAPK scaffold protein Ste5 fused to YFPx3 with 1 μM αF (isotropic stimulation) at time0 and then followed cells by time-lapse confocal fluorescence microscopy at the time points indicated. Images show polarization (formation of an Ste5 patch) in onecell, which is evident starting at 1min. Numbers correspond to time after αF addition (Movie S1 and SI Appendix, Fig. S4). (G) Fast relocalization of the polarization sitein response to an external cue.We stimulated yeast expressing Bem1 fused to threemNeonGreen fluorescent proteins (mNG)with a 0–50nM linear gradient (high onthe right side), such that cells experience a 1 nMdifference in pheromone concentration from “front” to “back” (SI Appendix, Fig. S5). Then,we followed cells by time-lapse confocal fluorescence microscopy at the time points indicated. Images show the location of the Bem1 patch. Light blue and yellow arrows mark the initial andnewBem1patch, respectively. Thewhite arrowmarks the daughter cell bud-neck Bem1patch. Redbarmarks spans the repositioning time. (Scale bars: F andG, 2 μm.)

Ventura et al. PNAS Early Edition | 5 of 10

BIOPH

YSICSAND

COMPU

TATIONALBIOLO

GY

PNASPL

US

Fus3 P

cytosol

Ste5

cell membrane F

Ste2 GG

Ste7Ste11

Fus3FSte5

Ste20G

Far1

Cdc42Cdc24

GG

Far1

Bem1

A

DKon

kon( ,t) k ,t)B

5

20

(in units)

F/K

d

INPUT SENSING C( ,t)

Time (min)

00.20.40.60.8

-0.5

00.5

Time (min)

(in

un

its)

OUTPUT

-0.5

00.5

(in

un

its)

Time (min)

0

20

4050

C

D

0

20

40

Binding dynamics(k -

Out

puts

fron

t qu

adra

nt (

%)

wt

0 0.50

50

200

(in units)

Num

ber

of p

olar

izat

ions

(firs

t 5 m

in)

0 0.50

50

200

(in units)

Fig. 4. Slow binding receptors efficiently convey gradient information to a cell polarization model. (A) Schema of the early events in the PRS (αF) related to gradientdetection. αF stimulates recruitment and activation of Cdc42 to sites of receptor activation. αF binds to the receptor (Ste2), causing dissociation of the G protein into Gα(Gpa1) and Gβγ (Ste4–Ste18) heterodimers. Gβγ recruits the MAPK scaffold (Ste5) to the membrane, leading to the activation of the MAPK Fus3. Gβγ also recruits Far1.Far1 recruits Cdc24, the activator of the small G protein Cdc42 (55, 56, 83). Cdc42 stimulates its own activation by binding to Bem1. Bem1 binds Cdc24, which furtheractivates Cdc42 (7, 49). αF stimulates this positive feedback loop further via recruited Ste5, which also binds Bem1 (green dotted line) (84). Active Cdc42 directs the as-sembly of actin filaments in a later phase of the gradient sensing process. Solid arrows correspond to activation, dotted arrows correspond tomoleculemovement (e.g.,membrane recruitment), dotted lines correspond to protein–protein interactions, and the double arrow indicates dissociation. (B) Schema of the Altschuler model forspontaneous emergence of cell polarity (54). It has four reactions: binding/unbinding of Cdc42 to/from the plasmamembrane, recruitment of cytoplasmic Cdc42 to siteswhere Cdc42 has already been recruited (positive feedback), and lateral diffusion of Cdc42 through themembrane. The associated parameters spontaneous associationrate (kon), random dissociation rate (koff), recruitment rate (kfb), lateral diffusion (D), and the total number of signalingmolecules (N) were estimated by Altschuler et al.(54) fromexperimental data. (C, Left) Input to thegradient sensingmodel: Steady-stateαF spatial profile, as in Fig. 3A, corresponds toa cell locatedatapoint inagradientwith an average αF concentration of 10 Kd. (C, Center) Sensing [C(θ,t)], representing normalized bound receptor. Plots correspond using a color scale on the right, atdifferent angles θ (as in Fig. 3A) vs. time, for slow [WT (wt), Left] or fast binding (fast, Right) receptors. For slow dynamics, we used published binding rates of αF to Ste2,k+=1.9×105M−1·s−1andk−=0.001 s−1 (15). For fastdynamics,weusedk+andk−thatare100-foldgreater,maintaining thesameKd. Thus, theparametersare q

4πDd=10,a/d= 0.3, and k−= 0.001 (slow) or 0.1 (fast) (SI Appendix, section 5.1). (C, Right) Example of simulation output: Plot corresponds to particle density at themembrane usinga color scale, indicatedbyabaron the right, atdifferent angles θ (as in Fig. 3A) vs. time. (D, Left) Percentageofpolarizations in the frontquadrant as a functionofbindingdynamics (ratek−at constantKd) foracell locatedas inCafter5min (black●) or15min (red■) of simulation.Of the testedk−rates, only those rates lower thanorequal to0.001 1/s resulted in polarizations in the front quadrant thatwere significantly different from random (25%, black line), P< 0.05. Arrows indicate the results shown in thehistograms.At the slowest rate tested, therewerenopolarizations in the first 5min. (D,Right)Histograms showing thenumberof stochastic simulationswithpolarizationat the indicatedangles θ, using slow (Left) or fast (Right) ligand-receptor bindingdynamics (2,000 simulation runs each). The sizeof thebins is π/6.5. Thepolarization statewas measured at t = 5 min. Coupling between occupied receptor and this model was done through parameters kon and kfb as follows: konðθ,tÞ=Aon *Cðθ,tÞ andkfbðθ,tÞ=Afb *Cðθ,tÞ, where Aon and Afb are the coupling parameters. Aon = 0.0012, Afb = 23, koff = 9 min−1, D = 1.2 μm2, and N = 103 (SI Appendix, Fig. S6).

6 of 10 | www.pnas.org/cgi/doi/10.1073/pnas.1322761111 Ventura et al.

We noted that the transient behavior of Delta (the difference inreceptor occupancy on the front and back sides) is functionallyequivalent to the transient behavior of X* in our toy model (Fig.2A). In the case of gradient sensing,Delta overshot because the timederivative of the occupied receptor at the front was initially largerthan at the back (because the concentration at the front was largerthan at the back), but it reached zero (steady state) faster at thefront than at the back (Fig. 3E). Therefore, cells have greatersensitivity for differences between the front and back during a timewindow (Fig. 3D) before binding equilibrium, and PRESS operatingduring this time window provides an opportunity for the cell toextract more information from the gradient than after equilibrium isreached. The slower the binding dynamics of the receptor, thelonger is the time window for accurate gradient detection (SI Ap-pendix, Fig. S3C and section 5), potentially providing a selectiveadvantage over evolutionary time for cells carrying mutations thatresulted in slower binding αF receptors.

Yeast Can Polarize Fast, Within the Time Window Compatible withPRESS. The spatial differences in receptor occupation result indifferential recruitment of the polarization machinery to the re-gion of higher binding, thus initiating cell polarization at that site.We again considered cell 1 in the model. For cell 1, there are ∼6min (Fig. 3D) during which it would be possible to improve po-larization in the direction of an αF gradient using PRESS. Wetested experimentally whether the downstreammachinery coupledto bound receptor operated quickly enough to polarize during thistime. To do so, we performed two experiments. In the first, wemeasured the timing of polarization in yeast stimulated with 1 μMαF (∼200 Kd). We used such a high level of αF to saturate itsreceptors within a few seconds, thus allowing us to follow the dy-namics of the polarization process itself, independent of the dy-namics of receptor binding. We applied αF isotropically. Duringisotropic stimulation, cells use the same core machinery as they doin gradients but use the internal marks otherwise utilized forbudding to choose the site to polarize (44). To assay polarization,we used cells expressing the MAPK scaffold protein Ste5 fused tothree YFPs, which localizes to the active G protein βγ dimer (Ste4/Ste18) (14, 45, 46) (Fig. 4A). Although the timing of YFP accu-mulation to a single pole was somewhat variable (partly due todifferences among cells in their position in the cell cycle), cells withpolarized Ste5 were visible within 1 min (Fig. 3F, Movie S1, and SIAppendix, Fig. S4).Weobtained the same result when following thepolarization of a second marker protein Ste20-YFP (SI Appendix,Fig. S4). These experimental results argued that the recruitment ofpolarization machinery downstream of the αF receptor is suffi-ciently rapid to use binding information within the time window ofthe overshoot of Delta, and is thus compatible with PRESS.In the second experiment, we asked if polarization was also

fast when cells do not use their internal cues, such as when theyhave to use gradient information. To answer this question, weperformed measurements in a gradient device (47), which ex-posed cells to a linear gradient from 0 to 50 nM across a chamberof 200 μm (SI Appendix, Fig. S5). Yeast could determine thegradient vector in these gradients (in the central region of thechamber, 54.8% yeast polarized in the front quadrant vs. 24.3%in the corresponding region of a control nongradient chamber;SI Appendix, Fig. S5D). We monitored early events in polariza-tion by observing the localization of the polarization machineryusing cells expressing Bem1 fused to a tandem repeat of threemNeonGreen fluorescent proteins (SI Appendix, section 7) (48).Bem1 is a scaffold protein that activates and clusters Cdc42,leading to the formation of the “polarity patch,” both during thecell cycle and during the mating response (49–51). We looked incells that had finished budding but in which mother and daughterwere still connected, in which the polarization machinery wasinitially localized to the bud neck. Two things happened to suchcells in such gradients. Some cells relocalized their polarization

machinery to one of two internal marks, which lie on each polealong the major axis of the cell (52, 53). Other cells relocalizedthe Bem1 patch to a new site that did not correspond to eithermark (SI Appendix, Fig. S5). We scored cells as having relo-calized Bem1 to either internal mark or to a new site. This lattergroup represented yeast that had repositioned their “internalcompass.” We found that the average time to relocalize the bud-neck Bem1 patch to these new sites was fast (10.8 ± 8.1 min).The fastest times were less than 5 min, in the time window duringwhich PRESS could benefit the cell. These results support theidea that yeast should be able to take advantage of PRESS whenmaking the gradient direction determination.

Slow Binding Receptors Efficiently Convey Gradient Information toa Cell Polarization Model. To investigate further if PRESS couldguide the yeast polarization machinery to orient in pheromonegradients, we fed the level of bound receptor at every positionand at every time point generated by our model to a downstreammodel of cell polarization (Fig. 4). For this model, we chose thestochastic neutral drift model developed by Altschuler et al. (54)because (i) it models the polarization of the very molecule(Cdc42) that operates during yeast mating (55, 56) (Fig. 4A); (ii)the model is modular, making it easy to couple to our input; and(iii) there are only four parameters (that have been estimatedexperimentally) describing four reactions: binding/unbinding ofCdc42 to/from the plasma membrane, recruitment of cytoplasmicCdc42 to sites where Cdc42 has already been recruited (positivefeedback), and lateral diffusion of Cdc42 on the membrane. Thus,the simplicity of this model limited the number of decisions weneeded to make about how to couple information from the re-ceptor binding model to it. We decided to use linear coupling,the simplest, and arguably the most restrictive, method. For thecoupling coefficient, we chose the smallest value that resulted inpolarizations in 100% of the simulations run with an average αFconcentration of 1 Kd (Fig. 4, legend) because this output is whatis observed experimentally.Our model predicted that for the αF gradient detection system, in

which PRESS is favored over equilibrium signaling, faster thannormal binding dynamics [in which the overshoot has a shorterduration (details are provided in SI Appendix, Fig. S3 and section 5)]should result in fewer polarizations oriented correctly. Our modelalso predicted that the increased ability to discriminate between thefront and back of the cell provided by PRESS should be mostpronounced at concentrations of αF close to the receptor saturationregion. To test this prediction, we simulated a range of αF bindingrates, from slower to faster than the published rates (15), spanningtime windows for PRESS from ∼90 min to ∼6 s. In Fig. 4C, we showthe time course of binding for the published rates, in which PRESScan operate, and for 100-fold faster rates, in which PRESS cannotoperate. We then ran thousands of stochastic simulations of a celllocated in a gradient with an average αF concentration of 10 Kd foreach binding rate (Fig. 4D and SI Appendix, Fig. S6). In Fig. 4D, weshow the polarization states arising in the first 5 min of simulation(t ≤ 5 min) for the two conditions shown in Fig. 4C. After 5 min, forthe WT receptor the advantage gained by PRESS has nearly dis-appeared (Figs. 3C and 4C). As predicted, simulations that usedslow binding kinetics resulted in significantly better-orientedpolarizations than those simulations using fast kinetics (Fig. 4D andSI Appendix, Fig. S5A). Specifically, using WT receptor dynamics,34% of the simulated polarizations localized to the front quadrantof the cell. This value is similar to what has been observed experi-mentally (11). In contrast, using fast dynamics, only 25.3% of thepolarizations localized to the front quadrant (WT vs. fast; P < 10−6),which is the value expected if polarization quadrant choice wasrandom. Rates slower than WT rates further increased the per-centage of outputs in the front quadrant (SI Appendix, Fig. S6A).These results indicate that in conditions where cells with fastreceptors are unable to use the gradient information, cells with slow

Ventura et al. PNAS Early Edition | 7 of 10

BIOPH

YSICSAND

COMPU

TATIONALBIOLO

GY

PNASPL

US

receptors can polarize significantly better than expected by chance.These results also suggest that PRESS is helpful for gradient sensingat high concentrations of αF.

DiscussionIn some cell signaling systems, including the yeast PRS, ligandsbind and unbind receptors relatively slowly. Here, we asked whatmight be gained from this slowness. We showed by simplemathematical analysis that in cells exposed to a sudden increasein ligand, the EC50 of the dose–response curve for receptor oc-cupation becomes progressively smaller as binding approachesequilibrium over time (Fig. 1 B and C). When receptor binding/unbinding is slow, the EC50 changes correspondingly slowly. Werealized that signaling system architectures that coupled the slowshift in the dose–response curve to faster events could enable thecell to react to information about the extracellular environment,which would be obscured after equilibrium was achieved. We calledthis ability to extract information from the time-evolving dose–response curve PRESS. For many system architectures, PRESS canmodulate the input dynamic range, simply by the difference in timingbetween ligand-receptor binding equilibrium and faster actingevents. When it operates, PRESS can shift the input dynamic rangeto higher doses, and it can also expand the input dynamic range,allowing cells to discriminate between high concentrations of ex-tracellular ligand that would otherwise be indistinguishable. Weshowed by simulation, and supported by experimentation, thatPRESS improves the orientation of yeast cells in response to pher-omone gradients.

Downstream Systems That Couple to Slow Receptors to Enable PRESS.We simulated the operation of PRESS in three toy models (Fig.2 and SI Appendix, Fig. S1) to illustrate the idea that PRESSworks with different downstream signaling architectures thatproduce a transient response, which are ubiquitous in eukar-yotes. The toy model in Fig. 2 was inspired by the operation offast-inactivating ligand-gated ion channels, common in the ner-vous system (57). The other two toy models corresponded toincoherent feedforward and negative feedback loops, respec-tively, which are commonly found in cellular signaling pathways(42, 43). These toy models had in common that their outputs weretransient. None of the toy models reflected the transient behaviorduring gradient sensing that enables PRESS in the mating pher-omone pathway. In this last case, the role of the transientdownstream signaling is played by the difference in ligand con-centration at the front and at the back, which causes a transientdifference in receptor occupancy in different parts of the cell,enabling gradient direction detection.

PRESS Expands and Shifts the Input Dynamic Range. The transientresponse enables the exploitation of pre-equilibrium information. Itdoes so by bringing about a shift in input dynamic range, an ex-pansion of input dynamic range, or both. The shift occurs becausethe peak in the transient response in the toy models takes place ata time when the binding dose–response curve has itself a higherEC50 than at binding equilibrium (Fig. 1B). The earlier that thesepeaks occur, the larger is the shift in the output EC50. The expan-sion happens when the peak of X* occurs at different times fordifferent stimulus levels: earlier when the stimulus is big and laterwhen it is small. This time shift for the maximum response, con-volved with the slow shift in the binding dose–response curve,expands the input dynamic range, as illustrated in Fig. 2C.

Other Ways Systems Modulate Input Dynamic Range. The enlarge-ment of the dynamic range in PRESS mode comes at a cost: Theresponse becomes subsensitive; therefore, the difference in theoutput for two different stimuli is smaller than it would have beenotherwise (39).Cells use othermechanisms to regulate input dynamicrange. For example, during its life cycle, the amoeba Dictyostelium

discoideum detects gradients of cAMP in a broad range of cAMPconcentrations. Here, binding of cAMP to its receptors is very quick(32), precluding PRESS.Dictyostelium uses a different approach thatexpands the input dynamic range: It has four homologous cAMPreceptors with different Kd values for cAMP (58). This multiple re-ceptor solution has the advantage overPRESS that it does not reducethe system’s sensitivity.Another way of increasing the input dynamic range beyond

receptor saturation is to avoid saturation directly by modulatingreceptor affinity, as in the case of the Escherichia coli chemotaxissystem, where the affinity of receptors for the ligand diminishesas cells adapt to higher concentrations of attractants (59, 60), orby fast turnover of the receptors, such as in the case of theerythropoietin receptor (61). Alternatively, a broad input dynamicrange might be achieved if the concentration of a ligand is trans-formed into the duration of the signal. This transformationmay beachieved, for example, by depleting or degrading the ligand ata constant rate, resulting in longer signaling times for higher doses(62, 63). Signal duration may, in turn, be converted back intoamplitude by the signaling pathway. In more general terms, it hasbeen shown in theory and by experimentation (64, 65) that neg-ative feedback could help to expand the input dynamic range.

Stimulus Dynamics and PRESS. In this work, for simplicity (and tomatch to earlier experiments), we focused our analysis of PRESS oncaseswhere the ligand concentrationoutside the cell (or the gradient)increases suddenly and stays stable thereafter (i.e., a step increase).However, in many physiological conditions, ligand concentrations(and gradients) might change over time in the time scale of receptoractivation. In these cases, the advantage of the PRESS mechanism(namely, the modulation of the input dynamic range) is still present.

PRESS Enables Signaling Systems to Match Different Input DynamicRanges to Different Outputs.Other outputs of the yeast PRS, suchas αF-induced gene expression, exhibit DoRA (i.e., match wellthe equilibrium binding curve of pheromone to the receptor) forseveral hours of continued stimulation (13, 14). This result sug-gests that the pheromone response pathway, using a single typeof receptor, operates simultaneously in two modes to elicit dif-ferent outputs: PRESS for gradient sensing and equilibriumsignaling for determining the level of steady-state gene expres-sion. We suggest that yeast evolved a system with slow bindingdynamics that combines PRESS with equilibrium binding, thusenabling the PRS to cover a broad input dynamic range, from thehigh concentrations sensed during orientation in gradients to thelower concentrations affecting fate choices and gene expression.We speculate that a similar combination of PRESS with equilib-rium signaling operates in other signaling systems with multipleoutputs, providing the ability to shift the input dynamic rangeaccording to physiological function.

PRESS and Other Mechanisms.We note that PRESS, which we havedescribed in cells responding to a steady spatial gradient, is notthe same as a mechanism(s) that enables sensing slowly changingspatial ligand gradients. During Drosophila melanogaster embryodevelopment, before cellularization, the dividing nuclei decodethe concentration of the morphogen Bicoid (start to expressthe Bicoid target gene Hunchback) before the Bicoid ante-roposterior gradient reaches steady state. This process in knownas “pre–steady-state decoding” (66, 67).PRESS mode may combine with other mechanisms postulated

to improve gradient sensing. Previous modeling studies sug-gested that the secreted protease Bar1 might improve matingpheromone gradient sensing. This protease degrades pheromone(68), and by doing so, it changes the shape of the gradient, po-tentially facilitating discrimination between similar partners (69,70). However, gradient detection is not significantly affected instrains deleted for bar1 (11). In any case, PRESS provides an

8 of 10 | www.pnas.org/cgi/doi/10.1073/pnas.1322761111 Ventura et al.

independent mechanism by which cells might sense gradientsmore efficiently. In addition, positive feedback loops (6, 7), to-gether with negative feedbacks (71, 72), have been shown to helpyeast establish a single polarization site. These mechanismsoperate downstream of receptor binding, and therefore down-stream of PRESS. In fact, we have used a polarization model inour work that incorporates positive feedback (Fig. 4), and we haveshown that it interacts appropriately with PRESS. Thus, PRESSdoes not rule out other mechanisms that have been proposed.

PRESS Mechanism Might Be Widespread. Systems with PRESS maybe widespread, given that many cell signaling systems have slowbinding receptors (73) (Table 1) and that there are severaldownstream network topologies that can result in transient signalresponses (74). A number of candidates are in the central ner-vous system, where fast-inactivating, ligand-gated ion channelsare commonplace (41), enabling the pairing of neurotransmitterbinding to fast and transient responses. PRESS may also operatein signaling systems at points other than ligand-receptor binding.The key necessary requirement is, as explained above, that thedose–response at a given step shifts over time. For example,cycles of activation and inactivation of substrates by phosphor-ylation and dephosphorylation are ubiquitous in signaling sys-tems (75). Such cycles are similar to binding/unbinding reactionsin the sense that the time to reach state–state concentrations ofphosphorylated substrate after a step increase in the input (inthis case, the kinase) depends on the size of the increase. If theslowly increasing concentration of activated substrate, in turn,activates a faster transient response, the overall system is capableof PRESS.

Materials and MethodsMathematical Models. The toy model (Fig. 2) and the ligand-receptor modelformulation and analysis for the case of an αF gradient (Figs. 3 and 4) aregiven in SI Appendix, sections 2 and 5.

Numerical Simulations. Stochastic simulations (Fig. 4) were performed usingthe routines kindly provided by S. Altschuler (University of California, SanFrancisco) and modified by us to receive the ligand-receptor information, asdescribed. Simulations were done using custom MATLAB (MathWorks)software. The number of simulations needed for the histograms was de-termined by increasing the sample size until obtaining no changes in the overalloutput. Polarization in simulations was determined following the same criteriaas Altschuler et al. (54); the position of polarizations was computed at the timewhenpolarization first appeared, andhistograms contain the polarizations thatappeared in the first 5 min.

Experimental Procedures. S. cerevisiae strains were of the W303a geneticbackground, derived from ACL379 (12) (MATa, Δbar1) by standard nucleicacid and yeast manipulation procedures (76). More information on strainsused and their construction is provided in SI Appendix, section 6.1.

Binding Dose–Response Curve Measurement. Yeast [strain YAB3725, Δbar1PSTE2-STE2(T305)-CFP] was grown to exponential phase in synthetic completemedium and briefly sonicated, and cycloheximide was added to a final con-centration of 100 μg/mL to block protein translation. The carboxyl-terminaldomain of Ste2 was truncated to eliminate receptor endocytosis (38). At time 0,cells were imaged using the CFP filter cube to estimate total Ste2 abundance,immediately followed by the addition of various amounts of a Hilyte488-labeled αF (37). Images were then taken approximately every 15 min for up to5 h. To calculate αF binding dynamics, images were first analyzed using Cell-IDsoftware (77), which calculates the fluorescence density at the membrane for

each cell. Then, a simple binding model (αF + R ↔ C) was globally fitted to thedata, resulting in Kd = 23 ± 3 nM and k− = 1.0 ± 0.2 10−4·s−1.

Assay of Polarization. We used S. cerevisiae W303a strains MWY003 andESY3136 [relevant genotypes: STE5-YFP(3x) and STE20-YFP, respectively],which are derivatives of ACL379 expressing the fusions under the control oftheir respective native promoters. For image cytometry, we affixed expo-nentially growing cells to the bottom of wells in a glass-bottomed 96-wellplate, as described (12), and stimulated them with 1 μM αF. We performedimage acquisition essentially as described (78) and quantified the results asdescribed in SI Appendix, section 6.2.

Chemotropism Assay in Microfluidic Devices. We fabricated microfluidicdevices designed for the generation of stable gradients in open chambersusing standard protocols for polydimethylsiloxane (PDMS)microfluidic deviceconstruction (47). Briefly, we generated silicon molds by three-layer SU-8photolithography, which were then used for making PDMS replicas of thedevice, by excluding PDMS from the tallest features of the mold, therebyproducing open (roofless) chambers. After polymerization, we peeled offthe patterned PDMS structure and bonded it onto glass cover slides byplasma-oxygen treatment (660 mtorr, 60 W, 60 s).

To improve adherence of cells to the glass, we treated the bottom of thechambers with poly-D-lysine (1 mg/mL; Sigma) at room temperature for atleast 3 h and then incubated the chambers with Con A (1 mg/mL; Sigma)overnight at 4 °C. We then filled the device with 0.22 μm of sterile, filteredwater using a vacuum-assisted method (79). Subsequently, we connected thetwo ports of the device with tubing and syringes filled with filtered syntheticcomplete medium alone or with 50 nM αF and 0.1 mg/mL bromophenol blue(BPB) as tracking dye [DBPB = 4.4 10−6·cm2·s−1 (3), DαF = 3.2 10−6·cm2·s−1 (4)].All media contained 100 ppm PEG3000 (Sigma) to prevent nonspecific αFbinding to the container’s surfaces (80). Water hydrostatic pressure (H = 5 cm)drove all flow. We evaluated the formation of the gradients by monitoringBPB fluorescence. Finally, we stopped the flow, washed the chambers withmedia, and loaded a mildly sonicated yeast exponential culture on top of thedevice. We allowed cells to settle and bind to the bottom glass before re-suming the flow. We performed imaging using an Olympus IX-81 micro-scope, with an Olympus UplanSapo objective with a magnification of 63×(N.A. = 1.35) coupled with an HQ2 (Roper Scientific) cooled CCD camera.

Quantification of Repositioning Time. We loaded a yeast strain expressingBEM1-3×–mNeonGreen (48) fusion integrated at the endogenous locus(YGV5097, derived from ACL379) in the microfluidic device. We monitoredthe Bem1 polarization patch imaging approximately every 3 min. We thenquantified polarization times as explained for Ste5, with the additionalanalysis of the angle of the different polarizations. We considered patchesas using “internal marks” (proximal or distal), or “no internal cue” based onthe angle between the position of the polarization at the neck and the newsite. We measured repositioning time as the interval between the last framewith Bem1 at the bud-neck and the first frame where it appeared asa patch elsewhere.

Statistics. P values and SEs for the mean were calculated using bootstrapmethods (81, 82).

ACKNOWLEDGMENTS. We thank C. G. Pesce, P. Aguilar, Peter Pryciak, AlbertoKornblihtt, M. González Gaitán, and Andreas Constantinou for helpful discus-sions and comments on the manuscript; Peter Pryciak (University of Massachu-setts) and Eduard Serra (Institute of Predictive and Personalized Medicine ofCancer) for providing yeast strains YPP3662 and ESY3136; S. Altschuler (Universityof California, San Francisco) for providing the simulation code that was the basisfor our simulations; and D. G. Drubin (University of California, Berkeley) forkindly providing the fluorescent αF. Work was supported by Grant PICT2010-2248 from the Argentine Agency of Research and Technology (to A.C.-L.) andGrant 1R01GM097479-01 from the National Institute of General Medical Scien-ces, National Institutes of Health (to R.B. and A.C.-L.).

1. Li R, Bowerman B (2010) Symmetry breaking in biology. Cold Spring Harb PerspectBiol 2(3):a003475.

2. Macnab RM, Koshland DE, Jr (1972) The gradient-sensing mechanism in bacterialchemotaxis. Proc Natl Acad Sci USA 69(9):2509–2512.

3. Swaney KF, Huang CH, Devreotes PN (2010) Eukaryotic chemotaxis: A network ofsignaling pathways controls motility, directional sensing, and polarity. Annu Rev Bio-phys 39:265–289.

4. Segall JE (1993) Polarization of yeast cells in spatial gradients of alpha mating factor.Proc Natl Acad Sci USA 90(18):8332–8336.

5. Zigmond SH (1989) Chemotactic response of neutrophils. Am J Respir Cell Mol Biol1(6):451–453.

6. Arkowitz RA (2009) Chemical gradients and chemotropism in yeast. Cold Spring HarbPerspect Biol 1(2):a001958.

7. Slaughter BD, Smith SE, Li R (2009) Symmetry breaking in the life cycle of the buddingyeast. Cold Spring Harb Perspect Biol 1(3):a003384.

8. Jackson CL, Hartwell LH (1990) Courtship in S. cerevisiae: Both cell types choosemating partners by responding to the strongest pheromone signal. Cell 63(5):1039–1051.

Ventura et al. PNAS Early Edition | 9 of 10

BIOPH

YSICSAND

COMPU

TATIONALBIOLO

GY

PNASPL

US

9. Jenness DD, Burkholder AC, Hartwell LH (1983) Binding of alpha-factor pheromone to yeasta cells: Chemical and genetic evidence for an alpha-factor receptor. Cell 35(2 Pt 1):521–529.

10. Ayscough KR, Drubin DG (1998) A role for the yeast actin cytoskeleton in pheromonereceptor clustering and signalling. Curr Biol 8(16):927–930.

11. Moore TI, Chou CS, Nie Q, Jeon NL, Yi TM (2008) Robust spatial sensing of matingpheromone gradients by yeast cells. PLoS ONE 3(12):e3865.

12. Colman-Lerner A, et al. (2005) Regulated cell-to-cell variation in a cell-fate decisionsystem. Nature 437(7059):699–706.

13. Yi TM, Kitano H, Simon MI (2003) A quantitative characterization of the yeast het-erotrimeric G protein cycle. Proc Natl Acad Sci USA 100(19):10764–10769.

14. Yu RC, et al. (2008) Negative feedback that improves information transmission inyeast signalling. Nature 456(7223):755–761.

15. Jenness DD, Burkholder AC, Hartwell LH (1986) Binding of alpha-factor pheromone toSaccharomyces cerevisiae a cells: Dissociation constant and number of binding sites.Mol Cell Biol 6(1):318–320.

16. Bajaj A, et al. (2004) A fluorescent alpha-factor analogue exhibits multiple steps onbinding to its G protein coupled receptor in yeast. Biochemistry 43(42):13564–13578.

17. Pruzansky JJ, Patterson R (1986) Binding constants of IgE receptors on human bloodbasophils for IgE. Immunology 58(2):257–262.

18. Mellman IS, Unkeless JC (1980) Purification of a functional mouse Fc receptor throughthe use of a monoclonal antibody. J Exp Med 152(4):1048–1069.

19. HerkenhamM, et al. (1990) Cannabinoid receptor localization in brain. Proc Natl AcadSci USA 87(5):1932–1936.

20. Wang HM, Smith KA (1987) The interleukin 2 receptor. Functional consequences of itsbimolecular structure. J Exp Med 166(4):1055–1069.

21. Hughes RJ, Boyle MR, Brown RD, Taylor P, Insel PA (1982) Characterization of coex-isting alpha 1- and beta 2-adrenergic receptors on a cloned muscle cell line, BC3H-1.Mol Pharmacol 22(2):258–266.

22. Horwitz EM, Jenkins WT, Hoosein NM, Gurd RS (1985) Kinetic identification of a two-state glucagon receptor system in isolated hepatocytes. Interconversion of homoge-neous receptors. J Biol Chem 260(16):9307–9315.

23. Marasco WA, Feltner DE, Ward PA (1985) Formyl peptide chemotaxis receptors on the ratneutrophil: Experimental evidence for negative cooperativity. J Cell Biochem 27(4):359–375.

24. Bajzer Z, Myers AC, Vuk-Pavlovi�c S (1989) Binding, internalization, and intracellularprocessing of proteins interacting with recycling receptors. A kinetic analysis. J BiolChem 264(23):13623–13631.

25. Ciechanover A, Schwartz AL, Dautry-Varsat A, Lodish HF (1983) Kinetics of in-ternalization and recycling of transferrin and the transferrin receptor in a humanhepatoma cell line. Effect of lysosomotropic agents. J Biol Chem 258(16):9681–9689.

26. Waters CM, Oberg KC, Carpenter G, Overholser KA (1990) Rate constants for binding,dissociation, and internalization of EGF: Effect of receptor occupancy and ligandconcentration. Biochemistry 29(14):3563–3569.

27. Lipkin EW, Teller DC, de Haën C (1986) Kinetics of insulin binding to rat white fat cellsat 15 degrees C. J Biol Chem 261(4):1702–1711.

28. Zigmond SH, Sullivan SJ, Lauffenburger DA (1982) Kinetic analysis of chemotacticpeptide receptor modulation. J Cell Biol 92(1):34–43.

29. Akiyama SK, Yamada KM (1985) The interaction of plasma fibronectin with fibro-blastic cells in suspension. J Biol Chem 260(7):4492–4500.

30. Matsui K, Boniface JJ, Steffner P, Reay PA, Davis MM (1994) Kinetics of T-cell receptorbinding to peptide/I-Ek complexes: Correlation of the dissociation rate with T-cellresponsiveness. Proc Natl Acad Sci USA 91(26):12862–12866.

31. Sklar LA, Sayre J, McNeil VM, Finney DA (1985) Competitive binding kinetics in ligand-receptor-competitor systems. Rate parameters for unlabeled ligands for the formylpeptide receptor. Mol Pharmacol 28(4):323–330.

32. Ueda M, Sako Y, Tanaka T, Devreotes P, Yanagida T (2001) Single-molecule analysis ofchemotactic signaling in Dictyostelium cells. Science 294(5543):864–867.

33. Morton T, Li J, Cook R, Chaiken I (1995) Mutagenesis in the C-terminal region ofhuman interleukin 5 reveals a central patch for receptor alpha chain recognition. ProcNatl Acad Sci USA 92(24):10879–10883.

34. Clements JD, Lester RA, Tong G, Jahr CE, Westbrook GL (1992) The time course ofglutamate in the synaptic cleft. Science 258(5087):1498–1501.

35. Hoffmann C, et al. (2005) A FlAsH-based FRET approach to determine G protein-coupled receptor activation in living cells. Nat Methods 2(3):171–176.

36. Krampfl K, et al. (2002) Control of kinetic properties of GluR2 flop AMPA-typechannels: Impact of R/G nuclear editing. Eur J Neurosci 15(1):51–62.

37. Toshima JY, et al. (2006) Spatial dynamics of receptor-mediated endocytic traffickingin budding yeast revealed by using fluorescent alpha-factor derivatives. Proc NatlAcad Sci USA 103(15):5793–5798.

38. Konopka JB, Jenness DD, Hartwell LH (1988) The C-terminus of the S. cerevisiae alpha-pheromone receptor mediates an adaptive response to pheromone. Cell 54(5):609–620.

39. Koshland DE, Jr, Goldbeter A, Stock JB (1982) Amplification and adaptation in reg-ulatory and sensory systems. Science 217(4556):220–225.

40. Basu S, Mehreja R, Thiberge S, Chen M-TT, Weiss R (2004) Spatiotemporal control of geneexpression with pulse-generating networks. Proc Natl Acad Sci USA 101(17):6355–6360.

41. Jones MV, Westbrook GL (1996) The impact of receptor desensitization on fast syn-aptic transmission. Trends Neurosci 19(3):96–101.

42. Tyson JJ, Novák B (2010) Functional motifs in biochemical reaction networks. AnnuRev Phys Chem 61:219–240.

43. Alon U (2006) An Introduction to Systems Biology: Design Principles of BiologicalCircuits (Chapman & Hall, London).

44. Dorer R, Pryciak PM, Hartwell LH (1995) Saccharomyces cerevisiae cells execute a defaultpathway to select amate in the absence of pheromone gradients. J Cell Biol 131(4):845–861.

45. Whiteway MS, et al. (1995) Association of the yeast pheromone response G proteinbeta gamma subunits with the MAP kinase scaffold Ste5p. Science 269(5230):1572–1575.

46. Pryciak PM, Huntress FA (1998) Membrane recruitment of the kinase cascade scaffoldprotein Ste5 by the Gbetagamma complex underlies activation of the yeast phero-mone response pathway. Genes Dev 12(17):2684–2697.

47. Keenan TM, Hsu C-H, Folch A (2006) Microfluidic “jets” for generating steady-state gradients of soluble molecules on open surfaces. Appl Phys Lett 89(11):114103-1–114103-3.

48. Shaner NC, et al. (2013) A bright monomeric green fluorescent protein derived fromBranchiostoma lanceolatum. Nat Methods 10(5):407–409.

49. Butty AC, et al. (2002) A positive feedback loop stabilizes the guanine-nucleotideexchange factor Cdc24 at sites of polarization. EMBO J 21(7):1565–1576.

50. Irazoqui JE, Gladfelter AS, Lew DJ (2003) Scaffold-mediated symmetry breaking byCdc42p. Nat Cell Biol 5(12):1062–1070.

51. Chenevert J, Corrado K, Bender A, Pringle J, Herskowitz I (1992) A yeast gene (BEM1)necessary for cell polarization whose product contains two SH3 domains. Nature356(6364):77–79.

52. Casamayor A, Snyder M (2002) Bud-site selection and cell polarity in budding yeast.Curr Opin Microbiol 5(2):179–186.

53. Zahner JE, Harkins HA, Pringle JR (1996) Genetic analysis of the bipolar pattern of budsite selection in the yeast Saccharomyces cerevisiae. Mol Cell Biol 16(4):1857–1870.

54. Altschuler SJ, Angenent SB, Wang Y, Wu LF (2008) On the spontaneous emergence ofcell polarity. Nature 454(7206):886–889.

55. Nern A, Arkowitz RA (1998) A GTP-exchange factor required for cell orientation.Nature 391(6663):195–198.

56. Nern A, Arkowitz RA (1999) A Cdc24p-Far1p-Gbetagamma protein complex requiredfor yeast orientation during mating. J Cell Biol 144(6):1187–1202.

57. Lisman JE, Raghavachari S, Tsien RW (2007) The sequence of events that underlie quantaltransmission at central glutamatergic synapses. Nat Rev Neurosci 8(8):597–609.

58. Kim JY, Borleis JA, Devreotes PN (1998) Switching of chemoattractant receptors pro-grams development and morphogenesis in Dictyostelium: Receptor subtypes activatecommon responses at different agonist concentrations. Dev Biol 197(1):117–128.

59. Sourjik V, Wingreen NS (2012) Responding to chemical gradients: Bacterial chemo-taxis. Curr Opin Cell Biol 24(2):262–268.

60. Friedlander T, Brenne N (2011) Adaptive response and enlargement of dynamicrange. Math Biosci Eng 8(2):515–528.

61. Becker V, et al. (2010) Covering a broad dynamic range: Information processing at theerythropoietin receptor. Science 328(5984):1404–1408.

62. Behar M, Hao N, Dohlman HG, Elston TC (2008) Dose-to-duration encoding and signalingbeyond saturation in intracellular signaling networks. PLOS Comput Biol 4(10):e1000197.

63. Edelstein SJ, Stefan MI, Le Novère N (2010) Ligand depletion in vivo modulates thedynamic range and cooperativity of signal transduction. PLoS ONE 5(1):e8449.

64. Nevozhay D, Adams RM, Murphy KF, Josic K, Balázsi G (2009) Negative autoregulationlinearizes the dose-response and suppresses the heterogeneity of gene expression.Proc Natl Acad Sci USA 106(13):5123–5128.

65. Madar D, Dekel E, Bren A, Alon U (2011) Negative auto-regulation increases the inputdynamic-range of the arabinose system of Escherichia coli. BMC Syst Biol 5:111.

66. Bergmann S, et al. (2007) Pre-steady-state decoding of the Bicoid morphogen gra-dient. PLoS Biol 5(2):e46.

67. Tamari Z, Barkai N (2012) Improved readout precision of the Bicoid morphogengradient by early decoding. J Biol Phys 38(2):317–329.

68. MacKay VL, et al. (1988) The Saccharomyces cerevisiae BAR1 gene encodes an ex-ported protein with homology to pepsin. Proc Natl Acad Sci USA 85(1):55–59.

69. Andrews SS, Addy NJ, Brent R, Arkin AP (2010) Detailed simulations of cell biologywith Smoldyn 2.1. PLOS Comput Biol 6(3):e1000705.

70. Rappaport N, Barkai N (2012) Disentangling signaling gradients generated byequivalent sources. J Biol Phys 38(2):267–278.

71. Kuo C-CC, et al. (2014) Inhibitory GEF phosphorylation provides negative feedback inthe yeast polarity circuit. Curr Biol 24(7):753–759.

72. Wu C-FF, Lew DJ (2013) Beyond symmetry-breaking: Competition and negativefeedback in GTPase regulation. Trends Cell Biol 23(10):476–483.

73. Lauffenburger DA, Linderman JJ (1993) Receptors: Models for Binding, Trafficking,and Signaling (Oxford Univ Press, New York).

74. Ma W, Trusina A, El-Samad H, Lim WA, Tang C (2009) Defining network topologiesthat can achieve biochemical adaptation. Cell 138(4):760–773.

75. Kholodenko BN, Hancock JF, Kolch W (2010) Signalling ballet in space and time. NatRev Mol Cell Biol 11(6):414–426.

76. Guthrie C, Fink GR (1991) Methods in Enzymology, Guide to Yeast Genetics andMolecular Biology (Academic, San Diego).

77. Gordon, et al. (2007) Single-cell quantification of molecules and rates using open-source microscope-based cytometry. Nat Methods 4(2):175–181.

78. Bush A, Colman-Lerner A (2013) Quantitative measurement of protein relocalizationin live cells. Biophys J 104(3):727–736.

79. Monahan J, Gewirth AA, Nuzzo RG (2001) A method for filling complex polymericmicrofluidic devices and arrays. Anal Chem 73(13):3193–3197.

80. Liu B, et al. (2013) Parts-per-million of polyethylene glycol as a non-interferingblocking agent for homogeneous biosensor development. Anal Chem 85(21):10045–10050.

81. Efron B, Tibshirani R (1994) An Introduction to the Bootstrap (Chapman & Hall, New York).82. Cedersund G, Roll J (2009) Systems biology: Model based evaluation and comparison

of potential explanations for given biological data. FEBS J 276(4):903–922.83. Butty AC, Pryciak PM, Huang LS, Herskowitz I, Peter M (1998) The role of Far1p in

linking the heterotrimeric G protein to polarity establishment proteins during yeastmating. Science 282(5393):1511–1516.

84. Leeuw T, et al. (1995) Pheromone response in yeast: Association of Bem1p withproteins of the MAP kinase cascade and actin. Science 270(5239):1210–1213.

10 of 10 | www.pnas.org/cgi/doi/10.1073/pnas.1322761111 Ventura et al.

Ventura  et  al,  SI  Appendix  

1  

 

Figure  Legends,  Supplemental  Computational/Experimental  Procedures/Analysis    

Utilization  of  extracellular  information  before  ligand-­‐receptor  binding  reaches  equilibrium    expands  and  shifts  the  input  dynamic  range.    

 

Alejandra  C  Ventura,  Alan  Bush,  Gustavo  Vasen,  Matías  A  Goldín,  Brianne  Burkinshaw,  Nirveek  

Bhattacharjee,  Albert  Folch,  Roger  Brent,  Ariel  Chernomoretz,  and  Alejandro  Colman-­‐Lerner  

 

Ventura  et  al,  SI  Appendix  

2  

 

 Table  of  contents.  

Supplemental  Figures  directly  called  from  the  main  text.  .........................................................................  3  Figure  S1.  Toy  models  with  transient  signaling  modules  perform  PRESS,  related  to  Figure  2.  .............  3  Figure  S2.  The  effect  of  noise  during  PRESS,  related  to  Figure  2.  .........................................................  4  Figure  S3.    Characterization  of  PRESS  applied  to  gradient  detection,  related  to  Figure  3.  ...................  5  Figure  S4.  Timing  of  yeast  polarization,  related  to  Figure  3.  .................................................................  6  Figure  S5.  Timing  of  polarization  patch  relocalization,  related  to  Figure  3.  ..........................................  7  Figure  S6.  The  αF  receptor  binding  dynamics  might  be  optimal  for  PRESS,  related  to  Figure  4.  ..........  8  

1.  Characterization  of  the  ligand-­‐receptor  time-­‐dependent  dose-­‐response  curve.  ................................  10  2.  Toy  models  with  transient  signaling  modules  perform  PRESS  .............................................................  12  

Model  1:  X  with  an  inactive  refractory  state.  ......................................................................................  12  Model  2:  X  controlled  by  an  incoherent  feed-­‐forward  loop  (IFFL);  and  Model  3:  X  controlled  by  a  negative  feed-­‐back  loop  (NFBL).  .........................................................................................................  12  

3.  Models  with  transient  signaling  may  be  subsensitive,  helping  to  expand  the  input  dynamic  range.  .  14  Sensitivity  in  Model  1:  X  with  an  inactive  refractory  state.  ................................................................  14  

Table  S1.  Model  1  sensitivity.  .........................................................................................................  15  Sensitivity  in  Model  3:  X  controlled  by  a  negative  feed-­‐back  loop  (NFBL).  .........................................  16  

Table  S2.  Model  3  sensitivity.  .........................................................................................................  16  General  conclusions  based  on  the  above  studies  and  the  results  in  the  main  text:  ...........................  16  

4.  The  effect  of  noise  during  PRESS  .........................................................................................................  19  5.  Mathematical  model  to  study  the  sensing  of  a  stationary  spatial  gradient  ........................................  21  

5.1  An  analytical  expression  for  the  αF  gradient  generated  by  a  point  source,  and  for  the  variable  Delta  (Description  1)  ...........................................................................................................................  21  5.1.1  A  simplified  formula  for  the  αF  gradient  generated  by  a  point  source,  and  the  variable  Delta  24  5.2    An  analytical  expression  for  the  variable  Delta  (Description  2)  ...................................................  26  

Table  S3.  Strains.  ............................................................................................................................  29  6  Supplemental  experimental  procedures  ..............................................................................................  29  

6.1  Strains  and  plasmids.  ....................................................................................................................  29  6.2  Quantification  of  Polarization  Times  .............................................................................................  30  

7.  Triple  monomeric  NeonGreen  DNA  sequence.  ....................................................................................  31  8.  Supplemental  References  ....................................................................................................................  32  

Ventura  et  al,  SI  Appendix  

3  

 

Supplemental  Figures  directly  called  from  the  main  text.  

Figure  S1.  Toy  models  with  transient  signaling  modules  perform  PRESS,  related  to  Figure  2.  In  the  three  models,  left  panels  contain  schemas  of  the  corresponding  model,  middle  panels  show  dose-­‐responses  and  right  panels  show   temporal   dynamics.   Parameters   were   selected   so   that   X*   has   a   transient   behavior.     Dose-­‐responses   are   included   for   the  equilibrium  value  of  C  (ligand-­‐receptor  complex)  in  solid  black  lines,  and  for  the  maximum  of  X*  in  dashed  blue  lines.  The  dose  of  the  ligand   is  normalized  by   the  Kd  of   the  binding/unbinding   reaction.  Temporal  profiles  of  C   (solid   lines)  and  of  X*   (dashed   lines)  are  included  at  ligand  concentrations  indicated  over  the  traces.    A.   X   with   an   inactive   refractory   state  (same   as   in   Figure   2).   Left.   Occupied  receptor   activates   effector   X;   after   a  while,   active   X   (X*)   converts   into   an  inactive   refractory   state   (X^),   which  slowly   converts   back   to   the   inactive  form   (X),   closing   the   cycle.   Middle.  Parameter   values:   k-­‐=0.001   1/sec,  r1=0.1,   r2=0.06-­‐0.8-­‐1.6   (increasing   from  left   to   right   in   the   middle   panel),  r3=0.001,   Xtot=10   (k+   is   not   needed  because   of   the   dimensionless   variables  chosen,   see   section   1).   EC50   for   each  response:   1   for   C,   and   for   X*   is   9.3   for  r2=0.06,   65.8   for   r2=0.8,   and   87   for  r2=1.6.    Right.  Temporal  profiles  for  the  same   parameters   as   in   the   middle  panel,  using  r2=0.8.    B.   X   controlled   by   an   incoherent   feed-­‐forward   loop   (IFFL).   Left.   Occupied  receptor  activates  both  B  and  X  into  B*  and   X*,   and   B*   increases   X*  inactivation.  Middle.   Parameter   values:  k-­‐=0.001-­‐0.1   1/sec   (from   right   to   left),  kCB=0.5,   KCB=0.01,   kFBB=10,   KFBB=10,  FB=0.5,   kCX=2,   KCX=0.1,   kBX=3,   KBX=0.1  (see  SI  Appendix,  Section  2  for  details  of  the   model,   ki   are   catalytic   rate  constants   and   Ki   are  Michaelis-­‐Menten  constants,   for   example,   kCB   and  KCB   are  the  pair   characterizing  how  C   regulates  B).   Right.   Temporal   profiles   for   the  same   parameters   as   in   the   middle  panel,  using  k-­‐=0.001  1/sec.    C.  X  controlled  by  a  negative   feed-­‐back  loop   (NFBL).   Left.   Occupied   receptor  activates   X   into   X*,   X*   activates   B   into  B*,   and   B*   increases   X*   inactivation.  Middle.   Parameter   values:   k-­‐=0.001-­‐0.1  1/sec   (from   right   to   left),   kXB=1,   KXB=0.01,   kFBB=5,   KFBB=100,   FB=0.5,   kCX=5,   KCX=0.1,   kBX=6,   KBX=0.01   (see   SI   Appendix,   Section   2   for  details  of  the  model).  Right.  Temporal  profiles  for  the  same  parameters  as  in  the  middle  panel,  using  k-­‐=0.001  1/sec.

L+R C

X X*

X^

0.01 1 100 100000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

mal

ized

Res

pons

e

A

r1

r2r3 increasing r2

Dose responses Dynamics

0 1 2 3 4 50

0.2

0.4

0.6

0.8

1

Time (min)

Nor

mal

ized

Res

pons

e

100

10

1X withrefractory state

CX*

k+

k-

0.001 0.1 10 10000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

mal

ized

Res

pons

eL+R C

X X*B B*

B

0 0.1 0.2 0.3 0.40

0.2

0.4

0.6

0.8

1

Time (min)

Nor

mal

ized

Res

pons

e

1000

100

10increasingk+ and k-

Dose responses Dynamics

X controlled byincoherentfeed-forward loop(IFFL)

CX*k+

k-

L+R C

X X*B B*

0.01 1 100 100000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

mal

ized

Res

pons

e

C

0 2 40

0.2

0.4

0.6

0.8

1

Time (sec)

Nor

mal

ized

Res

pons

e

500

1500

100increasingk+ and k-

k+

k-

Dose responses Dynamics

X controlled by ŶĞŐĂƟǀĞfeed-back loop(NFBL)

CX*

Ventura  et  al,  SI  Appendix  

4  

 

Figure  S2.  The  effect  of  noise  during  PRESS,  related  to  Figure  2.    Results   of   simulating   Model   1   (the   model  with   an   inactive   refractory   state)   for   low,  medium,   and  high  number  of   receptors   (30-­‐300-­‐3000),   using   stochastic   methods.  Simulations   were   done   in   Copasi   using   its  adaptive   tau-­‐leaping   algorithm.   Parameter  values:   k+=0.0002   ml/(pmol*sec),   k-­‐=0.001  1/s,   r1=0.01   ml/(pmol*sec),   r2=0.06   1/sec,  r3=0.001   1/sec   (with   1   pmol/ml   =   1   nM).  Initial   particle   numbers:   receptor=30-­‐300-­‐3000,  x=14536.5.  Cell  volume=  5e-­‐11  ml.  Left  panels  (A,  C  and  E)  contain  normalized  dose-­‐response  curves  and  right  panels  (B,  D  and  F)  contain   normalized   temporal   profiles,   A   and  B  correspond  to  3000,  C  and  D  to  300  and  E  and   F   to   30   receptors   per   cell,   as   indicated  over   the   figure.   Left   panels:   Equilibrium  value   of   the   ligand-­‐receptor   complex   C   (Css)  in  red  dashed  lines,  maximum  of  X*  (X*max)  in  black  dashed   lines,   and  C  at   the   time  where  X*   reaches   its   maximum   (C(tmax))   in   blue  dashed   lines.   The   dose   of   the   ligand   is  normalized   by   the   Kd   of   the  binding/unbinding  reaction.  Open  circles  and  error   bars   correspond   to   mean   value   and  standard   deviation   over   N=1001   stochastic  simulations   performed  with   the   same   set   of  parameter  values   for  each  value  of   ligand,  9  values  of  ligand  were  considered  in  the  range  0.05-­‐500   nM   (corresponding   to   0.01-­‐100  L/Kd).   Right   panels:   L/Kd=1-­‐10-­‐100   Kd   are  indicated   in   green,   red,   black,   respectively.  Monotonic   curves   correspond   to   the   ligand-­‐receptor   complex   C,   curves   with   transient  behavior   correspond   to   X*.   Both   in   left   and  right   panels   curves   were   normalized   as  follows:   Ceq   and   C(tmax)   were   divided   by   the  initial   receptor   particle   number   and   X*   was  divided   by   the   initial   X   particle   number.   G:  CV2

 (coefficient   of   variation,   the   ratio   of   the  standard   deviation   to   the   mean,   squared)  versus   normalized   ligand,   for   3000-­‐300-­‐30  receptors  per  cell  (dotted,  dashed,  solid  lines,  respectively).   CV2   for   C(tmax)   in   blue,   and   for  X*max  in  black.      

 

0.1 1 10 1000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

mal

ized

Res

pons

e Rtot=3000

C(tmax)

x*max

Css

A

0.1 1 10 1000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

mal

ized

Res

pons

e Rtot=30

C(tmax)

x*max

Css

E

0.1 1 10 1000

0.2

0.4

0.6

0.8

1

L/Kd

Nor

mal

ized

Res

pons

e Rtot=300

C(tmax)

x*max

Css

C

C(tmax)

Rtot= 30Rtot= 300 X*

max

Rtot= 3000

Rtot= 30Rtot= 300Rtot= 3000

G

0.01 0.1 1 10 1000.00001

0.001

0.1

10

L/Kd

Coe

ffici

ent o

f Var

iatio

n 2

0 2 4 6 80

0.2

0.4

0.6

0.8

1

Time (min)

Nor

mal

ized

Res

pons

e

B

0 2 4 6 80

0.2

0.4

0.6

0.8

1

Time (min)

Nor

mal

ized

Res

pons

e

D

0 2 4 6 80

0.2

0.4

0.6

0.8

1

Time (min)

Nor

mal

ized

Res

pons

e

F

Ventura  et  al,  SI  Appendix  

5  

 

min

5

10

15

0

0.05

0.10

0.15

0

L1L2

k- = 0.1 s -1

k = 0.001 s -1

-5 10 15 20Back (FF/Kd )Temporal window

5 10 15 20

5101520 0.8

0.6

0.4

0.2

0Fron

t (FF

/Kd

)

A

Back (FF/Kd )

B0.80.6

0.4

0.2

L1L2

Back (FF/Kd )5 10 15 20

00 5 10 15 20

0

0.2

0.4

0.6

Time (min)

Del

ta

L1L2

DC

Delta eq Deltamax  

Figure  S3.    Characterization  of  PRESS  applied  to  gradient  detection,  related  to  Figure  3.  

A-­‐C.  Graphs  are  heatmaps  of  Deltaeq  (A),  Deltamax  (B)  and  temporal  window  (C)  as  functions  of    and    (αFf  and  αFb  divided  by  the  Kd).  Since  we  are  considering  gradients,  αFf  has  to  be  greater  than  αFb  (and  this  is  why  the  figure  is  blank  below  the  diagonal).  Deltaeq  and  Deltamax  were  calculated  using  the  analytical  calculation  in  the  SI  Appendix,  and  are  plotted  using  the  same  color  scale.  The  temporal  window  was  numerically  calculated  as  the   interval  during  which  Delta   is  more  than  1.2  Deltaeq.  We  show  two  color  scales,  obtained  with  the   indicated  unbinding  rates  k-­‐.  The  black  solid   line  over  each  panels  corresponds  to  Deltaeq  =  0.23  and  the  black  dashed  line  corresponds  to  Deltamax  =  0.23.  The  solid  grey  line  represents  a  gradient  with  an  average  between  front  and  back  of  10  Kd  and  the  dashed  grey  line,  5  Kd.  A  cell  sensing  with  an  average  10  Kd  could  have  different  combinations  of   and   ,  some  of  them  will  result  in  Deltamax  >  0.23  and  Deltaeq  <  0.23  (those  that  correspond  to  the  grey  solid  line  in  the  region  between  the  solid  and  dashed  black  lines).  The  intersections  of  the  line  representing  an  average  of  10  Kd  and  the  lines  for  Deltaeq  =  0.23  and  Deltamax  =  0.23  are   identified   as   L1   and   L2,   respectively.   L1   is   characterized   by    and   ,   resulting   in   Deltaeq=0.23   and  Deltamax=0.59,  respectively,  and  a  temporal  window  of  12.9  min.  L2  is  characterized  by  by    and   ,  respectively,  resulting  in  Deltaeq  =  0.06  and  Deltamax  =  0.23  and  a  temporal  window  of  9.1  min.  D.  Delta(t)  vs  time,  for  L1  and  L2.  The  vertical  line  indicates  the  time  were  the  temporal  window  for  PRESS  ends.        

Ventura  et  al,  SI  Appendix  

6  

 

 

0.0

0.2

0.4

0.6

0.8

1.0

0 50 100 150Time (min)

Nor

mal

ized

Cum

ulat

ive

Pola

rizat

ions

Ste20-YFPSte5-(YFP)3x

A C

B

Figure S4, related to Figure 3

 

Figure  S4.  Timing  of  yeast  polarization,  related  to  Figure  3.  A.  Example  of  a  transmission  image  montage  used  to  determine  the  position  in  the  cell  cycle  (YPP3662  strain,  1μM  α-­‐factor).  G1  cells  (top)   respond   to   pheromone   forming   a   mating   projection.   S   cells   (bottom)   first   bud   and   then   respond   to   pheromone.   Time   in  minutes  is  shown  below  each  image,  and  the  ID  of  each  cell  is  shown  on  the  left.    B.  Example  of  a  YFP  image  montage  used  to  determine  the  polarization  time  (YPP3662  strain,  1μM  α-­‐factor).  The  slice  number  of  the  Z  stack  is  indicated  on  the  left.    Time  in  minutes  is  shown  below  each  image.    C.  Cumulative  fraction  of  polarized  cells  vs.  time  for  Ste5-­‐YFPx3  (YPP3662)  or  Ste20-­‐YFP  (ESY3136).  Cells  were  stimulated  with  1μM  α-­‐factor   at   time   zero.  Only   cells   in  G1  were   considered   for   this   analysis.   Error  bars   represent   the  95%  confidence   interval   of   the  mean  as  calculated  by  bootstrap.      

Ventura  et  al,  SI  Appendix  

7  

 

Figure  S5.  Timing  of  polarization  patch  relocalization,  related  to  Figure  3.  

A.  The  microfluidic  device  we  used.  It  contains  two  ports,  one  filled  with  a  red  dye  (labeled  input  source  1)  and  the  other  one  with  a  blue  dye  (labeled  input  source  2),  and  4  open  chambers  (200  x  500  µm).  Two  of  the   chambers   receive   different   inputs,  and  a  gradient  is  formed;  the  other  two  serve   as   controls.The   fluid   is   delivered  to   the   wells   by   a   binary   splitting  manifold   that   ends   in   a   line   of   38  microjets  of  10  µm  wide  and  2.5  µm  tall.    B.   The   dynamics   of   gradient   formation.  We   formed   a   gradient   using   a   solution  of   BPB   (bromophenol   blue,   0.01%)   as  source   1   and   water   as   source   2.   We  determined   the   gradient   profile  measuring   BPB   fluorescence   using   a  standard   Texas   Red   filter   cube.   At   t=0,  we   washed   away   the   chamber   with   a  pipet   and   then   we   monitored   gradient  formation  by  imaging  every  15  seconds.  The  gradient  is  linear  and  it  reaches  the  steady  state  in  ~3  min.    C.   Brightfield   image   of   yeast   strain  ACL379   exposed   to   a   gradient   of   0-­‐50nM   of   αF   for   3   hours.   The   middle  region   of   the   chamber   displays   the  highest   percentage   of   cells   oriented   in  the  direction  of  the  gradient.    D.  Histogram  showing  the  quantification  of   the   angles   between   the   shmoo   and  the   direction   of   the   gradient.   Perfectly  aligned  cells  display  an  angle  of  0°.  The  isotropic   chamber   (right)  has  no  bias   in  the   angle,   as   expected   for   random  polarization  (uniform  external  cue).    E.   Diagram   representing   the  measurement   of   patch   repositioning  time.   Newly   divided   cells   show   Bem1  polarization   at   the   bud-­‐neck.   In  response   to   α   factor   exposure,   Bem1  forms   a   new   patch   somewhere   in   the  cell   periphery.   Depending   on   the  position   of   the   new   polarization,  patches   (both   in  mothers  or  daughters)  were  classified  as  “proximal”,  “distal”  or  “no  internal  cue”.  The  time  between  the  last  frame  with  bud-­‐neck  polarization  and  the  first  with   a   new   visible   patch   was   used   to   estimate   the   repositioning   time.    We   show   an   example   of   each   group   of   yeast   YGV5097  (expressing  Bem1-­‐mNG(3x))   (blue  arrowheads   indicate   the  polarization   in   the  neck,   yellow  arrowheads   show  a  new  patch  and   the  repositioning  time  is  delineated  by  the  red  line).  Average  repositioning  time  is  presented  in  the  table  (mean  ±  standard  deviation).  

Ventura  et  al,  SI  Appendix  

8  

 

 

Figure  S6.  The  αF  receptor  binding  dynamics  might  be  optimal  for  PRESS,  related  to  Figure  4.  A.   Percentage   of   polarizations   in   the   front   quadrant   as   a   function   of  

binding   dynamics   (parameter   k-­‐),   using  

q4πDdKd

= 10  and  a

d= 0.3    

(corresponding   to  p1=10  and  p2=0.3,   see  SI  Appendix,   Section  5),   as   in  Fig.   4D.   For   each   value   of   k-­‐,   we   run   a   set   of   stochastic   numerical  simulations   (Altschuler’s   model   with   external   gradient   as   input),   and  counted  the  number  of  runs  that  resulted  in  a  polarized  state,  and  then  determined  how  many  polarized   in   the  quadrant   facing   the  αF   source  (front  quadrant).  We  computed  polarizations  that  took  place  in  the  first  five  minutes  of  the  simulation  as  in  Fig  4D  (black  circles)  or  in  the  total  simulation  time  of  fifteen  minutes  (red  circles).  We  varied  parameter  k-­‐  from   0.0001   1/sec   to   0.1   1/sec   (this   last   value   corresponding   to   the  “fast”  receptor   in  Fig.  4D).  A  horizontal   line  over  the  plot   indicates  the  expected  percentage  output  if  polarizations  were  randomly  located,  i.e.  25%.   For   fast   receptor   dynamics,   the   system   is   unable   to   use   the  gradient   information.     A   receptor   slower   than  wt   results   in   improved  sensing,  but  only  allowing  more  time  to  polarize.  When  evaluating  the  polarization  performance  in  5  min  (black  circles)  we  found  a  maximum  in   the  percentage  outputs   in   the   front  quadrant   at   k-­‐   of   about  0.0003  1/sec.  For  the  time  window  of  15  min  instead,  the  maximum  is  achieved  at  the  slower  dynamics  considered  for  this  figure  (k-­‐=0.0001  1/sec).    B.   Percentage   of   polarizing   cells   as   a   function   of   parameter  q4πDdKd

= 10  and  a

d= 0.3  (corresponding   to  p1=10  and  p2=0.3,   see  SI  

Appendix,   Section   5).   As   explained   in   (A),   if   the   receptor   dynamics   is  very  slow,  very  few  cells  polarize  in  5  min.    Combining   the   information   in   (A)   and   (B),  we  conclude   that   there   is   a  tradeoff   between   achieving   a   good   number   of   polarizing   cells   in   a  population   and   obtaining   a   good   performance   for   those   polarizations  (in   the   sense   that   those   cells   correctly   identify   the   position   of   the  partner).  If  the  decision  has  to  be  made  in  the  first  5  minutes,  then  the  dynamics   corresponding   to   the   a-­‐factor   receptor   (k-­‐=0.001   1/sec)  results   in   the   best   combination   of   polarizing   cells   with   successful  polarizations.   For   a   longer   time   interval,   like   the   one   of   15   minutes  analyzed,  the  optimal  occurs  for  a  still  slower  dynamics,  k-­‐  about  0.0003  1/sec.    C.   Summation  of   all   the  polarizations   that   appeared   for   a   time   less  or  equal  than  the  time  indicated  in  the  horizontal  axis,  normalized  by  the  total   number   of   polarizations   achieved   at   the   end   of   the   simulation  (t=15  min).  Each  value  of  k-­‐  is  indicated  with  a  different  color  as  shown  in  the  figure.    The   number   of   stochastic   numerical   simulations   performed   for   each  value  of  k-­‐  is:  19,000  for  k-­‐=0.0001  1/sec,  10280  for  k-­‐=0.0003  1/sec,  and  

0.0001 0.001 0.01 0.10

10

20

30

40

Outp

uts

fro

nt quadra

nt (%

)

Binding dynamics

(k-, unbinding rate, 1/s)

15 minutes5 minutes

0.0001 0.001 0.01 0.10

20

40

60

80

100

Outp

uts

with p

ola

rization (

%)

Binding dynamics

(k-, unbinding rate, 1/s)

15 minutes5 minutes

0 5 10 150

0.2

0.4

0.6

0.8

1

Time (min)

Cu

mu

lative

/To

tal

Po

lariza

tio

ns

0.0001

0.0003

0.001

0.004

0.01

0.1

A

B

C

Figure S6, related to Figure 4

Ventura  et  al,  SI  Appendix  

9  

 

2000  for  all  the  other  values  of  k-­‐.  Comparing  the  outputs  for  each  two  successive  values  of  k-­‐  indicate  that  only  the  following  pairs  are  significantly  different  (p  <  0.05):  0.0003-­‐0.001,  and  0.001-­‐0.004  at  5  min;  0.0001-­‐0.0003,  and  0.001-­‐0.004    at  15  min.      

Ventura  et  al,  SI  Appendix  

10  

 

1.  Characterization  of  the  ligand-­‐receptor  time-­‐dependent  dose-­‐response  curve.    

In   this   section   we   characterize   the   sensitivity   and   potency   (given   by   parameters   nH   and   EC50,  respectively)  of   the  curve  that  describes   the  time  evolution  of  occupied  receptor  at  different  doses  of  ligand  for  the  simple  case  of  one-­‐step  binding,  assuming  that  free  L   is  not  significantly  affected  by  the  reaction:  

   with  

     

the  steady-­‐state  value  for  C  and  

=  

the  characteristic  time  τ  (time  required  to  reach  63.2%  of  the  steady-­‐state  value).  L  is  the  ligand,  R  is  the  receptor,   C   is   the   complex   ligand-­‐receptor,   and   k+   and   k-­‐   are   the   binding   and   unbinding   rates,  respectively.  Normalizing  the  curve  C(t,L)  by  Rtot  and  defining  x=L/Kd  (normalized  ligand  concentration),  results  in  a  function  that  depends  only  on  x  and  k-­‐:  

 The  steady-­‐state  of  this  last  function,  which  we  obtained  by  taking  the  limit  t-­‐>∞,  has  an  nH  =  1  and  EC50  

=  1.  We  are  interested  in  obtaining  the  EC50  and  nH  for  normalized  C  versus  x,  for  any  given  time,  which  will   result   in   two   relationships:   EC50(t)   and   nH(t).   Instead   of   EC50(t)  we   could   easily   obtain   the   inverse  function,  t(EC50),  as  follows:  

 Based  on  this  last  expression,  we  plot  EC50  versus  time  in  Fig.  S7.  According  to  this  plot,  

1. EC50  >  1  for  every  time,  and  EC50  =  1  only  for  the  equilibrium  binding  curve;  

2. EC50  decreases  when  time  increases;  3. If   k-­‐   is   large   (fast   unbinding   reaction),   the  

time  at  which  the  curve  C  versus  x  has  a  given  EC50  is  small.    For  example,  if  the  desired  EC50  is   10*Kd,   then,   the   dose   response   curve   has  that  EC50  at  around  ~1  min   for  k-­‐=0.001  sec-­‐1  and  at  ~0.1  min  for  k-­‐=0.01  sec-­‐1,  see  Fig.  S7.      

Regarding   the   sensitivity   of   the   normalized   C  versus   x   curve   for   times   prior   to   reaching   the  equilibrium,   we   obtained   an   approximated  nH=1.4   in   the   limit   of   low   values   of   t*k-­‐.   We  

 

Figure  S7.  EC50  of  L-­‐receptor  binding  curve  vs.  Time.  EC50  is  in  units  of  L/Kd.    

Ventura  et  al,  SI  Appendix  

11  

 

obtained   this   value  by  deriving   t(EC90)   and   t(EC10)   in   the   same  way  we  derived   t(EC50),   and  neglecting  terms  as  follows:    

 

   These  approximations  are  valid  because  at   short   times   (t*k-­‐<<1),  both  EC90  and  EC10  are  much  greater  than  one,  similar  to  the  behavior  of  EC50  (Figure  S7).  From  the  last  two  formulas  we  obtained  the  ratio  EC90/EC10:  

 

which   results   in   21.85   when   t*k-­‐   is   neglected,   leading   to   nH=1.42   (nH=log(81)/log(EC90/EC10).     This  approximated  result  indicates  that  the  sensitivity  goes  from  about  1.4,  in  the  limit  of  low  values  of  t*k-­‐,  to  1,  in  the  limit  of  t-­‐>∞,  and  means  that  before  reaching  equilibrium  the  sensitivity  is  higher  than  when  it  is  reached.        

Ventura  et  al,  SI  Appendix  

12  

 

2.  Toy  models  with  transient  signaling  modules  perform  PRESS    

Model  1:  X  with  an  inactive  refractory  state.  

Occupied   receptor  activates  effector  X  with  a   rate   r1;  active  X   (X*)  converts   into   an   inactive   refractory   state   (X^)   with   rate   r2,   which  converts  back  to  the  inactive  form  (X)  with  rate  r3,  closing  the  cycle  (see  Fig.  S8).      Assuming  that  the  total  amount  of  X,  Xtot  =  X  +  X*  +  X^  is   conserved,   the   system   is   described   by   the   following   ordinary  differential  equations:    

  (1)  

      (2)  

 where  C  is  the  ligand-­‐receptor  complex.  The  parameters  used  in  the  numerical  simulations  reported  in  Figs.  2a-­‐b  in  the  main  text  are  r1  =  0.1,  r2  =  0.8,  r3  =  0.001,  Xtot  =  10,  and  the  initial  condition  was  such  that  all   the  effector  was   initially   in   its   inactive  state  X.  The  steady-­‐state  reached  by  this  systems  is  characterized  by  X*eq/Xtot  =  r3/(r3  +  r2  +  r2r3/(r1Ceq))  and  Xeq/Xtot  =  r2r3/(r2r3  +  r1Ceq(r2  +  r3)).    

Model  2:  X  controlled  by  an  incoherent  feed-­‐forward  loop  (IFFL);  and  Model  3:  X  controlled  by  a  

negative  feed-­‐back  loop  (NFBL).  

The  equations  for  the  network  depicted  in  Fig.  S9  and  Fig.  S10  were  adapted  from  those  in  Ma  et  al  (1)  for   the   case   where   one   node   is   a   ligand-­‐receptor   complex.   We  assume   that   each   node   (C,   B,   X)   has   a   fixed   concentration  (normalized  to  1)  but  has  two  forms:  C  is  the  ligand-­‐receptor  complex  and  (1-­‐C)  is  the  free  receptor;  B*  and  X*  represent  the  concentration  of   active   states,   (1-­‐B*)   and   (1-­‐X*)   are   the   concentration   of   the  inactive  states.  L  is  the  concentration  of  free  ligand  and  k+  and  k-­‐  are  the   binding   and   unbinding   rates,   respectively.   kCB   and   kCX   are   the  strength  with  which  C  activates  B  and  X  into  B*  and  X*,  and  kBX  is  the  strength  of  the  negative  regulation  from  B*  to  X*.  If  a  node  has  only  positive   incoming   links,   like   node   B   in   Fig.   S9   and   in   Fig.   S10,   it   is  assumed   that   there   is   a   background   (constitutive)   deactivating  enzyme   Fi   of   a   constant   concentration   to   catalyze   the   reverse  reaction   (see   Ma   et   al   (1)   for   details).   The   resulting   equations  

 

Figure   S8.   Toy   model   with   a  refractory  state.  Toy   model   of   a   downstream  response   activated   by   the   ligand-­‐receptor   complex.   Occupied  receptor  activates  effector  X;  after  a  while,  active  X  (X*)  converts  into  an  inactive  refractory  state  (X^),  which  slowly  converts  back  to  the  inactive  form  (X),  closing  the  cycle.  

 

Figure  S9.  Toy  model  of  an  incoherent  feedforward  loop.  Occupied   receptor   activates  both   B   and   X,   into   B*   and   X*,  respectively.   B*   regulates   the  inactivation  of  X*  into  X.  

Ventura  et  al,  SI  Appendix  

13  

 

describing  the  network  in  Fig.  S9  are:  

   

 

   

The  resulting  equations  describing  the  network  in  Fig.  S10  are:  

 

   

 

               

 

Figure  S10.  Toy  model  of  a  negative  feedback  loop.  Occupied   receptor   activates   X   into   X*,   X*  activates   B   into   B*   and   B*   regulates   the  inactivation  of  X*  into  X.  

 

Ventura  et  al,  SI  Appendix  

14  

 

3.  Models  with  transient  signaling  may  be  subsensitive,  helping  to  expand  the  input  dynamic  range.  

Modulation   of   the   input   dynamic   range   could   imply   its   expansion/compression   (i.e.,   an   increase   or  decrease   of   the   range   of   doses   where   the   system   produces   dose-­‐dependent   responses),   its  displacement   (i.e.,   the   ratio   EC90/EC10   is   the   same   but   it   is   centered   around   a   different   dose),   or   a  combination  of  both.  Expansion/compression  of   the  dynamic  range   is  determined  by  the  sensitivity  of  the  response,  quantified  by  the  change  in  the  Hill  coefficient  nH  (log(81)/log(EC90/EC10))  of  the  response:  the  higher   the  nH,   the   smaller   the   input  dynamic   range.  Displacement  of   the  dynamic   range   is  usually  quantified  by  the  change  in  the  EC50  of  the  response.  A  rightward  displacement  of  the  EC50  with  constant  nH  enables  dose-­‐dependent  responses  in  the  high  concentration  range  without  compromising  sensitivity,  while  a  decrease  in  nH  with  constant  EC50  results  in  an  extended  range  of  dose-­‐dependent  responses  and  a  corresponding  loss  of  sensitivity,  meaning  that  the  difference  in  the  output  you  get  for  two  different  stimuli  is  smaller  than  otherwise  would  have  been.  

The  common  Michaelis-­‐Menten   relationship  gives,  by  definition,  an  nH=1.  Subsensitive   responses   (i.e.,  nH  <  1)  may  arise,  for  example,  from  the  presence  of  negative  cooperativity  (2)  or  negative  feedback  (3).  Systems  with  transient  responses,  such  as  the  one  we  implemented  in  our  three  toy  models,  also  known  as  pulse-­‐generators,  can  be  subsensitive.  Given  that  one  of  the  effects  of  the  PRESS  signaling  mode  is  to  make   the   overall   response   subsensitive   (expand   the   input   dynamic   range),   we   asked   if   the   transient  responses  we   implemented  are   intrinsically   subsensitive   (i.e.,  when  studied  uncoupled   from  a  binding  mechanism);  and  if  so,  how  this  subsensitivity  contributes  to  the  overall  system  behavior,  when  coupled  to  a  slow  or  a   fast  binding  reaction.  With  this   in  mind,   in  what   follows  we  analyze   in  detail   the  model  with  an  inactive  refractory  state  and  the  model  controlled  by  a  negative  feedback  loop  (Models  1  and  3).  

Sensitivity  in  Model  1:  X  with  an  inactive  refractory  state.    

In   order   to   study   the   pulse   generator   of   Model   1   in   isolation,   that   is,   independently   of   the   ligand-­‐receptor  complex  (C)  formation,  we  first  studied  the  peak  response  of  X*  as  a  function  of  a  step  increase  in   C   (Fig.   S11a).   For   the   parameter   values   we   selected,   the   system   was   subsensitive,   with   a   Hill  coefficient  of  0.8,  corresponding  to  a  dynamic  range  (EC90/EC10)  of  243  (colored  lines  in  Fig.  S11a),  three  times  the  dynamic  range  of  a  Michaelian  response  (81,  black  line).    We  tested  three  values  of  r2  (the  rate  of   inactivation  of  X*,  the  key  parameter  controlling  the  duration  of  the  pulse  response),  which  had  no  significant  impact  on  the  nH,  but  had  a  strong  effect  on  the  EC50  of  the  response.    

Next,  we  introduced  a  simple  ligand-­‐receptor  binding  reaction  between  the  step  input  and  the  X  cycle,  using   fast  or   slow   ligand-­‐receptor  binding   rates   (Fig.  S11b  and  c).    The  addition  of   the   ligand-­‐receptor  binding   reaction   increased   the   sensitivity   from   0.8   to   0.9   when   we   used   fast   ligand   binding   rates  (reducing   the   dynamic   range   to   ~132),   and   decreased   it   to   0.7   when   we   used   slow   binding   rates  (increasing  the  dynamic  range  to  ~533)  (Table  S1).  In  both  cases,  parameter  r2  had  no  significant  impact  on  the  nH,  impacting  strongly  on  the  EC50  of  the  response.    

These   results   indicate   that   the   coupling   of   a   ligand-­‐receptor   reaction   interacts   with   the   subsensitive  pulse-­‐generator,   in   a   manner   that   depends   on   the   speed   of   this   reaction.  When   binding   is   fast,   the  

Ventura  et  al,  SI  Appendix  

15  

 

overall   sensitivity   is   increased,   but   when   binding   is   slow,   it   is   further   decreased.   The   speed   of   the  binding  reaction  also  affects  the  degree  of  the  dynamic  range  displacement;   it  was  much  greater  with  slow  binding  than  fast  binding.  Thus,  we  conclude  that  the  intrinsic  subsensitivity  of  the  pulse  generator  does  not  explain  the  effects  observed  in  PRESS  mode  of  signaling.      

    C   as   the  

stimulus       Fast  

binding     Slow  

binding    

r2     EC50     nH     EC50     nH     EC50     nH    0.06        1.2   0.79     0.4     0.92      9.3     0.71    0.8     15.9   0.80   1.3     0.94     65.1     0.75    1.6     31.3   0.81     1.7   0.92     87.2     0.73  

Table  S1.  Model  1  sensitivity.  Characterization  of  the  potency  (EC50)  and  sensitivity  (nH)  for  the  different  curves  in  Fig.  S11,  as  a  function  of  

parameter  r2,  corresponding  to  Model  1.            

 

Figure  S11.  Sensitivity  in  Model  1  X  with  an  inactive  refractory  state.  Parameters:  Xtot=10,  r1=0.1,  r2=0.06-­‐0.8-­‐1.6  (red,  blue,  green  lines),  r3=0.001.  a.  The  isolated  cycle  of  X  (no  dynamics  of  ligand-­‐receptor  complex,  C,  formation).  Plot  shows  peak  response  of  X*  as  a  function  of  step  increase  in  C.  A  Michaelian  response  (with  EC50=1  and  nH=1)  is  included  with  a  black  line.  b-­‐c.  Complete  Model  1  (including  the  dynamics  of  ligand-­‐receptor  complex,  C,  formation).  Plots  show  peak  response  of   X*   as   a   function   of   the   ligand   concentration,   L   (L   normalized   by   the   Kd   of   the   ligand-­‐receptor   reaction).  Different  values  of  parameter  r2  are  included  in  different  colors,  as  indicated.  Rates  in  b  are  k+=0.02  (nM*sec)  -­‐1,  k-­‐=0.1  sec-­‐1  and  in  c  k+=0.0002  (nM*sec)  -­‐1,  k-­‐=0.001  sec

-­‐1.  

Ventura  et  al,  SI  Appendix  

16  

 

Sensitivity  in  Model  3:  X  controlled  by  a  negative  feed-­‐back  loop  (NFBL).  

We  performed  the  same  study  for  Model  3.  When  considered  in  isolation,  for  the  parameter  values  we  selected,  the  peak  response  of  X*  to  a  step  increase  in  C  (Fig.  S12)  resulted  ultrasensitive,  with  an  nH  of  

1.7,  corresponding   to  a  dynamic   range  of  13.3,  and  an  EC50  of   0.3.  When  we   included   the   dynamics   of   ligand-­‐receptor  formation  (C),  the  sensitivity  of  the  overall  system  decreased  to   1.3   for   fast   binding,   and   even   further   to   1.2   for   slow  binding  (Table  S2),  corresponding  to  a  dynamic  range  of  29.4  and   38.9,   respectively.   In   addition,   the   dynamic   range   of   C  (EC50)  was  displaced  from  1  to  4.5  for  fast  binding  and  from  1  to  447.6  with  slow  binding.    

 

C  as   the  stimulus    

  Fast  binding  

  Slow  binding  

 

EC50     nH     EC50     nH     EC50     nH    0.3   1.71     4.5     1.31     447.6   1.24    

Table  S2.  Model  3  sensitivity.  Characterization  of  the  displacement  (EC50)  and  dynamic  range  (nH)  for  the  different  curves  in  Fig.  S1B  and  in  Fig.  S12,  corresponding  

to  Model  3.  

 

General  conclusions  based  on  the  above  studies  and  the  results  in  the  main  text:  

i)  The  ligand-­‐receptor  binding  reaction  has  an  equilibrium  dose-­‐response  curve  characterized  by  nH  =  1  and   EC50   =   1   (dose   =   ligand/Kd,   response   =   normalized   ligand-­‐receptor   complex).   The   dose-­‐response  curve   before   reaching   equilibrium   has   nH   >   1   and   EC50   >   1   (see   SI   Appendix,   Section   1),   that   is,   it   is  displaced  to  the  right  and  has  smaller  input  dynamic  range  (increased  sensitivity).  Therefore,  the  shape  of  the  binding  dose  response  curve  by  itself  cannot  explain  the  overall  characteristics  of  PRESS  systems.  ii)  Pulse  generators  may  by  subsensitive  (nH  <  1),  such  as  in  Model  1  or  ultrasensitive  (nH  >  1),  such  as  in  Model  3,  depending  on  architecture  and  parameter  values.  iii)   The   coupled   system   (ligand-­‐receptor   -­‐>   pulse   generator)   results   in   a   dose-­‐response   curve   (dose   =  ligand/Kd,   response   =   maximum   value   reached   by   the   pulse)   whose   characteristics   depend   on   the  coupling,  and  particularly  on  the  dynamics  of  the  ligand-­‐receptor  reaction.    

If  the  ligand-­‐receptor  reaction  is  fast  (compared  to  the  time-­‐scales  of  the  pulse  generator),  then  we  may  consider  the  equilibrium  ligand-­‐receptor  dose-­‐response  curve  (with  nH,A  fast  =  1)  as  the  input  to  the  pulse  generator.   Therefore,   the   upper   bound   of   the   overall   system   sensitivity   depends   exclusively   on   the  sensitivity  of  the  pulse  generator.   If  the  ligand-­‐receptor  reaction  is  slow,  one  may  be  tempted  to  think  

 

 

Figure  S12.  Sensitivity  in  Model  3  X  controlled  by  a  negative  feed-­‐back  loop  (NFBL).  Parameter  values:  kXB=1,  KXB=0.01,  kFBB=5,   KFBB=100,   FB=0.5,   kCX=5,   KCX=0.1,  kBX=6,   KBX=0.01.   A   Michaelian   response  (with  EC50=1  and  nH=1)   is   included  with  a  black   line.  The  response  of   the  NFBL  to  a  step   increase   in  C   (indicated  as   Stimulus)  

is  plotted  in  red.  

Ventura  et  al,  SI  Appendix  

17  

 

that   the   nH   of   the   dose-­‐reponse   that   acts   as   the   input   to   the   pulse   generator   has   an   nH,A   slow   >   1  (according  to  point  i)),  resulting  in  a  coupled  sensitivity  that  might  be  higher  for  the  slow  ligand-­‐receptor  reaction  than  for  the  fast  one.  This  is  clearly  in  disagreement  with  the  results  in  Tables  S1  and  S2,  where  we  show  that  the  coupled  sensitivity  is  consistently  lower  than  that  of  the  pulse  generator.    

What   happens   is   that   for   each   ligand   concentration   that   stimulates   the   coupled   system,   the   pulse  generator  reaches  its  maximum  at  a  different  time  tmax  (see  right  panels  in  Fig.  S1  and  also  Fig.  S13b-­‐d).  Particularly,  for  Model  1,  if  the  input  stimulus  is  large,  the  peak  of  X*  occurs  early;  if  the  input  stimulus  is  small,  the  peak  of  X*  occurs  late.  The  binding  dose-­‐response  curve  for  early  times  is  shifted  to  the  right  compared  to  those  of  later  times,  resulting  in  an  effective  ligand-­‐receptor  curve,  that  of  C(tmax),  that  has  lower  slope  than  that  of  C  in  early  times  or  even  than  that  of  C  in  equilibrium  (see  Fig.  S13a,  green  line).  

The   coupled   sensitivity   for  the   ligand-­‐receptor   -­‐>   pulse  generator   system   is   then  bounded   by   the   nH   of   the  curve   formed   by   the   C(tmax)  values   (Fig.   S13).   So,   for   a  pulse   generator   and  a   slow  ligand-­‐receptor   reaction,  the   coupled   sensitivity   will  be   lower   than   that   of   the  pulse  generator  in  isolation,  whether   or   not   the  sensitivity  of  this  last  one  is  lower   or   greater   than   1.  Therefore,   PRESS   itself  lowers   the   sensitivity  of   the  overall   response,   expanding  its  dynamic  range.  In   the   case   of  Model   3,   the  pulse   generator   reaches   its  maximum   at   similar   times  tmax:   0.88,   1.1,   1.3   sec   for  normalized   doses:   1500,  500,  100,  as  indicated  in  Fig.  S14,  resulting  in  an  effective  ligand-­‐receptor   curve,   that  of   C(tmax),   that   has   not   a  much   lower   slope   than   that  of  each  C(t)  curve  (curves  in  grey).   The   overall   result   for  

 

Figure  S13.  Graphical  explanation  of  the  expansion  of  the  input  dynamic  range.  Analysis   of   the   sensitivity   of   the   ligand-­‐receptor   system   coupled   to   the   pulse  generator   in  Model   1.     Parameters:   Xtot=10,   r1=0.1,   r2=0.06,   r3=0.001,   k-­‐=0.001  sec-­‐1.   a.   Normalized   ligand-­‐receptor   complex   C   in   equilibrium   (black),   at   four  different   times   indicated   over   the   figure   (grey),   and   for   the   times   where   the  maximum  of  the  pulse  occurs  for  each  dose,  C(tmax)  (green).  Peak  X*  is  included  in  red.  The  horizontal  axis  corresponds  to  the  dose  of  ligand  normalized  by  the  Kd  of   the   ligand-­‐receptor  reaction.  The  times  selected  to  plot  C  versus  L/Kd  before  equilibrium  are   the   tmax   for   L/Kd=0.1,  1,  10,  100:  6.24,  2.07,  0.83,  and  0.36  min  respectively.  Green  circles  indicate  the  intersection  of  doses  L/Kd=0.1,  1,  10,  100,  and   C   vs   L/Kd   at   the   listed   times,   therefore,   those   circles   are   over   the   curve  C(tmax).  The  curve  C(tmax)  vs  L/Kd  is  characterized  by  EC50=17.7  and  nH=0.9.  b-­‐e.  Temporal  profiles  for  normalized  C  (thin   line)  and  peak  X*  (thick   line)  for  doses  L/Kd=0.1,  1,  10,  100.    

Ventura  et  al,  SI  Appendix  

18  

 

this  model   is   then,  that  the  coupled  sensitivity   is  still   lower  than  the  sensitivity  of  the  pulse  generator  considered  in  isolation  but  the  effect  is  not  as  important  as  when  the  values  of  tmax  are  very  different  for  different  doses  (Model  1,  Fig.  S13).    

   

 

Figure  S14.    Graphical  explanation  of  the  behavior  of  Model  3.  Analysis  of  the  sensitivity  of  the  ligand-­‐receptor  system  coupled  to  the  pulse  generator  in  Model  3.    Parameters:  k-­‐=0.001   sec

-­‐1,   kXB=1,   KXB=0.01,   kFBB=5,   KFBB=100,   FB=0.5,   kCX=5,   KCX=0.1,   kBX=6,   KBX=0.01.   a.   Normalized   ligand-­‐receptor  complex  C  in  equilibrium  (black),  at  three  different  times  from  right  to  left  0.88,  1.1,  1.3  sec  (grey),  and  for  the  times  where  the  maximum  of  the  pulse  occurs  for  each  dose,  C(tmax)  (green).  Peak  X*  is  included  in  red.  The  horizontal  axis  corresponds  to  the  dose  of   ligand  normalized  by  the  Kd  of  the   ligand-­‐receptor  reaction.  The  times   selected   to   plot   C   versus   L/Kd   before   equilibrium  are   the   tmax   for   L/Kd=1500,   500,   and   100,   respectively.  Green  circles  indicate  the  intersection  of  doses  L/Kd=100,  500,  1500,  and  C  vs  L/Kd  at  the  listed  times,  therefore,  those  circles  are  over  the  curve  C(tmax).  The  curve  C(tmax)  vs  L/Kd   is  characterized  by  EC50=666.4  and  nH=1.15.  b.  Temporal  profiles  for  normalized  C  and  peak  X*  for  doses  L/Kd=100,  500,  1500.  

Ventura  et  al,  SI  Appendix  

19  

 

4.  The  effect  of  noise  during  PRESS  

To  determine  the  effect  of  noise  during  PRESS,  we  run  stochastic  simulations  of  the  toy  model  labeled  as  Model  1  in  this  SI  Appendix,  Section2  (the  model  with  an  inactive  refractory  state  in  Fig.  2  of  the  main  text)   varying   the   number   of   receptors.   It   is   well   known   that   at   low   concentrations   deterministic  simulations  based  on  the  law  of  mass  action  are,  in  general,  unlikely  to  be  applicable.  Therefore  we  used  a  stochastic  approach  to  describe   the  binding  kinetics  and  to  account   for   the  noise   resulting   from  the  probabilistic  character  of  the  (bio)chemical  reactions.  The  goal  of   the  study  was  to   introduce  different  degrees  of  noise  at  the  ligand-­‐receptor  binding  step  and  then  analyze  whether  this  noise  was  amplified  

or   not   by   the   downstream   response.   To  do  that,  we  run  simulations  with  different  number   of   receptors   (low,   medium   or  high)   while   maintaining   constant   the  number   of   the   rest   of   species   in   the  system.    

We   include   the   results   of   simulating  Model   1   for   low,   medium,   and   high  number  of  receptors  (30-­‐300-­‐3000),  using  stochastic   methods,   in   Fig.   S2.   We  performed   numerical   simulations   in  Copasi   4.11   (4)   using   its   adaptive   tau-­‐leaping   algorithm.   Importantly,   we  defined   parameter   r1   (x*   activation   rate)  as  0.1/initial   free  receptor,  and  therefore  its   value   was   0.1/30,   0.1/300,   and  0.1/3000,   respectively.   Thus,   each  bound  receptor   has   a   stronger   activation  capacity   when   there   are   only   30   total  receptors   than   when   there   are   300   or  3000.   In  this  way,  the  effective  activation  rate   of   X*   is   comparable   between   the  simulations   with   different   number   of  receptors.    

As   expected,   the   effect   of   noise   is  much  more   evident   when   the   number   of  receptors   is   low,   see   Fig.   S2,   as   can   be  seen  both  in  the  temporal  profiles  on  the  right   panels,   and   in   the   standard  deviation   from   1001   simulations   plotted  

   

Figure  S15.  Percentage  of  overlap  between  the  stochastic  outputs  of  two  consecutive  doses  for  Rtot=30.  Results   of   calculating   the   overlap   in   the   outputs   of   the   1001  stochastic   simulations   performed   for  Model   1   (the  model   with  an   inactive   refractory   state)   for   low   number   of   receptors   (30  receptors  per   cell),   using   stochastic  methods.   Simulations  were  done   in   Copasi   using   its   adaptive   tau-­‐leaping   algorithm.  Parameter   values:   k+=0.0002   ml/(pmol*sec),   k-­‐=0.001   1/s,  r1=0.01   ml/(pmol*sec),   r2=0.06   1/sec,   r3=0.001   1/sec   (with   1  pmol/ml   =   1   nM).   Initial   particle   numbers:   receptor=30,  x=14536.5.   Cell   volume=   5e-­‐11  ml.   The   overlap   was   calculated  with   three   different   variables:   the   equilibrium   value   of   the  ligand-­‐receptor  complex  C  (Ceq)  in  red,  maximum  of  X*  (X*max)  in  black,  and  C  at  the  time  where  X*  reaches  its  maximum  (C(tmax))  in  blue.   Labels   in  horizontal  axis   indicate   the   lower  dose  of   the  two  doses  being  compared  for  the  overlap,  normalized  by  the  Kd  of   the   ligand-­‐receptor   reaction.     Doses   being   compared   are   (in  units  of  Kd):  0.01,  0.03,  0.1,  0.32,  1.0,  3.16,  10,  31.62,  100.  

Ventura  et  al,  SI  Appendix  

20  

 

on  the  left  panels.  The  square  of  the  coefficient  of  variation  plotted  in  the  bottom  panel  indicates  that  for  the  three  variables,  Css  (red  curves),  C(tmax)  (blue  curves),  and  peak  X*  (black  curves),  lower  receptor  number  results   in  higher  CV2  for  the  whole  range  of  doses  considered.    Also,  for  the  three  amounts  of  receptors  considered  (30  in  solid  lines,  300  in  dashed  lines,  and  3000  in  dotted  lines),  the  curve  of  Css  is  lower  than  that  of  peak  X*,  and  peak  X*  is  lower  than  the  curve  for  C(tmax).  This  result  suggests  that  even  in  the  case  of  very  low  number  of  receptors  (30  per  cell),  which  introduce  a  significant  noise  at  the  level  of  occupied  receptor  (Fig.  S2F),  the  pulse  generator  downstream  of  the  receptor  did  not  amplify  it  and  therefore  the  improvement  gained  by  PRESS  is  not  lost  or  masked  by  the  effects  of  noise.    In  Fig.  S15,  we  show  the  percentage  of  overlap  between  the  outputs  of  1001  stochastic  simulations  for  two  consecutive  doses,  Li/Kd  and  Li+1/Kd,  being  Li+1/Kd~3*  Li/Kd  ,  evaluated  for  Ceq  (red),  C(tmax)  (blue)  and  X*max    (black),  in  the  case  of  low  number  of  receptors  (30  receptors  per  cell).  The  aim  of  this  figure  is  to  evaluate   if   two   doses   can   be   identified   as   different   doses   based   on   the   information   given   by   each  variable,   Ceq,   C(tmax),   X*max,   particularly   in   the   region   of   high   doses   (L/Kd   >1).   The   results   in   Fig.   S15  indicate   even   in   this   case   with   significant   noise   at   the   level   of   occupied   receptor,   variable   X*max   has  lower  overlap  than  Ceq,  supporting  that  the   improvement  gained  by  PRESS   is  not   lost  by  the  effects  of  noise.    

 

Ventura  et  al,  SI  Appendix  

21  

 

5.  Mathematical  model  to  study  the  sensing  of  a  stationary  spatial  gradient  

In   what   follows   we   present   two   alternative   mathematical   descriptions   for   the   variable   Delta.   In  Description   1   we   derived   the   steady-­‐state   concentration   of   αF   as   an   approximate   solution   to   the  diffusion  equation   from  a  point   source,  with,   and  also  without,   reflecting  boundary   conditions  on   the  sensing   sphere.  Combining   this   solution  with   the  normalized  bound   receptor   function  we  derived   the  variable   Delta   (difference   of   bound   receptors   at   the   front   and   at   the   back).   In   Description   1,   Delta  depends  on  three  combinations  of  parameters  related  the  average  concentration  sensed  by  the  cell,  the  ratio  between   the   radius  of   the   cell   and   the  distant   to   the   source,   and   the   ligand-­‐receptor  unbinding  rate.   In   Description   2,   we   derived   variable   Delta   in   terms   of   the   maximal   and   minimal   ligand  concentrations   sensed   by   the   cell,   without   relating   them   by   any   particular   function   (i.e.   without  specifying  the  profile  of  the  gradient).  In  this  case,  Delta  still  depends  on  the  ligand-­‐receptor  unbinding  rate,  and  on  the  maximal  and  minimal  ligand  concentrations.  Description  1  is  based  on  a  given  source-­‐sensing   cell   pair,   while   Description   2   is   based   on   the   local   environment   sensed   by   the   cell.   Both  descriptions  are  easily  related  to  one  another.  

5.1   An   analytical   expression   for   the  αF   gradient   generated   by   a   point   source,   and   for   the   variable  Delta  (Description  1)  

As  in  Segall  (5),  the  cell  is  modeled  as  an  impermeable  sphere  of  radius  a,  with  d  the  distance  from  the  center  of   the   sphere   to   the  point   source  of  αF.   There   is   azimuthal   symmetry  with   respect   to   the   line  between  the  center  of  the  sphere  and  the  point  source.  The  concentration  at  any  point  on  the  sphere  will  vary  with  the  angle,  θ,  between  the  line  connecting  the  center  of  the  sphere  to  the  point  source  and  the  line  connecting  the  center  of  the  sphere  to  the  point  on  the  surface.  The  point  over  the  sphere  that  is  nearest  the  source  is  indicated  with  an  f  (front,  θ  =  0)  and  the  point  that  has  the  greatest  distance  to  the  source  is  indicated  with  a  b  (back,  θ  =  π).  An   approximate   solution   to   the   diffusion   equation  with   a   point   source   and   given   reflecting   boundary  conditions  on  the  sphere  (5)  is:  

  (3)  

where  αF(θ)  is  the  steady-­‐state  concentration  of  αF  as  a  function  of  the  position  over  the  cell  surface,  q  is  the  rate  of  release  of  alpha-­‐factor  from  the  point  source,  D  is  the  αF  diffusion  coefficient,  and  Pj  is  the  Legendre  polynomial  of  order  j  (5).  We  model  the  ligand-­‐receptor  binding  with  the  following  reaction:  

αF  +  R    <-­‐>  C     (4)  where  αF   is   the   ligand  (alpha-­‐factor),  R   is   the  membrane  receptor  assumed  to  have  a  uniform  density  over  the  cell  surface,  C  is  the  complex  ligand-­‐receptor,  and  k+  and  k-­‐  are  the  binding  and  unbinding  rates,  respectively.  The  dissociation  constant  Kd  is  defined  as  k-­‐/k+  and  represents  the  ligand  concentration  that  produces  a  bound  receptor   level  that   is  one-­‐half  of   its  maximum.  The  total  concentrations  of  receptor  and  ligand  in  the  system  are  assumed  to  be  constant,  Rtot  =  [R]  +  [C],  αFtot  =  [αF]  +  [C].  Thus,  only  one  concentration  of  the  three  ([R],   [αF],  and  [C])   is   independent;   the  other  two  may  be  determined  from  

Ventura  et  al,  SI  Appendix  

22  

 

Rtot,  αFtot   and   the   independent   concentration.   The   kinetics   of   this   system   can   be   solved   analytically.  Choosing   [C]   as   the   independent   concentration   and   representing   the   concentrations   without   the  brackets  for  brevity  (e.g.  R  =  [R]),  the  kinetic  rate  equation  can  be  written  as:  

    (5)  

where  we  have  assumed  that  [αF]  ~  αFtot,  meaning  that  the  consumption  of  ligand  in  the  ligand-­‐receptor  reaction  is  negligible  compared  with  αFtot.  αFtot  is  then  the  function  αF(θ)  given  by  Eq.  (3).  The  solution  for  Eq.  (5)  is:    

,      (6)  

with    

  (7)  

being  the  steady-­‐state  value  for  C  and     (8)  

the  time  required  for  the  bound  receptor  C  to  rise  from  zero  to  1−1/e  (that  is,  63.2%)  of  its  final  steady-­‐state  value.  

Combining  Eqs.  (3),  (6),  (7),  and  (8)  we  obtain  that  the  normalized  bound  receptor,  C/Rtot  as  a  function  of  θ  and  t  depends  on  the  following  combination  of  parameters:  

  (9)  

Parameter  p1   combines  parameters   related   to   the   ligand   source   (q;D),   the  distance   to   the   source   (d),  and  Kd,  and  indicates  the  average  concentration  sensed  by  the  cell  in  multiples  of  Kd,  for  instance,  p1=1  means   that   the   cell   feels   a   concentration   of   about   1   ×   Kd.   Parameter   p2   is   a   geometrical   parameter  defined  as  the  ratio  between  the  radius  of  the  spherical  cell  and  the  distance  from  the  center  of  the  cell  to   the   point   source.   p1   and   p2   are   dimensionless   parameters.   Parameter   p3   is   the   ligand-­‐receptor  unbinding  rate  and  has  units  of  1/time.  The  three  parameters  are  always  positive,  and  p2  is  such  that  0  <  p2  <  1.  

For  p1  =  10  and  p2  =  0.3  (the  values  used  in  most  of  the  paper),  the  total  numbers  of  bound  receptors  in  steady  state,  for  the  hemispheres  facing  and  away  from  the  point  source,  were  calculated  following  (5)  by   integration   over   the   angles   0   to   π/2   radians   (0   to   90   degrees)   and   π/2   to   π   radians   (90   to   180  degrees),   respectively.   The   total   number   of   receptors   was   assumed   to   be   8000   (5).   The   hemisphere  nearest   the   source  has  3720  occupied   receptors   and   the  hemisphere  away   from   the   source  has  3531  occupied  receptors.  The  difference  between  the  two  sides,  189  receptors,  is  2.6%  of  the  total  number  of  occupied  receptors.    The  variable  Delta   is  defined  as  Delta(t)  =  C(θ  =  0,   t)  −  C(θ=π,   t)   (difference  of  bound  receptors  at   the  front  and  at   the  back).  The  effectiveness  of  PRESS   in   the  context  of  gradient  sensing  depends  on  how  much   (amplitude)   and   for   how   long   (duration)   lasts   the   overshoot   of   Delta(t).   Delta   depends   on  parameters  p1,  p2,  and  p3.      

Ventura  et  al,  SI  Appendix  

23  

 

By  combining  Eqs.  (3),  (6),  (7),  and  (8)  we  numerically  compute  Delta  and  determine  that  the  overshoot  amplitude  increases  with  the  concentration  of  ligand  (which  increases  p1).  This  is  because  Deltaeq  has  a  biphasic  behavior  with  p1  while  Deltamax  has  not  (Fig.  S16a  inset).  Deltaeq  peaks  at  about  p1=1,  i.e.  at  an  average  concentration  of  αF  of  about   the  Kd.  On   the  other  hand,  Deltamax   is  virtually   indistinguishable  

from   Deltaeq   for   low   values   of   p1,   but   at   p1   approaching   1   Deltamax   diverges   from   Deltaeq   increasing  asymptotically   to   a   maximum   Deltamax.   Consequently,   at   values   above   1,   increasing   p1   increases   the  amplitude  of   the  overshoot.   In  contrast,   its  duration  decreases  with  p1   (Fig.   S16b).  Thus,  even   though  high   levels   of   signal   provide   greater   overshoot   amplitude   for   PRESS,   the   time   window   available   for  PRESS  is  reduced.    

Increasing  p2  for  a  given  value  of  p1  (which  translates  into  increasing  the  sensing  cell  size),  increases  both  Deltamax  and  Deltaeq,  effectively  improving  the  chances  of  the  cell  of  detecting  differences  between  front  

! a b

 

Figure  S16.  Characterization  of  PRESS  applied  to  gradient  detection.  a.  Amplitude  of  Delta’s  overshoot  as  a  function  of  p1  (average  ligand  concentration  relative  to  the  Kd)  for  the  indicated   values   of   p2.   Overshoot   occurs   at   p1   greater   than   ~   1,   providing   an   advantage   of   PRESS   over  equilibrium   signaling.   p3   does   not   affect   Deltaeq   or   Deltamax.   Inset.   Dependency   of   Deltaeq   (solid   line)   and  Deltamax   (dashed   line)   on   parameter   p1   (average   concentration   of  αF)   for   p2=0.3.   Deltaeq   is   biphasic  with   p1  while  Deltamax  is  monotonic.  The  shaded  area  represents  Delta’s  overshoot.  b.  Temporal  window  during  which  Delta  overshoots  Deltaeq  as  a  function  of  parameter  p3  (the  time  scale  of  the  αF  binding  reaction),   for  p2=0.3  and  the  indicated  values  of  p1.  The  width  of  each  line  (shaded)  marks  the  region  between  two  extremes  yeast  cell  sizes  (cells  from  1.6  to  6.4  µm  diameter,  located  at  a  distance  from  the  source  of  6.7  µm),  corresponding  to  p2  from  0.12  to  0.48.  Inset  shows  Delta  versus  time  for  the  two  values  of  p3  indicated  in  the  x-­‐axis  (p1  =  10,  p2  =  0.3).  Graph  shows  that  the  temporal  profile  is  independent  of  p3,  except  for  the  scale  of  time.  We  defined  the  temporal  window  for  PRESS  as  the  interval  during  which  Delta  is  higher  in  more  than  20%  of  Deltaeq.  

 

Ventura  et  al,  SI  Appendix  

24  

 

and  back  bound  receptor,  both  at  equilibrium  and  prior   to   it,  but   it  does  not  make  PRESS  better   than  equilibrium  signaling  (Fig.  S16a).  

Most  interestingly,  the  duration  of  Delta’s  overshoot  is  most  strongly  affected  by  p3  (Fig.  S16b).  For  Cell  1   in  our  example  of  Fig  3B   in   the  main   text,  a  value  of  p3  100  times   faster   than  the  actual  αF  binding  reaction  reduces  the  time  window  for  PRESS  from  about  6  minutes  to  about  0.06  minutes  (3.6  seconds)  (inset  in  Fig.  S16b).  Thus,  the  slower  p3,  the  more  time  there  is  for  PRESS.  On  the  other  hand,  p3  does  not  affect  the  amplitude  of  the  overshoot.  

5.1.1  A  simplified  formula  for  the  αF  gradient  generated  by  a  point  source,  and  the  variable  Delta    

If   the   effect   of   the   reflecting   boundary   conditions   on   the   sensing   sphere   is   not   considered,   then   the  steady-­‐state  solution  for  the  problem  of  diffusion  from  a  point  source  is  simplified  to:  

,   (10)  

In  terms  of  parameters  p1,  p2  and  p3,  and  for  the  case  with  reflecting  boundary  conditions,  we  have:  

,   (11)  

and  for  the  case  without:  

    (12)  

This   last   case   results   in  αFf  =  αF(θ   =  0)  =  p1/(1  −  p2)  and  αFb  =  αF(θ   =  π)  =  p1/(1  +  p2).   In  Fig.   S17  we  compare   the   α-­‐factor   profiles   over   the   sensing   cell   surface   with   and   without   reflecting   boundary  conditions,  i.e.  we  compare  Eqs.  (11)  and  (12).  If  the  presence  of  the  sphere  is  taken  into  account  in  the  diffusion  problem  (Eq.  (11)),  the  differences  between  front  and  back  are  greater  than  if  not.  

 With  αF  given  by  Eq.  (12)  we  obtain:  

             (13)  

                 (14)  

where  

  (15)  

  (16)  

resulting  in                          (17)  

    (18)  

Ventura  et  al,  SI  Appendix  

25  

 

 (19)  

 

In  Fig.  S18  we  compare  the  variable  Delta  calculated  using  Eq.  (11)  or  Eq.  (12),  we  can  see  that  in  both  cases   there   is   a   time   window   where   Delta   is   greater   than   its   steady-­‐state   value,   meaning   that   the  feature  of  the  overshoot  in  Delta  does  not  depend  on  the  use  or  not  of  reflecting  boundary  conditions  over  the  sensing  sphere.  

We  make  use  of  this  simplified  scheme  to  analytically  explain  some  of  the  features  in  Fig.  S17.  From  Eq.  (18)  we  see   that  Deltaeq  depends  on  parameters  p1  and  p2  but  not  on  p3.  Computing   the  derivative  of  Deltaeq  with  respect  to  p1  and  equaling  to  zero,  we  obtain  that  Deltaeq  reaches  a  maximum  at    p1~  =(1  −  p2

2)1/2,  it  increases  with  p1  for  p1  <  p1~,  then  decreases.  For  a  small  value  of  parameter  p2,    p1~  is  close  to  1,  for  instance,  for  p2  =  0.3  (the  value  p2  used  in  the  majority  of  the  paper),  p1~  =  0.95.  The  derivative  of  Deltaeq  with  respect  to  p2  is  always  positive,  meaning  that  Deltaeq  always  increases  with  p2.  

Using  Eqs.  (17)-­‐(19)  we  can  explicitly  compute  the  time  t+  where  Delta  becomes  greater  than  Deltaeq,  i.e.  when  f(t)  becomes  positive:  

  (20)  

and  the  time  tmax  where  Delta  reaches  a  maximum:  

  (21)  

The  maximum  value  reached  by  Delta  is  given  by  Deltamax  =  Delta(t  =  tmax):  

.    (22)  

!  

Figure  S17.  α -­‐factor  profiles  over  the  sensing  cell  surface.  Including   (black)   not   (red)   reflecting   boundary  conditions.  αF  is  normalized  by  Kd  and  θ   is  plotted  in  π  units.  

!  

Figure  S18.  Delta  vs.  time  Data   was   calculated   with   (black)   and   without   (red)  reflecting  boundary  conditions,  for  p1=10,  p2=0.3,  and  p3=0.001  1/sec.  

 

Ventura  et  al,  SI  Appendix  

26  

 

Deltamax  depends  on  parameters  p1  and  p2  but  not  on  p3.  By  plotting  Deltamax  versus  parameter  p1   for  different  values  of  p2,  or  Deltamax  versus  parameter  p2  for  different  values  of  p1  (not  shown),  we  can  see  that  it  increases  monotonically  with  both  p1  and  p2.  From  this  last  equation  we  see  that  for  small  values  of  p1,  the  second  term  inside  the  parenthesis  is  very  small  resulting  in  Deltamax  ~  Deltaeq.  

5.2    An  analytical  expression  for  the  variable  Delta  (Description  2)  

The  effectiveness  of  PRESS  in  the  context  of  gradient  sensing  depends  on  the  amplitude  and  duration  of  the  overshoot  of  Delta(t).  Thus,  we  developed  a  mathematical  analysis   to  describe   the  evolution  over  time   of   Delta(t),   assuming   homogeneously   distributed   receptors   and   simple   binding   kinetics.   Delta(t)  depends  on  the  properties  of  the  gradient:  the  maximum  and  minimum  ligand  concentrations  sensed  by  the  cell  (in  the  case  of  αF,  αFf  and  αFb,  the  front  and  back  αF  concentrations,  respectively),  as  well  as  of  the  sensing  cell:  the  binding  and  unbinding  rate  for  the  ligand-­‐receptor  reaction,  k+  and  k-­‐.  Normalizing  by  Rtot  we  obtain  Delta(t)  as  follows:  

 

   Where    and    are   the   front   and   back   concentrations   of   αF   normalized   by   the   dissociation  constant,   Kd.   The   first   term   (in   black)   does   not   depend   on   time,   and   it   corresponds   to   the   value   of  Delta(t)  at  equilibrium,  Deltaeq.  The  second  and  third  terms  (in  blue  and  red)  correspond  to  the  dynamics  of   binding   at   the   back   (in   blue)   and   the   front   (in   red).   We   then   derived   a   formula   for   Deltamax,   the  maximum   value   of   Delta(t),   its   difference   with   Deltaeq   define   the   overshoot;   and   tmax,   the   time   of  Deltamax,   with   which   we   obtained   the   overshoot’s   duration   (arbitrarily   defined   as   temporal   window  during  which  Delta(t)  is  20%  larger  than  Deltaeq).    The  function  Delta(t)  can  be  characterized  by:  1)  its  initial  value,  Delta(t=0)=0;  2)  its  asymptotic  value,  Delta(t-­‐>∞):  

;  

3)  the  time  tmax  where  Delta(t)  reaches  a  maximum:  

;  

4)  the  maximum  value  reached  by  Delta(t)  ,  given  by  Deltamax  =  Delta(t  =  tmax):  

 

5)  the  time  t+  where  Delta  becomes  greater  than  Deltaeq:  

Ventura  et  al,  SI  Appendix  

27  

 

 

6)  the  temporal  window  where  Delta  is  greater  than  Deltaeq  ,  approximated  as  Δt=2*(tmax-­‐t+):  

 The  main  quantities  in  the  context  of  PRESS  and  the  overshoot  in  Delta(t)  are  Deltamax,  Deltaeq  and  the  temporal  window    ,  where   Delta(t)   is   greater   than  Deltaeq.   Deltamax  and   Deltaeq   depend   on    and  

 only,  and      depends  on  those  two  quantities  and  also  on  k-­‐.  Linear  and  exponential  gradients  are  represented  in  the    ( )  plane  as  lines  parallel  to  the  identity  with  different  y-­‐intercepts,  or  lines  with  zero  y-­‐intercept  and  different  slopes,  respectively,  as  described  with  the  following  equations:  

linear  gradient:   ,   ,   ,    

exponential  gradient: , ,  (with  const2>1)    The   biological   relevance   of   the   overshoot’s   duration   depends   on   the   time-­‐scale   of   the   processes  downstream  to  the   ligand-­‐binding  reaction  (i.e.,   is   the  downstream  process   fast  enough  to  generate  a  response  during  the  temporal  window  where  Delta(t)  >  Deltaeq?).  The  values  of  Delta(t)  are  differences  between   the   fraction   of   bound   receptor   at   the   front   and   back,   and   whether   a   given   value   of   Delta  provides   enough   directional   information   to   detect   a   given   gradient   depends   on   the   particular  mechanism  of  polarization.    To  show  the  behavior  of  Deltaeq,  Deltamax,  and  the  overshoot’s  duration  as  a  function  of  the  gradient,  we  focused  on  the  case  of  yeast  mating  pheromone  gradient  detection.   In  Fig.  S3  A-­‐C  we  show  heatmaps  where  the  color  scale  indicates  Deltaeq,  Deltamax  and  the  overshoot’s  duration,  respectively,  for  a  range  of  gradients  (αF  front,  αF  back  pair  combinations)  in  a  region  of  αF  concentrations  from  1  to  20  Kds.  For  Deltaeq  and  Deltamax,  given  a  value  of  αF  at   the  back,   increasing  αF  at   the   front   improves  both  Deltaeq  and  Deltamax  (and  the  same  is  true  for  a  given  value  of  αF  front  but  decreasing  αF  back),  indicating  that,  as  expected,  a  more  pronounced  gradient  improves  Deltaeq  and  Deltamax.    We  considered  the  profile  of  Delta(t)  for  a  cell   located  in  an  exponential  gradient  at  an  average  ligand  concentration  of  ~1⨯Kd  (see  Fig.  3,  main  text).   In  this  case,  a  cell  reaches  a  value  of  Deltaeq  =  0.23.  For  heuristic  reasons,  we  used  this  value  as  reference  point,  assuming  that  a  value  of  Delta  of  0.23  or  larger  provides  enough  front-­‐back  difference  to  distinguish  the  direction  of  the  gradient,  and  that  values  lower  than  0.23  result  in  suboptimal  gradient  detection.  The  region  to  the  left  of  the  contour  line  in  Fig.  S3  A  gives   all   combinations   of   αF   front-­‐αF   back   that   would   lead   to   consistent   gradient   detection   if   only  equilibrium  information  is  used.  However,  the  region  of  αF  front-­‐αF  back  combinations  where  Deltamax  is   greater   than   0.23   is  much   larger   than   thatobtained   if   we   consider   equilibrium   binding   information  (dashed  contour  line  in  Fig.  S3  B  marks  combinations  with  Deltamax  of  0.23).  In  the  region  between  the  Deltaeq   and   Deltamax   lines   of   0.23,   pre-­‐equilibrium   information   is   sufficient   (Deltamax>0.23)   but  equilibrium   information   is   not   (Deltaeq<0.23).   In   Fig.   S3   C   we   show   the   temporal   window  where   this  amplification  of  the  differences  between  front  and  back  occurs.  For  the  region  between  the  Deltaeq  and  

Ventura  et  al,  SI  Appendix  

28  

 

Deltamax   lines,   the   temporal  window   is   not   smaller   than  6.8  min.   The  duration  of  Delta’s   overshoot   is  strongly  affected  by  k-­‐:  a  value  of  k-­‐  100  times  faster  than  the  reported  αF  binding  reaction  reduces  it  by  a  factor  100  (duration  reaches  a  maximum  of  15  min  for  k-­‐=0.001  sec-­‐1,  but  only  of  0.15  min  for  k-­‐=0.1  sec-­‐1).  Thus,  the  slower  k-­‐,  the  more  time  available  for  PRESS.  On  the  other  hand,  k-­‐  does  not  affect  the  amplitude  of  the  overshoot  (see  formula  for  Deltamax    above).  With   this   analysis,   we   studied   Delta(t)   in   different   gradient   situations.   We   show   that   the   average  concentration  establishes  a  tradeoff  between  overshoot’s  amplitude  and  duration.  Deltamax   is   larger  at  higher  average  ligand  concentrations,  but  the  temporal  windows  are  shorter  (see  grey  lines  in  Fig.  S3  B  and  C).  The  optimal  condition  depends  on  the  downstream  process  time  scale  and  sensitivity.    Summarizing:  -­‐  Deltaeq   and  Deltamax  depend  only   on   and   .     A   higher   average  αF   concentration   and   a  more  pronounced  gradient  are  conditions  that  favor  Deltaeq  and  Deltamax.  -­‐   The   temporal   window   for   PRESS   depends   on   and    and   also   on   the   unbinding   rate   for   the  receptor-­‐αF  reaction,  k-­‐  (at  a  fixed  Kd).  It  is  shorter  for  higher  averages  of  αF  concentration  and  for  faster  unbinding  rates.    

Ventura  et  al,  SI  Appendix  

29  

 

 

6  Supplemental  experimental  procedures    

6.1  Strains  and  plasmids.  

Strains  are  detailed  in  Table  S3.  YACL379  was  used  as  the  parental  strain,  which  is  a  Δbar1  strain  derived  from   YAS245-­‐5C   (can1::HO-­‐CAN1   ho::HO-­‐ADE2   ura3   ade2   leu2   trp1   his3),  which   in   turn   is   a  W303-­‐1a  descendant  (6).  

ESY3136  was  made  by  transforming  YACL379  using  a  PCR  product  containing  the  YFP  gene  followed  by  the  S.  pombe  his5+  gene.  We  directed  recombination  downstream  of  the  STE20  gene  using  primers  with  40  nt  of  homology  with  the  3’end  of  the  STE20  ORF  (7).  MWY003  was  made  in  two  steps.  First,  STE5  was  deleted  using  a  PCR  product  containing  the  nourseothricin  (NAT)  resistance  gene  obtained  using  primers  with  40  nt  of  homology  with  the  5’  and  3’  ends  of  the  STE5  ORF.  Then,  this  strain  was  transformed  with  a  pRS404  plasmid  containing  the  STE5  gene  (from  -­‐1193  to  the  end  of  the  ORF)  followed  by  three  copies  of  YFP,  linearized  at  the  STE5  promoter  with  PacI.  The  fluorescence  level  of  MWY003  is  three  times  that  of   other   colonies   aislated   from   this   transformartion,   suggesting   that   three   integrations   of   the   STE5-­‐YFPx3   construct   ocurred.   YAB3725  was  made   by   transforming   YACL379  with   plasmid   pBB-­‐Ste2(T305)-­‐CFP  linearized  with  ClaI.  pBB-­‐Ste2(T305)-­‐CFP  is  a  pRS406  based  plasmid  that  contains  a  514  nt  fragment  of  STE2  (from  nt  +402  to  +916  counting  from  the  START  codon)  followed  by  CFP.  In  this  plasmid,  ClaI  cuts  within   the   STE2   fragment,   and   therefore,   it   integrates   at   the   STE2   locus,   generating   a   strain   that  expresses  only  a  Ste2  protein  truncated  at  T305  followed  by  CFP.  

 

 

We  made  YGV5097  by  PCR-­‐based  gene  tagging  (7)  of  the  BEM1  ORF  in  strain  ACL379.  For  the  PCR,  we  used  pFA6a-­‐mNG(3x)-­‐KANMX6  as  template  and  two  primers  with  40nt  homology  5’ends  to  the  3’  end  of  BEM1   ORF   and   its   3’UTR,   respectively.   We   made   pFA6a-­‐mNG(3x)-­‐KANMX6   in   two   steps.   First,   we  designed  and  ordered  the  synthesis  of  a  DNA  sequence  coding  for  a  tandem  repeat  of  three  monomeric  NeonGreen   (mNG)   fluorescent   proteins   (8)   flanked   by   a   5’   PacI   and   3’   AcsI   (Integrated   DNA  Technologies,  Inc.,  Coralville,  Iowa).  For  this  design  we  optimized  codon  usage  for  yeast.  For  preventing  

Table  S3.  Strains.  Strain   Relevant  Genotype   Reference  

YACL379   MATa  ∆bar1   (6)  

ESY3136   MATa  ∆bar1  ste20::STE20-­‐YFP::his5+   This  study  

MWY003   MATa  ∆bar1  ∆ste5::Nat,  (trp1::STE5-­‐YFPx3::TRP1)x3   Peter  Pryciak  lab  

YAB3725   MATa  ∆bar1  ste2::PSTE2  STE2(T305)-­‐CFP::URA3   This  study  

YGV5097   MATa  ∆bar1  BEM1::BEM1-­‐3xmNeonGreen-­‐HIS3MX6   This  study  

Ventura  et  al,  SI  Appendix  

30  

 

intra-­‐repeat   recombination,   we   designed   each   copy   of   mNG   to   use   slightly   different   codons.   The  resulting  plasmid  is  called  pUCIDT-­‐mNG(3x).  In  a  second  step,  we  subcloned  a  PacI  and  AscI  cut  fragment  containing  mNG-­‐(3X)   from  pUCIDT-­‐mNG(3x)   into   the  pFA6-­‐GFP-­‐HIS3MX6  backbone  also  digested  with  PacI   and   AscI   to   release   the   GFP   sequence.   Finally,   the   PCR   product   was   used   for   transformation   of  ACLY379.    

6.2  Quantification  of  Polarization  Times    

Microscopy  experiments  were  performed  as  described  elsewhere  (9,  10).  Exponential  growing  cells  were  placed  in  384-­‐well  glass-­‐bottom  plates  pre-­‐treated  with  1mg/ml  of  concanavalin  A  (Sigma)  to  affix  cells.  Images   were   acquired   using   an   Olympus   FV1000   confocal   module   mounted   on   an   Olympus   IX-­‐81  microscope,  with  an  Olympus  UplanSapo  63x  objective  (NA  =  1.35).    

Cells  were   stimulated  with   1μM  α-­‐factor   at   time   zero   and   the   same   field  was   followed   for   up   to   2.5  hours.   Transmission   and   YFP     (excitation   515nm,   emission   530-­‐630nm)   images  were   analyzed   by   our  software  Cell-­‐ID  (10),  identifying  individual  cells.  Using  our  package  Rcell  (http://cran.r-­‐project.org/),  we  created   transmission   image   montages   as   the   ones   showed   in   Fig   3F.   Based   on   these   montages,   the  position   in   the   cell   cycle   for   each   cell   was   manually   determined;   pre-­‐start   cells   (G1)   responded   to  pheromone   directly   while   post-­‐start   cells   (S)   budded   and   finished   a   round   of   replication   before  responding  to  pheromone  (Fig  S4A).  Analogously  image  montage  for  YFP  channel  (as  the  one  showed  in  Fig.  S2B)  were  created  with  Rcell.  To  avoid  missing  out-­‐of-­‐focus  polarizations,  a  Z  stack  was  acquired  at  each   time.   Polarization   time   was   determined   as   the   first   time   in   which   a   sustained   polarization   was  observed,   for  example  16.5  minutes   in  Fig.  S4B.  To  avoid  any  bias,   the  cell-­‐cycle  and  polarization-­‐time  determinations   were   done   independently,   and   the   order   in   which   the   cells   were   presented   was  randomized.    

Three   independent  repetitions  of  each  strains  were  done,   finding  very  good  repeatability.  To  estimate  the   uncertainty   in   the   cumulative   polarization   curves   (Fig.   S4C),   repetitions   were   pooled   and   95%  confidence  intervals  were  calculated  from  1000  bootstrap  samples  (11).    

Ventura  et  al,  SI  Appendix  

31  

 

7.  Triple  monomeric  NeonGreen  DNA  sequence.  

ATGGTCTCAAAAGGGGAGGAGGACAACATGGCATCCTTGCCAGCAACTCACGAATTACATATTTTCGGAAGTATTAATGGTGTTGACTTTGATATGGTTGGACAAGGCACTGGTAATCCAAACGATGGATACGAGGAACTTAATCTTAAATCAACAAAAGGCGATTTGCAATTTTCCCCCTGGATTCTTGTGCCCCATATTGGGTACGGCTTTCATCAATATTTACCTTACCCTGATGGAATGAGTCCATTTCAAGCCGCTATGGTAGACGGTTCTGGCTACCAAGTTCATAGAACCATGCAGTTTGAAGACGGCGCTTCCTTAACAGTAAACTACAGGTACACGTACGAGGGTTCCCATATTAAGGGAGAAGCCCAAGTTAAGGGTACTGGTTTTCCGGCTGATGGTCCCGTTATGACTAATAGCCTGACAGCTGCTGACTGGTGTAGGTCTAAAAAGACATATCCTAACGATAAAACAATAATTTCAACTTTTAAGTGGTCTTATACCACTGGCAATGGTAAAAGATACAGATCCACTGCCCGTACTACGTATACATTCGCTAAACCAATGGCCGCTAATTATCTTAAAAATCAACCCATGTACGTGTTTAGGAAGACTGAACTAAAACATAGTAAGACCGAGTTGAACTTTAAGGAATGGCAAAAAGCCTTCACCGATGTCATGGGAATGGATGAATTATATAAGATGGTTTCCAAGGGTGAGGAAGATAATATGGCATCTTTGCCAGCCACACACGAATTGCACATATTTGGTTCTATTAATGGAGTGGACTTCGATATGGTGGGTCAAGGTACTGGCAACCCTAATGACGGATATGAAGAACTTAATCTGAAGTCTACCAAAGGCGATTTACAGTTTTCTCCTTGGATCTTAGTTCCTCATATCGGCTATGGTTTCCATCAATACTTGCCTTACCCGGATGGTATGAGCCCGTTTCAAGCTGCGATGGTAGATGGTTCAGGTTATCAAGTCCATAGAACCATGCAATTTGAAGACGGTGCGTCTTTAACTGTAAATTACAGATACACTTATGAAGGTTCACACATTAAAGGAGAAGCTCAAGTTAAAGGTACAGGCTTTCCAGCAGATGGTCCTGTCATGACCAATTCTCTAACAGCTGCTGATTGGTGTAGATCAAAAAAGACTTACCCAAACGATAAAACAATTATATCTACCTTTAAATGGTCCTACACTACTGGCAATGGTAAAAGATACAGGAGCACTGCTCGTACTACTTACACGTTCGCTAAACCGATGGCTGCAAATTACTTGAAAAATCAACCAATGTATGTATTTAGAAAAACTGAATTGAAGCATTCTAAAACAGAGCTTAACTTTAAAGAATGGCAAAAAGCATTCACAGACGTTATGGGCATGGACGAATTGTATAAAATGGTCTCTAAGGGTGAAGAAGATAATATGGCGTCTCTACCAGCTACACACGAACTGCATATCTTCGGTAGCATCAACGGTGTAGACTTCGATATGGTCGGACAAGGCACCGGAAACCCAAATGATGGGTATGAAGAATTGAACCTAAAGTCAACGAAGGGTGACTTACAGTTTAGCCCTTGGATTCTAGTGCCACACATTGGTTATGGTTTTCATCAATACCTACCTTACCCCGACGGAATGTCACCATTTCAAGCCGCTATGGTTGATGGTTCCGGTTATCAAGTTCACAGAACTATGCAGTTTGAAGATGGAGCTTCCCTGACTGTCAACTACAGGTACACTTACGAAGGTTCTCATATCAAGGGTGAAGCACAAGTTAAAGGAACCGGTTTTCCTGCCGATGGACCCGTTATGACTAACTCTTTGACGGCTGCCGATTGGTGCAGGAGTAAAAAGACCTACCCAAACGATAAAACTATCATCAGCACTTTTAAGTGGTCATATACTACCGGGAACGGAAAGAGATATAGATCAACCGCTAGAACGACATACACATTTGCCAAGCCTATGGCAGCTAACTACTTAAAAAATCAACCTATGTACGTGTTTCGTAAAACCGAATTGAAGCATAGCAAGACAGAATTAAATTTCAAGGAATGGCAAAAAGCCTTCACCGATGTTATGGGTATGGACGAACTTTATAAATAG  

 

Ventura  et  al,  SI  Appendix  

32  

 

8.  Supplemental  References  

1.   Ma  W,  Trusina  A,  El-­‐Samad  H,  Lim  W,  &  Tang  C  (2009)  Defining  network  topologies  that  can  achieve  biochemical  adaptation.  Cell  138(4):760-­‐773.  

2.   Koshland  DE,  Jr.,  Goldbeter  A,  &  Stock  JB  (1982)  Amplification  and  adaptation  in  regulatory  and  sensory  systems.  Science  217(4556):220-­‐225.  

3.   Ferrell  JE  (2001)  Regulatory  Cascades:  Function  and  Properties.  eLS,    (John  Wiley  &  Sons,  Ltd).  

4.   Hoops  S,  et  al.  (2006)  COPASI-­‐-­‐a  COmplex  PAthway  SImulator.  Bioinformatics  22(24):3067-­‐3074.  

5.   Segall  JE  (1993)  Polarization  of  yeast  cells  in  spatial  gradients  of  alpha  mating  factor.  Proc  Natl  Acad  Sci  U  S  A  90(18):8332-­‐8336.  

6.   Colman-­‐Lerner  A,  et  al.  (2005)  Regulated  cell-­‐to-­‐cell  variation  in  a  cell-­‐fate  decision  system.  Nature  437(7059):699-­‐706.  

7.   Longtine  MS,  et  al.  (1998)  Additional  modules  for  versatile  and  economical  PCR-­‐based  gene  deletion  and  modification  in  Saccharomyces  cerevisiae.  Yeast  14(10):953-­‐961.  

8.   Shaner  NC,  et  al.  (2013)  A  bright  monomeric  green  fluorescent  protein  derived  from  Branchiostoma  lanceolatum.  Nature  methods  10(5):407-­‐409.  

9.   Bush  A,  Chernomoretz  A,  Yu  R,  Gordon  A,  &  Colman-­‐Lerner  A  (2012)  Using  Cell-­‐ID  1.4  with  R  for  microscope-­‐based  cytometry.  Current  protocols  in  molecular  biology  /  edited  by  Frederick  M.  Ausubel  ...  [et  al  Chapter  14:Unit  14  18.  

10.   Gordon  A,  et  al.  (2007)  Single-­‐cell  quantification  of  molecules  and  rates  using  open-­‐source  microscope-­‐based  cytometry.  Nat  Methods  4(2):175-­‐181.  

11.   Efron  B  &  Tibshirani  R  (1994)  An  introduction  to  the  bootstrap  (Chapman  &  Hall,  New  York)  pp  xvi,  436  p.  

 

Supporting InformationVentura et al. 10.1073/pnas.1322761111

Movie S1. Yeast cells polarize in minutes after pheromone stimulation (related to Fig. 3). We stimulated yeast expressing the MAPK scaffold protein Ste5fused to YFPx3 with 1 μM α-factor (isotropic stimulation) at time 0 and then followed cells by time-lapse confocal fluorescence microscopy at the time pointsindicated. We used Ste5-YFPx3 to visualize polarized patches in the membrane where the active G protein βγ dimer localizes (Ste4/Ste18) (Fig. 4A). After aninitial isotropic relocalization of cytoplasmic Ste5-YFPx3 to the plasma membrane, YFP accumulation to a single pole was evident in many cells starting at 2 min(white arrows indicate the first image with evident polarization).

Movie S1

Other Supporting Information Files

SI Appendix (PDF)

Ventura et al. www.pnas.org/cgi/content/short/1322761111 1 of 1