Abstract
Computer-assisted simulation is a promising approach for clarifying complicated signaling networks. However, this approach is currently limited by a deficiency of kinetic parameters determined in living cells. To overcome this problem, we applied fluorescence cross-correlation spectrometry (FCCS) to measure dissociation constant (Kd) values of signaling molecule complexes in living cells (in vivo Kd). Among the pairs of fluorescent molecules tested, that of monomerized enhanced green fluorescent protein (mEGFP) and HaloTag-tetramethylrhodamine was most suitable for the measurement of in vivo Kd by FCCS. Using this pair, we determined 22 in vivo Kd values of signaling molecule complexes comprising the epidermal growth factor receptor (EGFR)–Ras–extracellular signal-regulated kinase (ERK) mitogen-activated protein (MAP) kinase pathway. With these parameters, we developed a kinetic simulation model of the EGFR-Ras-ERK MAP kinase pathway and uncovered a potential role played by stoichiometry in Shc binding to EGFR during the peak activations of Ras, MEK, and ERK. Intriguingly, most of the in vivo Kd values determined in this study were higher than the in vitro Kd values reported previously, suggesting the significance of competitive bindings inside cells. These in vivo Kd values will provide a sound basis for the quantitative understanding of signal transduction.
INTRODUCTION
The epidermal growth factor (EGF) receptor (EGFR)–Ras–extracellular signal-regulated kinase (ERK) mitogen-activated protein (MAP) kinase pathway plays pleiotropic roles in cell functions such as cell proliferation, differentiation, survival, and tumorigenesis (1–3). This signaling pathway has been extensively studied, and vast amounts of proteins and regulations have been identified, resulting in an increase in the pathway's complexity. Computer-assisted simulation is one of the most promising approaches for the comprehensive understanding of the signal transduction pathway as a system. Indeed, a number of simulation models of the EGFR-Ras-ERK MAP kinase pathway have been reported over the past 10 years (4–7). In these simulation models, most of the kinetic parameters used for numerical simulations were not measured experimentally but rather were assumed by fitting the experimental data with the simulation data or simply determined arbitrarily. Consequently, there are substantial differences in the parameters among these studies, making it difficult to evaluate these simulation models quantitatively.
The kinetic parameters used for the simulation of intracellular signal transduction include protein concentrations, enzymatic kinetics, diffusion coefficients, and dissociation constants of the protein-protein interactions, which are denoted Kd. Among them, Kd is of central importance, because protein-protein interactions are a major constituent of signal transduction pathways (8). Under steady-state conditions, the Kd of the simple binding between protein A and protein B is defined as
| (1) |
| (2) |
| (3) |
where kf and kb are association and dissociation rate constants, respectively, and [Free A], [Free B], and [AB] correspond to the concentrations of unbound free protein A, protein B, and protein AB complexes, respectively. According to these definitions, the smaller the Kd values, the higher the affinity of the protein-protein interaction.
The Kd value has been determined by in vitro experiments such as coprecipitation experiments, sedimentation equilibrium using analytical ultracentrifugation, surface plasmon resonance (SPR), and isothermal titration calorimetry (ITC). All of these in vitro methods enable us to acquire the Kd value (here referred to as the “in vitro Kd” value) (Fig. 1A). The in vitro Kd value reflects the strength of the protein-protein interaction determined by the intrinsic properties of the two proteins. On the other hand, a few reports have measured dissociation constants in living cells (here referred to as the “in vivo Kd”) (Fig. 1B) by means of intermolecular fluorescence resonance energy transfer (FRET), single-molecule fluorescence imaging, and fluorescence cross-correlation spectroscopy (FCCS) (9–12). All of these techniques employ two-color fluorescence imaging. Theoretically, the in vivo Kd can be affected mainly by two factors: competitive binding and molecular crowding (Fig. 1B). In the former case, non-fluorescently labeled proteins, including endogenous and other interacting proteins, bind competitively to fluorescently labeled molecules and consequently appear to lead to an overestimation of the in vivo Kd values relative to the in vitro Kd values (Fig. 1B, top). The overestimated Kd is also known as the “apparent Kd.” Meanwhile, in the latter case, macromolecular crowding can occur because the cytosol contains numerous molecular species, such as proteins, lipids, and nucleotides, and so on, which occupy a substantial fraction of the volume within the cytoplasm (13). Such molecular crowding can profoundly influence protein-protein binding through an excluded-volume effect (14, 15) (Fig. 1B, bottom). The in vivo Kd confers a potential advantage to kinetic simulation models, because the in vivo Kd authentically includes the effects of all intracellular environments, such as competitive bindings and molecular crowding, on protein-protein interactions within a cell. However, to date, only a few in vivo Kd values have been made available for computer simulation for the EGFR-Ras-ERK MAP kinase pathway, possibly due to the technical difficulties.
FIG 1.

Strategy for measuring in vivo Kd by FCCS. (A and B) Comparison between in vitro Kd (A) and in vivo Kd (B) values. In general, the in vivo Kd was affected by competitive binding proteins (B, top) and molecular crowding (B, bottom), leading to increased and decreased of Kd values compared to the in vitro Kd values (A), respectively. (C) Experimental procedure for measurement of in vivo Kd by FCCS. HeLa cells were cotransfected with two plasmids expressing fluorescent protein-fused proteins A and B. Two days after transfection, the cells were subjected to FCCS measurements. At least 5 points in the cytoplasm of each cell were analyzed by FCCS. In each cell, the concentrations of total A protein ([Atotal]), total B protein ([Btotal]), and the protein AB complex ([AB]) were obtained by the y intercepts of the autocorrelation curves (red and green curves) and cross-correlation curve (blue curve), respectively. The averaged values of the ratios of [AB] to [Btotal] in each cell, with standard deviations, are plotted against the unbound A protein concentration ([Atotal] − [AB]), with standard deviations. The Kd value was obtained by fitting the data to equations 16 and 17.
FCCS allows the measurement of protein mobility, protein concentrations, and protein-protein interactions by exploiting the temporal fluorescence fluctuations of two diffusing fluorescently labeled particles under a confocal laser scanning microscope with a tiny focal volume, called the effective volume (16). As a distinct number of fluorescently labeled molecules diffuse through the effective volume (approximately 1 fl), the fluorescence signals fluctuate in a manner dependent on the mobility and concentration. An autocorrelation function of the fluctuating fluorescence signal provides the diffusion coefficient and concentration of molecules. FCCS utilizes two spectrally different fluorophores to label a pair of proteins. If the differently labeled particles are associated with each other, they pass through the effective volume in a synchronized way. Therefore, the simultaneous fluctuations of their fluorescence signals lead to an increase in the amplitude of the cross-correlation function. The amplitude provides the concentration of the protein-protein complex.
In this study, we established a method for obtaining in vivo Kd values in living cells by FCCS and determined >20 in vivo Kd values for the EGFR-Ras-ERK MAP kinase pathway in HeLa cells. In addition, we built a simulation model of the EGFR-Ras-ERK MAP kinase pathway based on the in vivo Kd values. This model suggested that multiple bindings of Shc to phosphorylated EGFR (pEGFR) are required for the peak activation of Ras, MEK, and ERK in response to EGF stimulation. Intriguingly, most of the in vivo Kd values measured in this study were higher than those measured previously by in vitro experiments, suggesting that competitive bindings play a major role in the in vivo Kd values within the cytoplasm.
MATERIALS AND METHODS
Plasmids.
The cDNA of HaloTag was amplified by PCR using the pFC14A (HaloTag 7) cytomegalovirus (CMV) Flexi vector (Promega, Madison, WI) as the template. The cDNA of monomerized enhanced green fluorescent protein (mEGFP), HaloTag, or the GGSGGS linker was inserted into the pCAGGS, pCXN2, pCX4puro, or pCX4neo vector (17, 18) to generate pCAGGS-HaloTag-GGSGGS-mEGFP, pCAGGS-mEGFP, pCAGGS-FLAG-mEGFP, pCXN2-mEGFP, pCX4puro-HaloTag, pCX4neo-mEGFP, pCAGGS-FLAG-HaloTag, and pCXN2-HaloTag. pDONR223-RPS6KA1 (RSK1), pDONR223-RPS6KA2 (RSK3), and pDONR223-KSR (KSR1) were gifts from William Hahn (19) (Addgene plasmids 23860, 23530, and 23443, respectively). The cDNAs of human RSK1 and human RSK3 were amplified by PCR. The cDNAs of human Shc1 (20), human Grb2 (21), mouse Sos1 (22), mouse p85, cow p110α (23), human HRasdCT-G12V (where dCT stands for deleted C terminus) (24), human BRaf, human BRaf-S364A (25), human CRaf, human CRaf-S259A (24), human RSK2 (26), human RSK1, and human RSK3 were subcloned into mEGFP or HaloTag vector to generate pCXN2-mEGFP-Shc1-stop, pCX4puro-HaloTag-Shc, pCAGGS-FLAG-Grb2-HaloTag, pCAGGS-mEGFP-Sos1, pCAGGS-FLAG-p85-HaloTag, pCAGGS-mEGFP-p110α, pCXN2-HaloTag-HRasdCT-G12V, pCAGGS-FLAG-BRaf-mEGFP, pCAGGS-FLAG-BRaf-S364A-mEGFP, pCAGGS-FLAG-CRaf-mEGFP, pCAGGS-FLAG-CRaf-S259A-mEGFP, pCXN2-HaloTag-RSK1, pCXN2-HaloTag-RSK2, and pCXN2-HaloTag-RSK3, respectively. Human MEK1, human MEK2, human ERK1, and human ERK2 were obtained by reverse transcription-PCR (RT-PCR) with a HeLa cDNA library and subcloned into the mEGFP or HaloTag vector to generate pCAGGS-FLAG-MEK1-HaloTag, pCAGGS-FLAG-MEK2-HaloTag, pCXN2-mEGFP-ERK1, and pCXN2-mEGFP-ERK2, respectively. The ccdB gene and chloramphenicol resistance genes sandwiched with aatR1 and aatR2 were amplified by PCR and inserted into the pCAGGS-mEGFP vector. pCAGGS-mEGFP-KSR1 was constructed with pCAGGS-mEGFP-ccdB and pDONR223-KSR by using the Gateway vector conversion system according to the manufacturer's instructions (Invitrogen, San Diego, CA). The cDNA of human EGFR was subcloned into the pPBbsr2-mEGFP vector to generate pPBbsr2-EGFR-mEGFP (27).
Cells, reagents, and antibodies.
HeLa cells were purchased from the Human Science Research Resources Bank (Sennanshi, Japan) and maintained in Dulbecco's modified Eagle medium (DMEM) (Sigma-Aldrich, St. Louis, MO) supplemented with 10% fetal bovine serum (FBS). HeLa cells stably expressing mEGFP, HaloTag, or both HaloTag-Shc and EGFR-mEGFP were established in accordance with the conventional retroviral gene transfer and piggyBac transposon systems (17, 28). EGF and sodium orthovanadate (Na3VO4) were purchased from Sigma-Aldrich. Tetradecanoyl phorbol acetate (TPA) was purchased from Calbiochem (La Jolla, CA). HaloTag-TMR (tetramethylrhodamine) ligand and anti-HaloTag antibody (G9281) were purchased from Promega. The following antibodies were also used in this study: anti-Grb2 (catalog number sc-255), anti-RSK1 (catalog number sc-231), anti-RSK2 (catalog number sc-9986), and anti-RSK3 (catalog number sc-1431) antibodies (purchased from Santa Cruz Biotechnology, Santa Cruz, CA); anti-Shc (catalog number 610879), anti-MEK1 (catalog number 610121), and anti-MEK2 (catalog number 610235) antibodies (BD Transduction Laboratories, San Jose, CA); anti-green fluorescent protein (GFP) (catalog number 632375) antibody (Clontech, Palo Alto, CA); and anti-p44/p42 MAP kinase (ERK1/2) antibody (catalog number 4695; Cell Signaling Technology, Danvers, MA). PD-184352 was purchased from Toronto Research Chemicals (Toronto, Canada).
FCCS measurements.
HeLa cells were plated onto 35-mm glass-base dishes (Asahi Techno Glass, Tokyo, Japan). One day after plating, plasmids were transfected by using 293fectin transfection reagent according to the manufacturer's instructions (Invitrogen, San Diego, CA). Two or three days after transfection, the cells were labeled with HaloTag-TMR ligand, as described below.
FCCS measurements were performed with the LSM780 Meta/ConfoCor 2 system (Carl Zeiss, Oberkochen, Germany) and the FV-1000 confocal imaging system (Olympus, Tokyo, Japan), equipped with gallium arsenide phosphide (GaAsP) detectors. With respect to the Zeiss system, the excitation lines were set at 488 nm and 561 nm. The excitation laser power under our microscopic settings was determined in accordance with a previous report (29). The excitation beam was reflected by an MBS 488/561 dichroic mirror and focused by a water immersion objective lens (C-Apochromat 40×/1.2 W Corr M27; Carl Zeiss). The emitted light was collimated and then split by an NFT 565 dichroic mirror. Emission signals were detected through a BP 505-540 infrared (IR) emission filter for mEGFP and a BP 615-680 IR emission filter for HaloTag-TMR. With respect to the Olympus system, the excitation lines were set at 488 nm and 559 nm. The excitation beam was reflected by a DM 405/488/559 dichroic mirror and focused by an oil immersion objective lens (60×/1.35-numerical-aperture [NA] Uplsapo 60XO; Olympus). The emitted light was detected through a diffraction grating to measure the fluorescence at wavelengths of 495 to 540 nm for mEGFP and 575 to 630 nm for HaloTag-TMR. FCCS measurements were performed for 28 s (Zeiss) and 17 s (Olympus) for each point and repeated at 7 different points per cell. At least 10 cells were examined under all conditions. All FCCS experiments were performed at room temperature.
Spectroscopy by confocal microscopy.
HeLa cells were transfected with a plasmid encoding mEGFP- or HaloTag-fused proteins. Two days after transfection, the cells were incubated for 15 min with 5 μM HaloTag-TMR ligand at 37°C, washed twice with phosphate-buffered saline (PBS), and incubated for 30 min in DMEM lacking all vitamins (DMEM-V) (30) at 37°C. Before imaging, cells were washed twice and fed with DMEM-V. Fluorescence spectra were acquired by using an FV-1000 confocal imaging system (Olympus, Tokyo, Japan). The excitation line was set at 488 nm or 559 nm. The excitation beam was reflected by a DM 405/488/559 dichroic mirror and focused by an oil immersion objective lens (Uplsapo 60XO, 1.35 NA; Olympus). Emission signals were detected in the lambda scanning mode.
Calculation of the effective detection volume.
The effective detection volume of the confocal laser scanning microscope, Veff, is given by the following equation:
| (4) |
where wxy and wz are the lateral and axial 1/e2 distances, respectively. These are calculated as
| (5) |
| (6) |
where D and t indicate the diffusion coefficient and the average time for detected molecules to diffuse out of this volume, respectively, and k is a structure parameter defined as k = wz/wxy. The D value of rhodamine 6G in water at room temperature was determined previously to be 280 μm2/s (31). The t and k values were obtained by fluorescence correlation spectroscopy (FCS) measurements with rhodamine 6G in water at room temperature and fitting of the autocorrelation curve using ZEN software (Carl Zeiss). Thus, wxy and wz were calculated to be 0.24 μm and 1.81 μm, respectively, and Veff was determined to be 0.58 ± 0.20 fl under our experimental conditions.
Data analysis for FCCS.
Fluorescent signals were analyzed as described previously (29, 32). The fluorescence autocorrelation function Gauto(τ), from the mEGFP and HaloTag-TMR channels, GG(τ) and GH(τ), respectively, and the fluorescence cross-correlation function, Gcross(τ), were calculated according to the normalized correlation function:
| (7) |
where τ indicates the time delay; Ii is the fluorescence intensity of the mEGFP channel (i = G) or the HaloTag-TMR channel (i = H); the angle brackets denote the time average; and GG(τ), GH(τ), and Gcross(τ) denote the autocorrelation functions of mEGFP (i = j = G), HaloTag-TMR (i = j = H), and the cross-correlation function (i = G or H; j = H or G), respectively. The calculated G(τ) values were fitted to the following equation:
| (8) |
where τD, k, G(0), and G(∞) indicate the correlation time, the structural parameter, the amplitude of the correlation curve, and the G(τ) value when a correlation functions under steady state, respectively. For fitting, we used the Excel Solver tool to obtain τD, G(0), and G(∞). The concentrations of mEGFP ([EGFPtotal]), HaloTag-TMR ([HaloTagtotal]), and the complex ([Complex]) are given as
| (9) |
| (10) |
| (11) |
where NA corresponds to the Avogadro number and Veff is the effective detection volume, determined as described above. The value obtained by dividing the cross-correlation amplitude by one of the autocorrelation amplitudes, called the relative cross-correlation (RCC), provides a direct measure of the fraction of molecular binding, as follows (16):
| (12) |
| (13) |
where RCCG and RCCH are the relative cross-correlations of EGFP and HaloTag, respectively. These RCC values were then corrected by the RCC values obtained from the positive control (RCCPC_G and RCCPC_H), in which two fluorescent proteins were linked by a short peptide, and the negative control (RCCNC_G and RCCNC_H), in which two fluorescent proteins were expressed separately and would not bind each other. Thus, the corrected complex concentration, [cComplex], was given as
| (14) |
| (15) |
The values of RCCPC_G, RCCPC_H, RCCNC_G, and RCCNC_H were measured in every experiment and used to obtain the [cComplex] value.
Calculation of in vivo Kd.
According to the definition of Kd, the fraction of the complex in total EGFP or total HaloTag proteins is derived from equations 2, 14, and 15:
| (16) |
| (17) |
The fractions of the bound protein in total EGFP and total HaloTag, i.e., the left-hand terms of equations 16 and 17, were plotted as a function of free (unbound) HaloTag and EGFP, respectively. The in vivo Kd value was obtained by nonlinear fitting of the experimental data with equations 16 and 17 and by averaging of these fitted values. For fitting, we used the Excel Solver function.
Quantification of concentrations of endogenous proteins in HeLa cells.
The protein concentrations of endogenous Shc1, Grb2, MEK1, MEK2, ERK1, ERK2, RSK1, RSK2, and RSK3 in HeLa cells were determined as previously described (23). The 5myc-Flag-3HA-tagged (where HA stands for hemagglutinin) yellow fluorescent protein (YFP) and glutathione S-transferase (GST)-tagged HaloTag were purified from Escherichia coli. The lysates of the cells expressing mEGFP-Shc1, Grb2-mEGFP, MEK1-PAGFP, MEK2-m1YFP, mEGFP-ERK2, HaloTag-RSK1, HaloTag-RSK2, and HaloTag-RSK3 were used as references. The standard protein (10 ng, 5 ng, and 2.5 ng of 5myc-Flag-3HA-tagged YFP or GST-tagged HaloTag), the reference protein (a serial dilution of the reference lysate), and the total cell lysate obtained from 5.0 × 104, 2.5 × 104, or 1.25 × 104 cells were separated by SDS-PAGE, followed by immunoblotting with anti-GFP or anti-HaloTag antibodies (standard and reference) or specific antibodies (reference and total cell lysate). The bound antibodies were detected and quantified with an Odyssey infrared imaging system (Li-Cor). The amount of reference protein was determined by the relative intensity of the reference to 5myc-Flag-3HA-tagged YFP or GST-tagged HaloTag, and the amount of endogenous protein was calculated from the intensity of the reference protein. Finally, the concentration of endogenous protein in an individual HeLa cell was determined by dividing the amount of endogenous protein by the volume of a HeLa cell, which was 3.4 × 10−12 liters (9).
Kinetic modeling and numerical simulation.
All kinetic reactions were described with mass action kinetics with CellDesigner (version 4.1) (33, 34) (see Fig. 6 and Tables 3 and 4.). Most of the enzymatic reactions were described as first-order reactions for simplicity. The reactions that follow the first-order enzymatic kinetics in our model (see Fig. 6) are as follows: Shc phosphorylation by EGFR (reaction 4), pShc dephosphorylation (reaction 5), Ras activation (reactions 11 and 12), Ras inactivation (reaction 13), Shoc2 activation by EGFR and inactivation (reaction 14), MEK phosphorylation by Raf (reaction 19), MEK dephosphorylation (reaction 20), and ERK dephosphorylation (reactions 26 to 29 and 38 to 41). Among them, MEK dephosphorylation and ERK dephosphorylation were measured in our previous study (9, 22). The other reactions were approximated as first-order reactions for simplicity. The ordinary differential equations and parameters were exported to MATLAB software (version R2008b or R2010b; Mathworks Inc., Natick, MA) through Systems Biology Workbench (version 2.7.8) (35). Numerical simulation was performed by using the MATLAB function ode23, a numerical solver of differential equations for solving nonstiff problems. The simulation data calculated by ode23 did not differ from those obtained by ode15s, a numerical solver for stiff differential equations (data not shown). The concentrations of each isoform were summed to obtain the total concentrations. Dissociation constants of the summed proteins were weighted and summed as follows:
| (18) |
where Ai and Bj correspond to the concentration of isoform i (i = 1, 2, …, m) of protein A and the concentration of isoform j (j = 1, 2, …, n) of protein B, respectively, and Kd ij is the dissociation constant of the binding of Ai and Bj.
FIG 6.
Kinetic simulation model of the EGFR-Ras-ERK MAP kinase pathway. The schematic of the EGFR-Ras-ERK pathway, which was created with CellDesigner (33), includes four modules: the EGFR-Shc-Grb2-Sos module (1), the Ras-Raf activation module (2), the MEK-ERK processive phosphorylation module (3), and the ERK-Sos negative-feedback module (4). The reaction identification (Re) number is indicated for each reaction. For more details, see Tables 3 and 4.
TABLE 3.
Initial concentrations of signaling molecules in simulationsa
| Protein | Concn (μM) | Description or reference |
|---|---|---|
| EGF | 0.016 | 10 ng/ml |
| EGFR | 0.42 | 22 |
| Shoc2 | 0.67 | 72 |
| Shc | 0.15 | This study |
| Grb2 | 0.14 | This study |
| Sos1 | 0.12 | 22 |
| Ras | 0.43 | 51 |
| BRaf | 0.037 | This study |
| CRaf | 0.013 | 51 |
| MEK | 0.87 | This study |
| ERK | 0.79 | This study |
| RSK | 0.28 | This study |
| RasGEF | 1.0 | 22 |
| RasGAP | 1.0 | 22 |
The initial concentrations of species in the simulation model are listed. The concentrations of isoforms, e.g., MEK1 and MEK2, are summed simply.
TABLE 4.
Kinetic reactions and parametersf
| Reaction identification | Reaction and equation | Parameter |
Description or reference | ||
|---|---|---|---|---|---|
| Abbreviation | Value | Unit of measure | |||
| 1 | EGF + EGFR ⇔ pEGFR | kf | 100 | μM/s | Estimated from Fig. 7 |
| kf · EGF · EGFR − kb · pEGFR | kb | 0.15 | s | Estimated from Fig. 7 | |
| 2 | EGF_EGFR → pEGFR_degra | kf | 0.002 | s | Estimated from Fig. 7 |
| kf · pEGFR | |||||
| 3 | pEGFR_degra → EGFR | kf | 0.001 | s | Estimated from Fig. 7 |
| kf · pEGFR_degra | |||||
| 4 | Shc + pEGFR → pShc | kf | 0.031 | μM/s | Quantified in Fig. 7a |
| kf · pEGFR · Shc | |||||
| 5 | pShc → Shc | kf | 0.028 | s | Quantified in Fig. 7a |
| kf · pShc | |||||
| 6 | pShc + Grb2 ⇔ pShc_Grb2 | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · pShc · Grb2 − kb · pShc_Grb2 | kb | 0.14 | s | ||
| 7 | Grb2 + Sos1 ⇔ Grb2_Sos1 | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · Grb2 · Sos1 − kb · Grb2_Sos1 | kb | 0.17 | s | ||
| 8 | pShc_Grb2 + Sos1 ⇔ pShc_Grb2_Sos1 | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · pShc_Grb2 · Sos1 − kb · pShc_Grb2_Sos1 | kb | 0.17 | s | ||
| 9 | pShc + Grb2_Sos1 ⇔ pShc_Grb2_Sos1 | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · pShc · Grb2_Sos1 − kb · pShc_Grb2_Sos1 | kb | 0.14 | s | ||
| 10 | pEGFR + n · pShc_Grb2_Sos1 ⇔ pEGFR_(pShc_Grb2_Sos1)n | kf | 0.1 | μM/s | See text |
| (pEGFR) kf · pEGFR · pShc_Grb2_Sos1 − kb · pEGFR_(pShc_Grb2_Sos1)n | kb | (0.1c, 0.0056d, 0.02e) | s | ||
| [pShc_Grb2_Sos1) n · (kf · pEGFR · pShc_Grb2_Sos1 − kb · pEGFR_(pShc_Grb2_Sos1)n] | n | (1c,d, 5.9e) | |||
| 11 | RasGDP + pEGFR_(pShc_Grb2_Sos1)n (enzyme) → RasGTP | kf | 1 | μM/s | 73 |
| kf · n · pEGFR_(pShc_Grb2_Sos1)n · RasGDP | n | (1c,d, 5.9e) | See text | ||
| 12 | RasGDP + RasGEF → RasGTP | kf | 0.00005 | μM/s | 22 |
| kf · RasGEF · RasGDP | |||||
| 13 | RasGTP + RasGAP → RasGDP | kf | 0.05 | μM/s | 22 |
| kf · RasGAP · RasGTP | |||||
| 14 | Inactive_Shoc2 + pEGFR (enzyme) ⇔ Active_Shoc2 | kf | 0.1 | μM/s | 52 |
| kf · inactive_Shoc2 · pEGFR − kb · active_Shoc2 | kb | 0.1 | s | ||
| 15 | RasGTP + cRaf ⇔ Ras_cRaf | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · RasGTP · cRaf − kb · Ras_cRaf | kb | 0.23 | s | ||
| 16 | RasGTP + BRaf ⇔ Ras_BRaf | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · RasGTP · BRaf − kb · Ras_BRaf | kb | 0.31 | s | ||
| 17 | RasGTP + active_Shoc2 ⇔ Ras_Shoc2 | kf | 1 | μM/s | 52 |
| kf · RasGTP · active_Shoc2 − kb · Ras_Shoc2 | kb | 0.1 | s | ||
| 18 | Ras_Shoc2 + cRaf ⇔ Ras_Shoc2_cRaf | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · Ras_Shoc2 · cRaf − kb · Ras_Shoc2_cRaf | kb | 0.23 | s | ||
| 19 | MEK_cyt + (Ras_cRaf + Ras_BRaf + Ras_Shoc2_cRaf) → pMEK_cyt | kf | (1c, 10d,e) | μM/s | See text |
| kf · MEK_cyt · (Ras_cRaf + Ras_BRaf + Ras_Shoc2_cRaf) | |||||
| 20 | pMEK_cyt → MEK_cyt | kf | 0.0096 | s | 9 |
| kf · pMEK_cyt | |||||
| 21 | MEK_cyt + ERK_cyt ⇔ MEK_ERK_cyt | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · MEK_cyt · ERK_cyt − kb · MEK_ERK_cyt | kb | 0.66 | s | ||
| 22 | ERK_cyt + RSK_cyt ⇔ ERK_RSK_cyt | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · ERK_cyt · RSK_cyt − kb · ERK_RSK_cyt | kb | 0.12 | s | ||
| 23 | pMEK_cyt + ERK_cyt ⇔ pMEK_ERK_cyt | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · pMEK_cyt · ERK_cyt − kb · pMEK_ERK_cyt | kb | 0.66 | s | ||
| 24 | pMEK_ERK_cyt → pMEK_pYERK_cyt | kf | (0.073c, 0.32d,e) | s | See text |
| kf · pMEK_ERK_cyt | |||||
| 25 | pMEK_pYERK_cyt → pTpYERK_cyt + pMEK_cyt | kf | 0.05 | μM/s | 9 |
| kf · pMEK_pYERK_cyt | |||||
| 26 | pTpYERK_cyt → pYERK_cyt | kf | 0.004 | μM/s | 9 |
| kf · pTpYERK_cyt | |||||
| 27 | pTpYERK_cyt → pTERK_cyt | kf | 0.0055 | μM/s | 9 |
| kf · pTpYERK_cyt | |||||
| 28 | pYERK_cyt → ERK_cyt | kf | 0.0067 | μM/s | 9 |
| kf · pYERK_cyt | |||||
| 29 | pTERK_cyt → ERK_cyt | kf | 0.0068 | μM/s | 9 |
| kf · pTERK_cyt | |||||
| 30 | pYERK_cyt + pMEK_cyt → pTpYERK_cyt | kf | 0.021 | μM/s | 9 |
| kf · pYERK_cyt · pMEK_cyt | |||||
| 31 | pTERK_cyt + pMEK_cyt → pTpYERK_cyt | kf | 0.02 | μM/s | 9 |
| kf · pYERK_cyt · pMEK_cyt | |||||
| 32 | ERK_cyt ⇔ ERK_nuc | kf | 0.0017 | s | 9 |
| kf · ERK_cyt − kb · ERK_nuc | kb | 0.013 | s | ||
| 33 | pYERK_cyt ⇔ pYERK_nuc | kf | 0.0025 | s | 9 |
| kf · pYERK_cyt − kb · pYERK_nuc | kb | 0.017 | s | ||
| 34 | pTERK_cyt ⇔ pTERK_nuc | kf | 0.0022 | s | 9 |
| kf · pTERK_cyt − kb · pTERK_nuc | kb | 0.049 | s | ||
| 35 | pTpYERK_cyt ⇔ pTpYERK_nuc | kf | 0.0082 | s | 9 |
| kf · pTpYERK_cyt − kb · pTpYERK_nuc | kb | 0.0076 | s | ||
| 36 | RSK_cyt ⇔ RSK_nuc | kf | 0.001 | s | Quantified in Fig. 7F and G |
| kf · RSK_cyt − kb · RSK_nuc | kb | 0.11 | s | Data not shown | |
| 37 | ERK_nuc + RSK_nuc ⇔ ERK_RSK_nuc | kf | 0.1 | μM/s | Calculated from in vivo Kd valueb |
| kf · ERK_nuc · RSK_nuc − kb · ERK_RSK_nuc | kb | 0.12 | s | ||
| 38 | pTpYERK_nuc → pYERK_nuc | kf | 0.0032 | s | 9 |
| kf · pTpYERK_nuc | |||||
| 39 | pTpYERK_nuc → pTERK_nuc | kf | 0.0038 | s | 9 |
| kf · pTpYERK_nuc | |||||
| 40 | pYERK_ nuc → ERK_nuc | kf | 0.0077 | s | 9 |
| kf · pYERK_nuc | |||||
| 41 | pTERK_ nuc → ERK_nuc | kf | 0.0058 | s | 9 |
| kf · pTERK_nuc | |||||
| 42 | Sos1 + pTpYERK_cyt ⇔ pSos1 | kcat | 0.02 | μM/s | 22 |
| kcat · pTpYERK_cyt · Sos1/(Km + Sos1) − kb · pSos1 | Km | 1 | M | ||
| kb | 0.0025 | s | |||
| 43 | pSos1 + pTpYERK_cyt ⇔ ppSos1 | kcat | 0.02 | μM/s | 22 |
| kcat · pTpYERK_cyt · pSos1/(Km + pSos1) − kb · ppSos1 | Km | 1 | μM | ||
| kb | 0.0025 | s | |||
| 44 | ppSos1 + pTpYERK_cyt ⇔ pppSos1 | kcat | 0.02 | μM/s | 22 |
| kcat · pTpYERK_cyt · ppSos1/(Km + ppSos1) − kb · pppSos1 | Km | 1 | M | ||
| kb | 0.0025 | s | |||
| 45 | pppSos1 + pTpYERK_cyt ⇔ ppppSos1 | kcat | 0.02 | μM/s | 22 |
| kcat · pTpYERK_cyt · pppSos1/(km + pppSos1) − kb · ppppSos1 | Km | 1 | μM | ||
| kb | 0.0025 | s | |||
The enzymatic reaction was treated as a first-order enzymatic reaction.
The kf value was set to 0.1 mM/s. The kb value was calculated from the kf and in vivo Kd values obtained in this study (Table 1).
The value is used in Fig. 8A.
The value is used in Fig. 8D.
The value is used in Fig. 8E.
Reaction identification numbers correspond to the numbers in Fig. 6. Boldface type denotes enzymes.
RESULTS
Strategy for measuring in vivo Kd.
In this study, we attempted to measure the in vivo Kd between signaling proteins in the EGFR-Ras-ERK MAP kinase pathway with FCCS and to build a numerical model based on the experimentally verified parameters. Figure 1C provides a schematic of the procedure for measuring the in vivo Kd by FCCS. HeLa cells were cotransfected with two plasmids expressing chimeric proteins A and B, which were fused with different fluorescent proteins. Two or three days after transfection, we performed FCCS measurements at seven points in the cytoplasm of a single HeLa cell and obtained the concentrations of the fluorescently labeled molecules, [Atotal] and [Btotal], and their complex, [AB], from y intercepts, G(0), of the auto- and cross-correlation curves, respectively (Fig. 1C). This analysis was repeated with at least 10 cells, and the fractions of [AB] to [Btotal] were plotted as a function of [Atotal] − [AB]. These plots were fitted to equations 16 and 17 to obtain the in vivo Kd value. It should be noted that this measurement requires the assumption of steady-state conditions, under which the binding reactions are in equilibrium kinetics.
Optimization of FCCS in HeLa cells.
First, we examined which pairs of fluorescent proteins were most suited for the application of FCS/FCCS to living cells. In line with previous studies using GFP and red fluorescent protein (RFP) as a pair of fluorescent proteins for FCCS measurements (12, 36, 37), we tested the brightness, the degree of bleaching, and triplet formation. As reported previously, mEGFP exhibited sufficiently bright fluorescence, a small triplet fraction, and resistance to photobleaching. In contrast, substantial triplet fractions were detected in tdTomato and mCherry, as reported previously (38). Furthermore, Foo et al. reported previously that the low maturation efficiency of mCherry affected FCCS analysis (38).
To overcome the drawbacks of red fluorescent proteins in FCCS measurements, we took advantage of HaloTag, which is a modified haloalkane dehalogenase designed to covalently bind to synthetic ligands (HaloTag ligands) (39). The complex formed by conjugation of the HaloTag ligand to the membrane-permeable HaloTag-tetramethylrhodamine (TMR) ligand, here referred to as HaloTag-TMR, did not show detectable photobleaching or a detectable triplet fraction under these imaging conditions, indicating that this complex would have practical advantages for FCCS measurements. We acquired the fluorescence spectra of mEGFP and HaloTag-TMR in HeLa cells (Fig. 2A) and selected completely separable emission filters for mEGFP (505 to 540 nm) and HaloTag-TMR (615 to 680 nm).
FIG 2.

Experimental conditions for in vivo Kd measurement by FCCS. (A) Relative fluorescence emission spectra of mEGFP and HaloTag-TMR represented as a function of wavelength. Lines of 488 nm and 559 nm represent the excitation wavelengths for mEGFP and HaloTag-TMR, respectively. (B and C) HeLa cells expressing a fusion protein comprised of mEGFP, the GGSGGS linker, and HaloTag-TMR were subjected to FCCS measurement as a positive control (B). As a negative control (C), HeLa cells expressing individual mEGFP and HaloTag-TMR proteins were measured by FCCS. The autocorrelation curves for mEGFP and HaloTag-TMR and cross-correlation curves for the complex of the positive control and negative control are plotted as a function of the delay time, τ. Narrow and bold lines represent raw and fitted correlation curves, respectively. (D) The relative cross-correlation values for positive controls and negative controls are regarded as the values in the case of 100% binding and 0% binding, respectively. These values were measured at every experiment and used as reference values for correcting the fraction of the complex.
We next performed FCCS by using a positive control and a negative control (Fig. 2B and C). A chimeric protein consisting of mEGFP, the GGSGGS linker, and HaloTag-TMR was used as a positive control. mEGFP and HaloTag expressed separately in cells were used as a negative control. We measured the autocorrelation functions of mEGFP (Fig. 2B and C, green lines) and HaloTag-TMR (red lines) and the cross-correlation function (blue lines). Ideally, with the positive-control sample, the y intercept of the autocorrelation function of GFP, GG(0), must be identical to that of HaloTag-TMR, GH(0), and also identical to the y intercept of the cross-correlation function, Gcross(0). In the case of the negative-control sample, Gcross(0) must be equal to 0. Thus, the relative cross-correlation (RCC) values, which represented the fraction of the complex, theoretically must be 1 and 0 for the positive and negative controls, respectively. In our FCCS setup, the RCC values were within the range of 0.7 to 1.0 and 0.02 to 0.15 for the positive and negative controls, respectively (Fig. 2D). This deviation could be attributed to the imperfect overlap of the effective volumes of the two excitation lasers (38) and the spectral bleedthrough of EGFP fluorescence into the HaloTag-TMR emission channel. Therefore, we corrected the fraction of bound protein by using the RCC values of the positive and negative controls as 100% binding and 0% binding on each day of imaging, respectively (for more details, see Materials and Methods).
Measurement of in vivo Kd values of protein complexes in the EGFR-Ras-ERK MAP kinase pathway.
By using the method described above, we quantified the in vivo Kd values of protein complexes that transmit signals in the EGFR-Ras-ERK MAP kinase pathway in living HeLa cells (Fig. 3 and see Table 2).
FIG 3.
Measurement of in vivo Kd values for bindings of signaling molecules in the EGFR-Ras-ERK pathway. The measurements of the in vivo Kd values of the binding between Shc1-Grb2 (A and B), Grb2-Sos1 (C and D), HRasdCT-G12V–CRaf WT (E), HRasdCT-G12V–BRaf WT (F), MEK2-ERK1 (G), and ERK1-RSK2 (H) are shown. HeLa cells were transfected with plasmids expressing the indicated fusion proteins. Two or three days after transfection, the cells were treated with the TMR ligand, followed by serum starvation. The cells were then subjected to FCCS measurements. The fraction of the complex compared to the total HaloTag-TMR-fused proteins was plotted as a function of unbound mEGFP-fused proteins. EGF with Na3VO4 (B and D) was treated 30 min before FCCS measurements were started.
TABLE 2.
Concentrations of signaling molecules in the EGFR-Ras-ERK pathway in a HeLa cella
| Protein | No. of molecules/cell | Concn (μM) | Reference |
|---|---|---|---|
| EGFR | 0.42 | 22 | |
| p46 Shc | 1.4 × 105 | 0.068 | This study |
| p52 Shc | 1.7 × 105 | 0.083 | This study |
| Grb2 | 2.9 × 105 | 0.14 | This study |
| Sos1 | 0.12 | 22 | |
| Ras | 0.43 | 51 | |
| BRaf | 7.5 × 104 | 0.037 | This study |
| CRaf | 0.013 | 51 | |
| MEK1 | 2.6 × 105 | 0.13 | This study |
| MEK2 | 1.5 × 106 | 0.74 | This study |
| ERK1 | 2.4 × 105 | 0.12 | This study |
| ERK2 | 1.4 × 106 | 0.68 | This study |
| RSK1 | 1.5 × 105 | 0.073 | This study |
| RSK2 | 3.1 × 105 | 0.15 | This study |
| RSK3 | 1.1 × 105 | 0.054 | This study |
Protein concentrations in a single HeLa cell were determined by dividing the number of molecules by a cell volume of 3.4 pl in a HeLa cell (9).
The adaptor protein Grb2 binds to another adaptor protein, Shc1, in a phosphotyrosine-dependent manner (40). In agreement with this property, the in vivo Kd value of the Grb2-Shc1 complex was >15 μM without growth factor stimulation (Fig. 3A). To maximize the tyrosine phosphorylation of Shc1, HeLa cells were stimulated with EGF in the presence of Na3VO4, a tyrosine phosphatase inhibitor. Under these conditions, the Kd value of the Grb2-Shc1 complex was 1.4 μM (Fig. 3B and Table 1). The in vivo Kd value was slightly higher than or comparable to those obtained in vitro (40, 41). Next, the Kd value of the Grb2-Sos1 complex mediated by the SH3 domain was determined to be approximately 1.7 μM (Fig. 3C and Table 1). The Kd was increased to 6.5 μM in EGF- and Na3VO3-treated cells (Fig. 3D and Table 1). This increase in the Kd was cancelled by the MEK inhibitor (Table 1). These data were consistent with the negative-feedback mechanism showing that Sos1 phosphorylation by ERK reduced the affinity of binding between Sos1 and Grb2 (42–44).
TABLE 1.
Comparison of in vivo Kd values obtained in this study with in vitro Kd values measured previously by in vitro experimentsa
| Binding | This study |
Previous studies |
||||
|---|---|---|---|---|---|---|
| In vivo Kd (μM) | 95% CIb | No. of measured cells | Mean in vitro Kd (μM) ± SD | Methodc | Reference | |
| Grb2-pShc | 0.23 ± 55 | SPR | 61 | |||
| 1.4d | 0.7, 2.1 | 30 | 0.031 ± 8 | SPR | 61 | |
| 0.023 | SPR | 41 | ||||
| Grb2-Sos1 | 1.7 | 1.2, 2.2 | 23 | 0.001 | Co-IP | 62 |
| 6.5d | 3.3, 9.7 | 21 | 0.4 | ITC | 63 | |
| 2.8e | 1.7, 3.9 | 18 | 21.4 ± 5.9 | ITC | 64 | |
| 2.1f | 1.2, 1.5 | 16 | 1.7 ± 0.1 | SPR | 64 | |
| p110α-p85 | <0.1 | 24 | NR | |||
| HRas-CRafg | 6.9 | 4.2, 9.6 | 34 | 0.12 | ITC | 65 |
| 2.3h | 1.7, 2.9 | 27 | 0.13 | GDI assay | 66 | |
| 0.0021 | SPR | 67 | ||||
| HRas-BRaff | 3.1 | 2.5, 3.7 | 33 | 0.0011 | SPR | 67 |
| 3.5i | 2.7, 4.3 | 28 | ||||
| MEK1-ERK1 | 6.6 | 3.1, 10.1 | 22 | 0.058 | Stopped flow | 68 |
| MEK1-ERK2 | 0.046–0.476 | Michaelis constant | 69 | |||
| 11 | 7.2, 15 | 24 | 0.34 ± 0.06 | Michaelis constant | 70 | |
| 1.5 | SPR | 9 | ||||
| MEK2-ERK1 | 8.1 | 5.8, 10 | 24 | NR | ||
| MEK2-ERK2 | 5.6 | 3.8, 7.3 | 26 | NR | ||
| ERK1-RSK1 | 1.7 | 0.85, 2.5 | 22 | NR | ||
| ERK2-RSK1 | 0.87 | 0.54, 1.2 | 25 | 0.15 ± 0.01 | SPR | 71 |
| ERK1-RSK2 | 2.1 | 1.5, 2.8 | 36 | NR | ||
| ERK2-RSK2 | 1.3 | 0.93, 1.7 | 36 | NR | ||
| ERK1-RSK3 | 1.0 | 0.51, 1.4 | 23 | NR | ||
| ERK2-RSK3 | 0.7 | 0.47, 0.91 | 27 | NR | ||
| MEK1-KSR1 | 1.5 | 0.99, 2.0 | 25 | NR | ||
| MEK1-KSR2 | 1.2 | 0.79, 1.6 | 23 | NR | ||
| MEK2-KSR1 | 1.3 | 0.77, 1.9 | 35 | NR | ||
| MEK2-KSR2 | 1.1 | 0.54, 1.6 | 22 | NR | ||
The in vivo Kd values quantified by FCCS in this study were compared with in vitro Kd values that were previously determined by in vitro experiments. NR, not reported.
CI, confidence interval.
SPR, surface plasmon resonance; co-IP, coimmunoprecipitation; ITC, isothermal calorimetry; GDI, guanine nucleotide dissociation inhibitor.
In the presence of 10 ng/ml EGF and Na3VO4.
In the presence of 1 mM TPA.
In the presence of 10 μM PD-184352 (MEK inhibitor), 10 ng/ml EGF, and 10 mM Na3VO4.
Carboxyl-terminus-deleted cytoplasmic HRas-G12V mutant.
CRaf-S259A mutant.
BRaf-S364A mutant.
HRas localizes at the plasma membrane and Golgi apparatus via the lipid-modified C terminus (45, 46). We attempted to measure the binding between HRas-G12V, which was a constitutively active mutant, and CRaf. However, because of the slow diffusion of HRas at the plasma membrane, we failed to measure the in vivo Kd value for the binding of HRas-G12V to CRaf. To overcome this problem, we deleted the carboxyl terminus of HRas and prepared the cytoplasmic HRasdCT-G12V mutant (24). The Kd value of the HRasdCT-G12V–CRaf complex was 6.9 μM (Fig. 3E). Interestingly, the CRaf-S259A mutant, which adopts an open active conformation (24), bound to HRasdCT-G12V more strongly than did the CRaf wild type (WT), with an in vivo Kd of 2.3 μM (Table 1). In contrast, an equivalent mutation in BRaf, S364A, did not decrease the Kd value of the complex formed by HRasdCT-G12V and BRaf (Fig. 3F and Table 1), suggesting that the phosphorylation of BRaf at S364 plays a different role in Ras binding than the phosphorylation of CRaf at S259.
MEK has been shown to bind to and sequester ERK in the cytoplasm (47). ERK also binds to RSK through the D domain of the RSK LXL (DEJL) motif (48–50). There were also isoforms of them, MEK1, MEK2, ERK1, ERK2, RSK1, RSK2, and RSK3. We measured 10 possible combinations of the bindings among them (Fig. 3G and H and Table 1). We did not find any remarkable difference among the isoforms; therefore, the isoforms of MEK, ERK, and RSK were handled as single proteins in the simulation model described below.
The phosphatidylinositol 3-kinase (PI3-K) consists of a p85 regulatory subunit and a p110 catalytic subunit. The Kd value of the PI3-K complex was <0.1 μM (Table 1). We could not precisely determine the Kd value by nonlinear curve fitting, because the lower detection limit of the concentration for mEGFP and HaloTag-TMR was ca. 0.1 μM under our experimental conditions.
Evaluation of the effects of competitive bindings on in vivo Kd values by computer simulation.
Most of the in vivo Kd values obtained in this study were higher than the in vitro Kd values reported previously by an order of 1 or 2 (Table 1). This discrepancy strongly suggested that competitive binding had a greater effect on the in vivo Kd values than did molecular crowding. It should be noted that the large variability of each experimental plot from the fitting curve was caused by competitive binding. We simulated the effect of a fixed concentration of endogenous competitors on the in vivo Kd calculation by randomly changing the concentrations of two fluorescently labeled molecules (Fig. 4A and B). As expected, the in vivo Kd value was increased with the increase in the concentrations of endogenous competitors (Fig. 4C to E). The simulated plots were largely varied and dispersed, even though the concentrations of competitor proteins were fixed at constant values, as shown by the experimental data (Fig. 3). These simulation data were basically consistent with data reported in a previous study by Foo et al. (38). We compared fitting methods 1 and 2, which were used in this study, with fitting method 3, which was used in the study by Foo et al., and found that all of these methods provided almost the same in vivo Kd values (Fig. 4F and G). However, fitting methods 1 and 2 showed higher values for the coefficient of determination (R2) between the fitting and simulation data than fitting method 3 (Fig. 4H and I). Therefore, we applied these fitting methods to the measurement of in vivo Kd values in this study.
FIG 4.
Computer simulation of the effects of endogenous competitors on measurements of in vivo Kd values. (A) Schematic competitive binding model. Green fluorescent protein-fused exogenous protein A binds to red fluorescent protein-fused exogenous protein B. Endogenous proteins C and D competitively bind to proteins A and B, respectively. Dissociation constants for all bindings in this model were assumed to be 0.1 μM. The species measured by the autocorrelation function (ACF) and cross-correlation function (CCF) are indicated in the insets. (B) Comparison among three fitting methods for in vivo Kd calculations. While fitting methods 1 and 2 were applied in this study, Foo et al. (38) previously employed fitting method 3. (C to E) The initial concentrations of proteins A and B were set randomly with a fixed concentration set of competitor proteins C and D. The concentrations of the AB protein complex, [AB], at steady state were calculated in accordance with the competitive binding model shown in panel A. Simulated results were plotted and fitted according to fitting method 1 (C), fitting method 2 (D), and fitting method 3 (E). Even though the concentrations of endogenous competitor proteins C and D were fixed values, the simulated plots varied widely from the fitting curves. (F and G) Heat maps showing the logarithmic in vivo Kd values obtained by averaging fitting methods 1 and 2 (F) and by fitting method 3 (Kd3) (G), with the indicated concentration set of competitor proteins C and D. (H and I) Heat maps showing the coefficient of determination (R2) values obtained by averaging fitting methods 1 and 2 (H) and by fitting method 3 (I) with the indicated concentration sets of competitor proteins C and D. In several concentration sets of competitor proteins C and D, fitting methods 1 and 2 showed higher values for the coefficient of determination than those obtained by fitting method 3, indicating the advantage of the former method.
Quantitative in vivo Kd map of the EGFR-Ras-ERK MAP kinase pathway.
Quantitative Western blot analyses were used to measure the endogenous concentrations of proteins comprising the EGFR-Ras-ERK MAP kinase pathway in individual HeLa cells, including Shc, Grb2, MEK1, MEK2, ERK1, ERK2, RSK1, RSK2, and RSK3 (Fig. 5A and Table 2). These results, taken together with our previous data (22, 25, 51), are summarized in Table 2. Taking the in vivo Kd data (Table 1) and endogenous protein concentrations in a HeLa cell (Table 2) into account, we drew a quantitative interaction map of the EGFR-Ras-ERK MAP kinase pathway (Fig. 5B). In this scheme, the in vivo Kd values of Ras-CRaf and Ras-BRaf were the same as those of constitutively active Ras. Therefore, with the low abundance of Raf proteins, we speculated that signaling from Ras to Raf is the most inefficient part of the signal transduction pathway.
FIG 5.

Quantitative map of the EGFR-Ras-ERK signaling pathway in a HeLa cell. (A) Protein concentrations of endogenous MEK1 in a HeLa cell were determined by quantitative Western blotting, as described in Materials and Methods. (B) Schematic representation of the in vivo Kd map of the EGFR-Ras-ERK pathway. The diameter of the circles represents the concentration of the indicated protein, and the line thickness indicates the in vivo Kd values in a HeLa cell.
Model of the EGFR-Ras-ERK MAP kinase pathway.
We built a simulation model of the EGFR-Ras-ERK MAP kinase pathway (Fig. 6 and Tables 3 and 4). This model was comprised of 4 modules, the EGFR-Shc-Grb2-Sos module, the Ras-Raf activation module (52), the MEK-ERK processive phosphorylation module (9), and the ERK-Sos1 negative-feedback module (22). EGFR was phosphorylated (pEGFR) upon EGF stimulation. Shc was phosphorylated by pEGFR, followed by binding to Grb2 and/or the Grb2-Sos complex. The pShc-Grb2-Sos ternary complex was associated with pEGFR. The bindings of Grb2 or the Grb2-Sos complex to pEGFR were ignored because we could not observe Grb2 binding to pEGFR in HeLa cells and A431 cells upon EGF stimulation (our unpublished observations), and we did not include the bindings of Shc, pShc, or pShc-Grb2 to pEGFR to avoid unnecessary complexity of the model.
The in vivo Kd values obtained in this study were used as the kinetic parameters to define 14 reactions in the model (Fig. 6, magenta lines), in addition to Shc phosphorylation/dephosphorylation rates (reactions 4 and 5) (Fig. 6 and 7A and B), the time course of EGFR and MEK phosphorylation upon EGF stimulation (Fig. 7C to E), and RSK1/RSK2/RSK3 nuclear import/export rates (reaction 36) (Fig. 6 and 7F and G). To constrain the parameters during simulation, we measured the stoichiometry of MEK phosphorylation by Phos-tag Western blotting (53) (Fig. 7H to K). Roughly 33% of MEK1/2 proteins were phosphorylated at Ser217/221 residues by Raf 10 min after EGF stimulation. Other kinetic parameters were either determined experimentally (Fig. 6, green lines) or estimated from experimental data from previous studies (22, 51) (Fig. 6, cyan lines).
FIG 7.
Quantifications of kinetic parameters for EGFR-Ras-ERK signaling. (A and B) Shc1 phosphorylation and dephosphorylation rates (reactions 4 and 5) (Fig. 6). Serum-starved HeLa cells were treated with 50 ng/ml EGF, followed by the addition of an EGFR inhibitor, 10 μM PD-153035, 5 min later. Cell lysates were subjected to Western blotting with antiphosphotyrosine and anti-pY317-Shc antibodies. Representative results of three independent experiments are shown (A). Averaged pShc intensity values were plotted as a function of time, with standard deviations (B). First, the pShc dephosphorylation rate constant was measured by fitting the experimental data with a single exponential function (blue line). The time constant of the fitted exponential function corresponds to the pShc dephosphorylation rate constant. Second, the Shc phosphorylation rate constant was calculated as follows. The time constant of phosphorylation was obtained by fitting the experimental data with a single exponential function (red line). The time constant value was subtracted from the pShc dephosphorylation rate constant and then divided by the EGFR concentration, providing the Shc phosphorylation rate by pEGFR. AU, arbitrary units. (C to E) Time courses of EGF-induced phosphorylation of EGFR and MEK1/2 in HeLa cells were examined by Western blotting (C) and quantified by densitometry (D and E). (F and G) Nuclear import and export rates of RSK1, RSK2, and RSK3 proteins. HeLa cells expressing HaloTag-fused RSK1, RSK2, or RSK3 were treated with the TMR ligand, followed by confocal microscopy (F), and the ratio of fluorescence intensity in the nucleus to that in the cytoplasm was quantified (G). The concentration ratio is equal to the ratio of the nuclear import rate to the nuclear export rate. Therefore, we set the nuclear import rate at 0.001/s, and the nuclear export rates were calculated by the ratio of nuclear intensity to cytoplasmic intensity. Bar, 15 μm. (H to K) Stoichiometry of phosphorylated MEK. Recombinant GST-MEK1 protein was phosphorylated in vitro by constitutively active CRaf kinase for the indicated time periods. The eluates were subjected to Phos-tag Western blotting. The fraction of phospho-GST-MEK1 was measured by the reduction of nonphosphorylated GST-MEK1 (H, left). The same eluates were also subjected to conventional Western blotting (I). The fractions of phosphorylated MEK1 were plotted as a function of the pMEK/MEK ratio and fitted with a linear function (K). EGF-stimulated HeLa cell lysates were analyzed by conventional Western blotting (J), and the fraction of phosphorylated MEK1/2 was calculated by using the reference function obtained in panel K.
Numerical simulation of the EGFR-Ras-ERK MAP kinase pathway based on in vivo Kd values.
Numerical simulation of EGF-induced changes in the activity of EGFR (pEGFR), Ras (Ras-GTP), MEK (pMEK), and ERK (pTpY-ERK) in this original model demonstrated much lower values for Ras, MEK, and ERK than those obtained in experiments (Fig. 8A). This could be reasonable because in vivo Kd values obtained in this study were substantially higher than those used in the original models. We examined which reactions were rate-limiting steps and found three responsible reactions: from upstream, association of the pShc-Grb2-Sos complex with pEGFR (Fig. 8B, left), MEK phosphorylation by Raf (Fig. 8B, middle), and ERK phosphorylation by MEK (Fig. 8B, right). The latter two reactions were updated with the in vivo Kd values obtained in this study, resulting in decreases of the affinity of Ras-Raf binding and MEK-ERK binding, respectively. These reductions were corrected by multiplying the constant values, which approximately corresponded to the fold change in the reduction. Meanwhile, the detailed mechanisms of the association of the pShc-Grb2-Sos complex with pEGFR were not addressed, because we could not measure the dissociation constant of the reaction that took place at the plasma membrane.
FIG 8.
Effect of multivalent binding of Shc1 to pEGFR on EGFR-Ras-ERK signaling. (A) EGF-induced kinetics of phospho-EGFR (pEGFR), Ras-GTP, phospho-MEK (pMEK), and phospho-ERK (pERK) were simulated in the models according to data shown in Fig. 6, without any corrections, and showed low signal transmission from Ras-GTP in comparison to the experimental data. The experimental time courses of Ras-GTP and pTpY-ERK were obtained in our previous studies (9, 22). The time courses of pEGFR and pMEK were quantified in this study (Fig. 7). (B) Three reactions involved in rate-limiting steps are shown. (C) Heat map representing the logarithmic residual sum square values between the experiments and numerical simulations, which varied with respect to both the dissociation constant and stoichiometry values of the reaction for Shc1-pEGFR binding. Blue indicates a better fit of the simulation to experiments. D and E indicate parameter sets used in panels D and E (best fit), respectively. (D) After parameter corrections for MEK phosphorylation (B, middle) and ERK phosphorylation (B, right), the EGF-induced kinetics of Ras-GTP, pMEK, and pERK were calculated with a stoichiometry value of 1.0 and fitted to experiments by changing the dissociation constant value (B, left). The best-fit dissociation constant was 0.057 μM. (E) EGF-induced kinetics of Ras-GTP, pMEK, and pERK were calculated with a dissociation constant of 0.25 μM and the indicated stoichiometry values. (F to H) Multiple bindings of Shc to pEGFR. HeLa cells expressing mEGFP, HaloTag, or both EGFR-mEGFP and HaloTag-Shc were cocultured at equal ratios. The boxed region in panel F is magnified in panel G. After incubation with the TMR ligand and serum starvation for 1 h, the cells were stimulated with 10 ng/ml EGF for 15 min, followed by fixation. By FCS, the concentrations of mEGFP and HaloTag-TMR were determined to be 2.5 μM and 12 μM, respectively. These values were used to calculate the ratio of HaloTag-Shc to EGFR-mEGFP at each endosome in panel H (n = 78 endosomes from 28 cells). Bars, 20 μm (F) and 10 μm (G).
Role of multiple Shc bindings to pEGFR.
To assess the role of pEGFR-Shc binding, we examined the dissociation constant and stoichiometry of binding by employing our quantitative simulation model. The phosphotyrosine binding (PTB) domain of Shc1 has been reported to interact with multiple sites of phosphorylated tyrosine residues of pEGFR. Jones et al. previously measured the dissociation constants of bindings between the Shc1 PTB domain and 6 phosphotyrosine residues of pEGFR by using protein microarrays. The in vitro Kd values were 0.040 μM for pY1172, 0.133 μM for pY1138, 0.317 μM for pY1192, 0.321 μM for pY1110, 0.628 μM for pY1016, and 1.152 μM for pY998 (54). To examine the role of these multiple binding sites, we numerically changed the parameters of binding between pEGFR and Shc, with ranges of 0.001 μM to 10 μM for the dissociation constant and 0 to 6 for stoichiometry, and calculated the residual sum square (RSS) values, which were a measure of how well the simulated data set fit with the experimental data. In Fig. 8C, blue represents a better fit to the experimental data. If a single binding site, namely, a stoichiometry value of 1, was assumed in the model, the best fit was obtained with a Kd value of 0.057 μM, which seemed to be implausible because this value was almost comparable to the lowest in vitro Kd value of pEGFR-Shc binding, 0.040 μM (54). Furthermore, the time courses of activation of Ras, MEK, and ERK were very dull under this condition (Fig. 8D). On the other hand, both the dissociation constant and stoichiometry were fitted to the experimental data, providing values of 0.250 μM and 5.9, respectively (Fig. 8C and E). We set the Kd value at 0.25 μM and examined the effect of stoichiometry on the time courses of activation of Ras, MEK, and ERK. We found that high stoichiometry is essential to recapitulate the peak activations of Ras, MEK, and ERK (Fig. 8E).
The prediction by the simulation model urged us to confirm experimentally the stoichiometry of Shc binding to EGFR. For this, we first established two stable HeLa cell lines expressing mEGFP or HaloTag, which were used to calibrate the concentrations of mEGFP and HaloTag by FCS. Next, HeLa cells stably expressing both EGFR-mEGFP and HaloTag-Shc were stimulated with EGF for 15 min to induce accumulation of EGFR-mEGFP and HaloTag-Shc on the endosomes (Fig. 8F and G). The concentration ratio of EGFR-mEGFP to HaloTag-Shc was determined on each endosome and plotted in a histogram (Fig. 8H). We found that approximately 3 Shc molecules associated with EGFR at endosomes upon EGF stimulation.
DISCUSSION
In this study, we established a method for obtaining the in vivo Kd in living cells by FCCS and used this method to determine 22 in vivo Kd values for the EGFR-Ras-ERK MAP kinase pathway. By combining these in vivo Kd values with intracellular concentrations quantified by Western blotting, we built a simulation model of this signaling pathway. This model indicated the possible effect of multivalent binding of Shc proteins to pEGFR on sufficient peak activations of Ras, MEK, and ERK. Therefore, to the best of our knowledge, this study is the first to show the essential role played by multiple Shc bindings to pEGFR in the EGFR signaling pathway. These results clearly demonstrate the validity of this method and quantitative simulation model. However, there are substantial discrepancies between the experimental and simulated results, especially for the transient peak activity and sustained basal activity of MEK and ERK upon EGF stimulation (Fig. 8E), suggesting the existence of other possible regulations to be addressed in the future.
The in vivo Kd is influenced by both competitive binding by a number of proteins and molecular crowding (Fig. 1B). Competitive bindings increase the in vivo Kd value, whereas molecular crowding decreases the in vivo Kd value through the excluded-volume effect in general (14, 15). Intriguingly, almost all of the in vivo Kd values determined in this study were higher than those measured previously in in vitro assays (Table 1). These findings suggested that competitive bindings, rather than molecular crowding, made a more significant contribution to in vivo Kd values. Of note, we could not exclude other potential explanations for the increased in vivo Kd values, such as changes in conformation and molecular crowding, under certain circumstances. Foo et al. reported previously that the in vivo Kd values of Cdc42 and its effectors could become close to the in vitro Kd values by introducing correcting factors, such as the maturation efficiency of fluorescent proteins, photobleaching, and fluorescence resonance energy transfer (FRET) (38). The application of these corrections to our data might lead to an improvement in the discrepancies between the in vivo and in vitro Kd values to some extent. However, the increase in in vivo Kd values is unavoidable and remarkable in the binding of a protein containing multiple domains for protein-protein bindings, such as adaptor molecules. Although we could not entirely exclude the possibility that other factors such as ionic strength alter Kd values, in vivo Kd values have a potential advantage to directly demonstrate how much proteins form complexes in living cells.
The quantitative in vivo Kd measurements clarified several regulatory mechanisms underlying the EGFR-Ras-ERK pathway. Research groups have studied the role of growth factor-induced Sos1 phosphorylation but have not reached the same conclusion. Sos1 has at least 4 serine residues phosphorylated mainly by ERK in the C-terminal domain, resulting in negative regulation of Sos1 as a Ras activator. Two regulatory mechanisms have been suggested for the phosphorylation-mediated negative regulation of Sos1: a phosphorylation-induced disassembly of the Grb2-Sos1 complex (42, 43) and a phosphorylation-induced dissociation of the Grb2-Sos1 complex from the activated EGF receptors (40, 55). However, most of those experiments were performed under in vitro conditions. Our results revealed that pretreatment with TPA or EGF and Na3VO4 increased the in vivo Kd value of the Grb2-Sos1 complex (Fig. 3C and D and Table 1), supporting the former mechanisms for negative feedback from ERK to Sos1. In addition, we found that HRasdCT-V12 bound preferentially to the CRaf-S259A mutant rather than to the CRaf WT (Fig. 3E and Table 1). This result agreed with previously reported findings showing a sequential state transition model of CRaf for its Ras binding; CRaf phosphorylated on Ser259 adopts a closed inactive conformation, and dephosphorylation of Ser259 induces CRaf to adopt a semiclosed inactive state to associate with Ras-GTP (24). In contrast, the in vivo Kd values for HRas binding to the BRaf WT and the BRaf-S364A mutant, which corresponded to the CRaf-S259A mutant, did not show any difference (Table 1). These results indicated a clear difference between Ras-CRaf WT binding and Ras–CRaf-S259A binding but not between Ras-BRaf WT and Ras–BRaf-S364A binding (Fig. 3 and Table 1). It has been well characterized that 14-3-3 binds to the CRaf S259 phosphorylation site to form a closed inactive state (24), and the BRaf S364 phosphorylation site is thought to be equivalent to the CRaf S259 phosphorylation site. We suggest that the CRaf S259 site is highly phosphorylated even in the basal state, and therefore, most CRaf WT proteins form a closed state, while the BRaf S364 site is not sufficiently phosphorylated, and therefore, a substantial amount of BRaf WT proteins form an open state. The other possibility is that another phosphoserine residue serves as the 14-3-3 binding site in BRaf (25). Notably, for technical reasons, we had to omit some important regulatory mechanisms. For example, in vitro analyses of CRaf-lipid interactions have revealed two distinct phospholipid binding regions within CRaf, at amino acid residues 139 to 184 and 390 to 423, for phosphatidylserine and phosphatidic acid, respectively (56, 57). These two regions function to promote translocation of CRaf to the membrane and to increase the affinity for GTP-Ras. Therefore, we might have overestimated in vivo Kd values of Ras-Raf binding, which was measured only in the cytoplasm. Even with this kind of limitation, we believe that measurements of in vivo Kd values by FCCS will provide compelling quantitative data for interactions in living cells.
We improved and optimized some of the experimental conditions of FCCS for our in vivo Kd measurements. EGFP and mCherry have hitherto been utilized as a fluorescent protein pair for FCCS measurements (12, 36, 37). However, the mCherry protein has been reported to exhibit slower chromophore maturation, a much greater triplet component, weaker photostability than the EGFP protein, and, consequently, a decrease in the signal-to-noise ratio of FCCS measurements (38). We found that HaloTag-TMR overcame all these flaws. The membrane-permeable TMR ligand needs to be added after HaloTag expression in cells. Based on the positive-control data, almost all HaloTag proteins were covalently attached to the TMR ligand (Fig. 2B), and the amount of free TMR ligand in the medium was negligible after the medium change. Further improvements will be needed to overcome the following limitations in our system. First, there is a limitation associated with the measurement of the in vivo Kd value for binding on the membrane. As we mentioned above, FCCS measurements of HRas and Raf binding on the membrane did not provide accurate and reproducible results, possibly because of the low diffusion constant, heterogeneous membrane structure, and leakage of the fluorescence signal of Raf from the cytoplasm. Second, the lower limit for the measurements of the in vivo Kd is also problematic. In our experimental setup, we could not reproducibly calculate the in vivo Kd at values of <0.1 μM. This was due mainly to the sensitivity of fluorescence detection by imaging of cells with low EGFP and HaloTag-TMR expression levels, in which autofluorescence and substantial photobleaching interfered with data acquisition. The third limitation concerns the measurement of the in vivo Kd values in ternary protein complexes. There are many scaffold proteins in the EGFR-Ras-ERK MAP kinase pathway that regulate signaling efficiency and specificity (58–60). Measurements of the in vivo Kd values in ternary protein complexes by FCCS will shed new light on how these scaffold proteins regulate efficiency, specificity, and divergence in the signaling pathway more quantitatively.
We established a method for measuring the in vivo Kd in living cells by FCCS and quantified >20 in vivo Kd values for protein-protein interactions involved in the EGFR-Ras-ERK MAP kinase pathway in living HeLa cells. These in vivo Kd values provide quantitative data that will help to elucidate various signaling pathways.
ACKNOWLEDGMENTS
We thank Kazunari Kaizu and Koichi Takahasi (QBiC, Japan) and the members of the Matsuda Laboratory for helpful discussions. We are also grateful to the Center for Meso-Bio Single Molecule Imaging and the Institute for Integrated Cell-Material Sciences at Kyoto University for the use of the microscope.
K.A. and M.M. were supported by the Research Program of Innovative Cell Biology by Innovative Technology (Cell Innovation) and the Platform for Dynamic Approaches to Living System from the Ministry of Education, Culture, Sports, and Science, Japan. K.A. was supported by the JST PRESTO program and by JSPS KAKENHI (23136504 and 25136706). W.S. was supported by the Yoshida Scholarship Foundation.
Footnotes
Published ahead of print 23 June 2014
REFERENCES
- 1.Chang L, Karin M. 2001. Mammalian MAP kinase signalling cascades. Nature 410:37–40. 10.1038/35065000 [DOI] [PubMed] [Google Scholar]
- 2.Chen Z, Gibson TB, Robinson F, Silvestro L, Pearson G, Xu B, Wright A, Vanderbilt C, Cobb MH. 2001. MAP kinases. Chem. Rev. 101:2449–2476. 10.1021/cr000241p [DOI] [PubMed] [Google Scholar]
- 3.Qi M, Elion EA. 2005. MAP kinase pathways. J. Cell Sci. 118:3569–3572. 10.1242/jcs.02470 [DOI] [PubMed] [Google Scholar]
- 4.Bhalla US, Ram PT, Iyengar R. 2002. MAP kinase phosphatase as a locus of flexibility in a mitogen-activated protein kinase signaling network. Science 297:1018–1023. 10.1126/science.1068873 [DOI] [PubMed] [Google Scholar]
- 5.Nakakuki T, Birtwistle MR, Saeki Y, Yumoto N, Ide K, Nagashima T, Brusch L, Ogunnaike BA, Okada-Hatakeyama M, Kholodenko BN. 2010. Ligand-specific c-Fos expression emerges from the spatiotemporal control of ErbB network dynamics. Cell 141:884–896. 10.1016/j.cell.2010.03.054 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Sasagawa S, Ozaki Y, Fujita K, Kuroda S. 2005. Prediction and validation of the distinct dynamics of transient and sustained ERK activation. Nat. Cell Biol. 7:365–373. 10.1038/ncb1233 [DOI] [PubMed] [Google Scholar]
- 7.Schoeberl B, Eichler-Jonsson C, Gilles ED, Muller G. 2002. Computational modeling of the dynamics of the MAP kinase cascade activated by surface and internalized EGF receptors. Nat. Biotechnol. 20:370–375. 10.1038/nbt0402-370 [DOI] [PubMed] [Google Scholar]
- 8.Bandyopadhyay S, Chiang CY, Srivastava J, Gersten M, White S, Bell R, Kurschner C, Martin C, Smoot M, Sahasrabudhe S, Barber DL, Chanda SK, Ideker T. 2010. A human MAP kinase interactome. Nat. Methods 7:801–805. 10.1038/nmeth.1506 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Aoki K, Yamada M, Kunida K, Yasuda S, Matsuda M. 2011. Processive phosphorylation of ERK MAP kinase in mammalian cells. Proc. Natl. Acad. Sci. U. S. A. 108:12675–12680. 10.1073/pnas.1104030108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Kasai RS, Suzuki KG, Prossnitz ER, Koyama-Honda I, Nakada C, Fujiwara TK, Kusumi A. 2011. Full characterization of GPCR monomer-dimer dynamic equilibrium by single molecule imaging. J. Cell Biol. 192:463–480. 10.1083/jcb.201009128 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Shi X, Foo YH, Sudhaharan T, Chong SW, Korzh V, Ahmed S, Wohland T. 2009. Determination of dissociation constants in living zebrafish embryos with single wavelength fluorescence cross-correlation spectroscopy. Biophys. J. 97:678–686. 10.1016/j.bpj.2009.05.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Sudhaharan T, Liu P, Foo YH, Bu W, Lim KB, Wohland T, Ahmed S. 2009. Determination of in vivo dissociation constant, KD, of Cdc42-effector complexes in live mammalian cells using single wavelength fluorescence cross-correlation spectroscopy. J. Biol. Chem. 284:13602–13609. 10.1074/jbc.M900894200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Minton AP. 2001. The influence of macromolecular crowding and macromolecular confinement on biochemical reactions in physiological media. J. Biol. Chem. 276:10577–10580. 10.1074/jbc.R100005200 [DOI] [PubMed] [Google Scholar]
- 14.Minton AP. 1983. The effect of volume occupancy upon the thermodynamic activity of proteins: some biochemical consequences. Mol. Cell. Biochem. 55:119–140. 10.1007/BF00673707 [DOI] [PubMed] [Google Scholar]
- 15.Zimmerman SB, Minton AP. 1993. Macromolecular crowding: biochemical, biophysical, and physiological consequences. Annu. Rev. Biophys. Biomol. Struct. 22:27–65. 10.1146/annurev.bb.22.060193.000331 [DOI] [PubMed] [Google Scholar]
- 16.Bacia K, Kim SA, Schwille P. 2006. Fluorescence cross-correlation spectroscopy in living cells. Nat. Methods 3:83–89. 10.1038/nmeth822 [DOI] [PubMed] [Google Scholar]
- 17.Akagi T, Sasai K, Hanafusa H. 2003. Refractory nature of normal human diploid fibroblasts with respect to oncogene-mediated transformation. Proc. Natl. Acad. Sci. U. S. A. 100:13567–13572. 10.1073/pnas.1834876100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Niwa H, Yamamura K, Miyazaki J. 1991. Efficient selection for high-expression transfectants with a novel eukaryotic vector. Gene 108:193–199. 10.1016/0378-1119(91)90434-D [DOI] [PubMed] [Google Scholar]
- 19.Johannessen CM, Boehm JS, Kim SY, Thomas SR, Wardwell L, Johnson LA, Emery CM, Stransky N, Cogdill AP, Barretina J, Caponigro G, Hieronymus H, Murray RR, Salehi-Ashtiani K, Hill DE, Vidal M, Zhao JJ, Yang X, Alkan O, Kim S, Harris JL, Wilson CJ, Myer VE, Finan PM, Root DE, Roberts TM, Golub T, Flaherty KT, Dummer R, Weber BL, Sellers WR, Schlegel R, Wargo JA, Hahn WC, Garraway LA. 2010. COT drives resistance to RAF inhibition through MAP kinase pathway reactivation. Nature 468:968–972. 10.1038/nature09627 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Gotoh N, Toyoda M, Shibuya M. 1997. Tyrosine phosphorylation sites at amino acids 239 and 240 of Shc are involved in epidermal growth factor-induced mitogenic signaling that is distinct from Ras/mitogen-activated protein kinase activation. Mol. Cell. Biol. 17:1824–1831 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Matuoka K, Shibata M, Yamakawa A, Takenawa T. 1992. Cloning of ASH, a ubiquitous protein composed of one Src homology region (SH) 2 and two SH3 domains, from human and rat cDNA libraries. Proc. Natl. Acad. Sci. U. S. A. 89:9015–9019. 10.1073/pnas.89.19.9015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Kamioka Y, Yasuda S, Fujita Y, Aoki K, Matsuda M. 2010. Multiple decisive phosphorylation sites for the negative feedback regulation of SOS1 via ERK. J. Biol. Chem. 285:33540–33548. 10.1074/jbc.M110.135517 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Aoki K, Nakamura T, Inoue T, Meyer T, Matsuda M. 2007. An essential role for the SHIP2-dependent negative feedback loop in neuritogenesis of nerve growth factor-stimulated PC12 cells. J. Cell Biol. 177:817–827. 10.1083/jcb.200609017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Terai K, Matsuda M. 2005. Ras binding opens c-Raf to expose the docking site for mitogen-activated protein kinase kinase. EMBO Rep. 6:251–255. 10.1038/sj.embor.7400349 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Terai K, Matsuda M. 2006. The amino-terminal B-Raf-specific region mediates calcium-dependent homo- and hetero-dimerization of Raf. EMBO J. 25:3556–3564. 10.1038/sj.emboj.7601241 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Torii S, Kusakabe M, Yamamoto T, Maekawa M, Nishida E. 2004. Sef is a spatial regulator for Ras/MAP kinase signaling. Dev. Cell 7:33–44. 10.1016/j.devcel.2004.05.019 [DOI] [PubMed] [Google Scholar]
- 27.Komatsu N, Aoki K, Yamada M, Yukinaga H, Fujita Y, Kamioka Y, Matsuda M. 2011. Development of an optimized backbone of FRET biosensors for kinases and GTPases. Mol. Biol. Cell 22:4647–4656. 10.1091/mbc.E11-01-0072 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Yusa K, Rad R, Takeda J, Bradley A. 2009. Generation of transgene-free induced pluripotent mouse stem cells by the piggyBac transposon. Nat. Methods 6:363–369. 10.1038/nmeth.1323 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Kinjo M, Sakata H, Mikuni S. 2011. Basic fluorescence correlation spectroscopy setup and measurement. Cold Spring Harb. Protoc. 2011:1262–1266. 10.1101/pdb.prot065938 [DOI] [PubMed] [Google Scholar]
- 30.Bogdanov AM, Bogdanova EA, Chudakov DM, Gorodnicheva TV, Lukyanov S, Lukyanov KA. 2009. Cell culture medium affects GFP photostability: a solution. Nat. Methods 6:859–860. 10.1038/nmeth1209-859 [DOI] [PubMed] [Google Scholar]
- 31.Magde D, Elson EL, Webb WW. 1974. Fluorescence correlation spectroscopy. II. An experimental realization. Biopolymers 13:29–61 [DOI] [PubMed] [Google Scholar]
- 32.Kogure T, Karasawa S, Araki T, Saito K, Kinjo M, Miyawaki A. 2006. A fluorescent variant of a protein from the stony coral Montipora facilitates dual-color single-laser fluorescence cross-correlation spectroscopy. Nat. Biotechnol. 24:577–581. 10.1038/nbt1207 [DOI] [PubMed] [Google Scholar]
- 33.Funahashi A, Morohashi M, Kitano H, Tanimura N. 2003. CellDesigner: a process diagram editor for gene-regulatory and biochemical networks. Biosilico 1:159–162. 10.1016/S1478-5382(03)02370-9 [DOI] [Google Scholar]
- 34.Kitano H, Funahashi A, Matsuoka Y, Oda K. 2005. Using process diagrams for the graphical representation of biological networks. Nat. Biotechnol. 23:961–966. 10.1038/nbt1111 [DOI] [PubMed] [Google Scholar]
- 35.Sauro HM, Hucka M, Finney A, Wellock C, Bolouri H, Doyle J, Kitano H. 2003. Next generation simulation tools: the Systems Biology Workbench and BioSPICE integration. OMICS 7:355–372. 10.1089/153623103322637670 [DOI] [PubMed] [Google Scholar]
- 36.Maeder CI, Hink MA, Kinkhabwala A, Mayr R, Bastiaens PI, Knop M. 2007. Spatial regulation of Fus3 MAP kinase activity through a reaction-diffusion mechanism in yeast pheromone signalling. Nat. Cell Biol. 9:1319–1326. 10.1038/ncb1652 [DOI] [PubMed] [Google Scholar]
- 37.Slaughter BD, Schwartz JW, Li R. 2007. Mapping dynamic protein interactions in MAP kinase signaling using live-cell fluorescence fluctuation spectroscopy and imaging. Proc. Natl. Acad. Sci. U. S. A. 104:20320–20325. 10.1073/pnas.0710336105 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Foo YH, Naredi-Rainer N, Lamb DC, Ahmed S, Wohland T. 2012. Factors affecting the quantification of biomolecular interactions by fluorescence cross-correlation spectroscopy. Biophys. J. 102:1174–1183. 10.1016/j.bpj.2012.01.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Los GV, Encell LP, McDougall MG, Hartzell DD, Karassina N, Zimprich C, Wood MG, Learish R, Ohana RF, Urh M, Simpson D, Mendez J, Zimmerman K, Otto P, Vidugiris G, Zhu J, Darzins A, Klaubert DH, Bulleit RF, Wood KV. 2008. HaloTag: a novel protein labeling technology for cell imaging and protein analysis. ACS Chem. Biol. 3:373–382. 10.1021/cb800025k [DOI] [PubMed] [Google Scholar]
- 40.Rozakis-Adcock M, McGlade J, Mbamalu G, Pelicci G, Daly R, Li W, Batzer A, Thomas S, Brugge J, Pelicci PG, Schlessinger J, Pawson T. 1992. Association of the Shc and Grb2/Sem5 SH2-containing proteins is implicated in activation of the Ras pathway by tyrosine kinases. Nature 360:689–692. 10.1038/360689a0 [DOI] [PubMed] [Google Scholar]
- 41.Skolnik EY, Lee CH, Batzer A, Vicentini LM, Zhou M, Daly R, Myers MJ, Jr, Backer JM, Ullrich A, White MF, Schlessinger J. 1993. The SH2/SH3 domain-containing protein GRB2 interacts with tyrosine-phosphorylated IRS1 and Shc: implications for insulin control of ras signalling. EMBO J. 12:1929–1936 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Cherniack AD, Klarlund JK, Conway BR, Czech MP. 1995. Disassembly of Son-of-sevenless proteins from Grb2 during p21ras desensitization by insulin. J. Biol. Chem. 270:1485–1488. 10.1074/jbc.270.4.1485 [DOI] [PubMed] [Google Scholar]
- 43.Corbalan-Garcia S, Degenhardt KR, Bar-Sagi D. 1996. Insulin-induced dissociation of Sos from Grb2 does not contribute to the down regulation of Ras activation. Oncogene 12:1063–1068 [PubMed] [Google Scholar]
- 44.Rozakis-Adcock M, van der Geer P, Mbamalu G, Pawson T. 1995. MAP kinase phosphorylation of mSos1 promotes dissociation of mSos1-Shc and mSos1-EGF receptor complexes. Oncogene 11:1417–1426 [PubMed] [Google Scholar]
- 45.Chiu VK, Bivona T, Hach A, Sajous JB, Silletti J, Wiener H, Johnson RL, II, Cox AD, Philips MR. 2002. Ras signalling on the endoplasmic reticulum and the Golgi. Nat. Cell Biol. 4:343–350. 10.1038/ncb783 [DOI] [PubMed] [Google Scholar]
- 46.Hancock JF, Paterson H, Marshall CJ. 1990. A polybasic domain or palmitoylation is required in addition to the CAAX motif to localize p21ras to the plasma membrane. Cell 63:133–139. 10.1016/0092-8674(90)90294-O [DOI] [PubMed] [Google Scholar]
- 47.Fukuda M, Gotoh Y, Nishida E. 1997. Interaction of MAP kinase with MAP kinase kinase: its possible role in the control of nucleocytoplasmic transport of MAP kinase. EMBO J. 16:1901–1908. 10.1093/emboj/16.8.1901 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Dimitri CA, Dowdle W, MacKeigan JP, Blenis J, Murphy LO. 2005. Spatially separate docking sites on ERK2 regulate distinct signaling events in vivo. Curr. Biol. 15:1319–1324. 10.1016/j.cub.2005.06.037 [DOI] [PubMed] [Google Scholar]
- 49.Smith L, Parizi-Robinson M, Zhu MS, Zhi G, Fukui R, Kamm KE, Stull JT. 2002. Properties of long myosin light chain kinase binding to F-actin in vitro and in vivo. J. Biol. Chem. 277:35597–35604. 10.1074/jbc.M206483200 [DOI] [PubMed] [Google Scholar]
- 50.Zhao Y, Bjorbaek C, Moller DE. 1996. Regulation and interaction of pp90(rsk) isoforms with mitogen-activated protein kinases. J. Biol. Chem. 271:29773–29779. 10.1074/jbc.271.47.29773 [DOI] [PubMed] [Google Scholar]
- 51.Fujioka A, Terai K, Itoh RE, Aoki K, Nakamura T, Kuroda S, Nishida E, Matsuda M. 2006. Dynamics of the Ras/ERK MAPK cascade as monitored by fluorescent probes. J. Biol. Chem. 281:8917–8926. 10.1074/jbc.M509344200 [DOI] [PubMed] [Google Scholar]
- 52.Matsunaga-Udagawa R, Fujita Y, Yoshiki S, Terai K, Kamioka Y, Kiyokawa E, Yugi K, Aoki K, Matsuda M. 2010. The scaffold protein Shoc2/SUR-8 accelerates the interaction of Ras and Raf. J. Biol. Chem. 285:7818–7826. 10.1074/jbc.M109.053975 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Kinoshita E, Kinoshita-Kikuta E, Takiyama K, Koike T. 2006. Phosphate-binding tag, a new tool to visualize phosphorylated proteins. Mol. Cell. Proteomics 5:749–757. 10.1074/mcp.T500024-MCP200 [DOI] [PubMed] [Google Scholar]
- 54.Jones RB, Gordus A, Krall JA, MacBeath G. 2006. A quantitative protein interaction network for the ErbB receptors using protein microarrays. Nature 439:168–174. 10.1038/nature04177 [DOI] [PubMed] [Google Scholar]
- 55.Porfiri E, McCormick F. 1996. Regulation of epidermal growth factor receptor signaling by phosphorylation of the ras exchange factor hSOS1. J. Biol. Chem. 271:5871–5877. 10.1074/jbc.271.10.5871 [DOI] [PubMed] [Google Scholar]
- 56.Ghosh S, Strum JC, Sciorra VA, Daniel L, Bell RM. 1996. Raf-1 kinase possesses distinct binding domains for phosphatidylserine and phosphatidic acid. Phosphatidic acid regulates the translocation of Raf-1 in 12-O-tetradecanoylphorbol-13-acetate-stimulated Madin-Darby canine kidney cells. J. Biol. Chem. 271:8472–8480 [DOI] [PubMed] [Google Scholar]
- 57.Ghosh S, Xie WQ, Quest AF, Mabrouk GM, Strum JC, Bell RM. 1994. The cysteine-rich region of raf-1 kinase contains zinc, translocates to liposomes, and is adjacent to a segment that binds GTP-ras. J. Biol. Chem. 269:10000–10007 [PubMed] [Google Scholar]
- 58.Kolch W. 2005. Coordinating ERK/MAPK signalling through scaffolds and inhibitors. Nat. Rev. Mol. Cell Biol. 6:827–837. 10.1038/nrm1743 [DOI] [PubMed] [Google Scholar]
- 59.Shaul YD, Seger R. 2007. The MEK/ERK cascade: from signaling specificity to diverse functions. Biochim. Biophys. Acta 1773:1213–1226. 10.1016/j.bbamcr.2006.10.005 [DOI] [PubMed] [Google Scholar]
- 60.Shaw AS, Filbert EL. 2009. Scaffold proteins and immune-cell signalling. Nat. Rev. Immunol. 9:47–56. 10.1038/nri2473 [DOI] [PubMed] [Google Scholar]
- 61.Chook YM, Gish GD, Kay CM, Pai EF, Pawson T. 1996. The Grb2-mSos1 complex binds phosphopeptides with higher affinity than Grb2. J. Biol. Chem. 271:30472–30478. 10.1074/jbc.271.48.30472 [DOI] [PubMed] [Google Scholar]
- 62.Sastry L, Lin W, Wong WT, Di Fiore PP, Scoppa CA, King CR. 1995. Quantitative analysis of Grb2-Sos1 interaction: the N-terminal SH3 domain of Grb2 mediates affinity. Oncogene 11:1107–1112 [PubMed] [Google Scholar]
- 63.Houtman JC, Yamaguchi H, Barda-Saad M, Braiman A, Bowden B, Appella E, Schuck P, Samelson LE. 2006. Oligomerization of signaling complexes by the multipoint binding of GRB2 to both LAT and SOS1. Nat. Struct. Mol. Biol. 13:798–805. 10.1038/nsmb1133 [DOI] [PubMed] [Google Scholar]
- 64.Lemmon MA, Ladbury JE, Mandiyan V, Zhou M, Schlessinger J. 1994. Independent binding of peptide ligands to the SH2 and SH3 domains of Grb2. J. Biol. Chem. 269:31653–31658 [PubMed] [Google Scholar]
- 65.Kiel C, Filchtinski D, Spoerner M, Schreiber G, Kalbitzer HR, Herrmann C. 2009. Improved binding of raf to Ras. GDP is correlated with biological activity. J. Biol. Chem. 284:31893–31902. 10.1074/jbc.M109.031153 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Block C, Janknecht R, Herrmann C, Nassar N, Wittinghofer A. 1996. Quantitative structure-activity analysis correlating Ras/Raf interaction in vitro to Raf activation in vivo. Nat. Struct. Biol. 3:244–251. 10.1038/nsb0396-244 [DOI] [PubMed] [Google Scholar]
- 67.Fischer A, Hekman M, Kuhlmann J, Rubio I, Wiese S, Rapp UR. 2007. B- and C-RAF display essential differences in their binding to Ras: the isotype-specific N terminus of B-RAF facilitates Ras binding. J. Biol. Chem. 282:26503–26516. 10.1074/jbc.M607458200 [DOI] [PubMed] [Google Scholar]
- 68.Barbero N, Napione L, Visentin S, Alvaro M, Veglio A, Bussolino F, Viscardi G. 2011. A transient kinetic study between signaling proteins: the case of the MEK-ERK interaction. Chem. Sci. 2:1804–1809. 10.1039/c1sc00268f [DOI] [Google Scholar]
- 69.Haystead TA, Dent P, Wu J, Haystead CM, Sturgill TW. 1992. Ordered phosphorylation of p42mapk by MAP kinase kinase. FEBS Lett. 306:17–22. 10.1016/0014-5793(92)80828-5 [DOI] [PubMed] [Google Scholar]
- 70.Mansour SJ, Candia JM, Matsuura JE, Manning MC, Ahn NG. 1996. Interdependent domains controlling the enzymatic activity of mitogen-activated protein kinase kinase 1. Biochemistry 35:15529–15536. 10.1021/bi961854s [DOI] [PubMed] [Google Scholar]
- 71.Burkhard KA, Chen F, Shapiro P. 2011. Quantitative analysis of ERK2 interactions with substrate proteins: roles for kinase docking domains and activity in determining binding affinity. J. Biol. Chem. 286:2477–2485. 10.1074/jbc.M110.177899 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Yoshiki S, Matsunaga-Udagawa R, Aoki K, Kamioka Y, Kiyokawa E, Matsuda M. 2010. Ras and calcium signaling pathways converge at Raf1 via the Shoc2 scaffold protein. Mol. Biol. Cell 21:1088–1096. 10.1091/mbc.E09-06-0455 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Gureasko J, Galush WJ, Boykevisch S, Sondermann H, Bar-Sagi D, Groves JT, Kuriyan J. 2008. Membrane-dependent signal integration by the Ras activator Son of sevenless. Nat. Struct. Mol. Biol. 15:452–461. 10.1038/nsmb.1418 [DOI] [PMC free article] [PubMed] [Google Scholar]





