Top Banner
Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application Juliane Mai, * Saskia Trump, Rizwan Ali, § R. Louis Schiltz, { Gordon Hager, { Thomas Hanke, § Irina Lehmann, and Sabine Attinger k Department of Computational Hydrosystems and Department of Environmental Immunology, UFZ - Helmholtz Centre for Environmental Research, Leipzig, Germany; § Max Bergmann Center of Biomaterials and Institute of Materials Science, Dresden University of Technology, Dresden, Germany; { Laboratory of Receptor Biology and Gene Expression, National Cancer Institute, National Institutes of Health, Bethesda, Maryland; and k Institute for Geosciences, University of Jena, Jena, Germany ABSTRACT At present, fluorescence recovery after photobleaching (FRAP) data are interpreted using various types of reac- tion-diffusion (RD) models: the model type is usually fixed first, and corresponding model parameters are inferred subsequently. In this article, we describe what we believe to be a novel approach for RD modeling without using any assumptions of model type or parameters. To the best of our knowledge, this is the first attempt to address both model-type and parameter uncertainties in inverting FRAP data. We start from the most general RD model, which accounts for a flexible number of molecular fractions, all mobile, with different diffusion coefficients. The maximal number of possible binding partners is identified and optimal parameter sets for these models are determined in a global search of the parameter-space using the Simulated Annealing strategy. The numerical performance of the described techniques was assessed using artificial and experimental FRAP data. Our general RD model outperformed the standard RD models used previously in modeling FRAP measurements and showed that intracel- lular molecular mobility can only be described adequately by allowing for multiple RD processes. Therefore, it is important to search not only for the optimal parameter set but also for the optimal model type. INTRODUCTION In recent years, interest in using noninvasive methods to observe and analyze intracellular molecular mobility has increased dramatically (1–7). One technique widely used for this purpose is fluorescence recovery after photobleach- ing (FRAP) (8–16). To study the behavior of a fluorescent molecule in a live cell by FRAP, a specific region (i.e., bleaching spot) in the cell is defined and exposed to an intense excitation pulse sufficient to irreversibly inactivate fluorescence emission (Fig. 1). The recovery of fluorescence reflects the movement of new fluorescent molecules into the photobleached region. This increase in fluorescence over time is described by a so- called recovery curve, which can be used to extract informa- tion on mobility and binding of the monitored fluorescent molecules. Several reaction-diffusion models have been suggested for the analysis of such recovery curves. However, diffusion coefficients and reaction parameters can be deduced only if an analytical or numerical solution for these models can be determined. Until now, it was possible to calculate an analytical solution only by simplifying the model using assumptions about particular parameter values (8,9,17– 26). These assumptions are reflected in various model types, such as the reaction-dominant model, where the diffusion coefficients of the acting molecules are fixed, or the diffu- sion-dominant model, where all reaction processes are neglected. As implied by the model names, particular mech- anisms are supposed for the underlying biological process. To eliminate the necessity for these restrictions, we herein introduce the solution of the general reaction-diffusion model, which includes an unconfined number of reacting and diffusing compounds. A generalization is attended by an increasing number of model parameters, which raises the concern of overfitting. To address this potential problem, we adapted a method introduced by de Prony (27–29) to include a preprocessing step in which the maximal number of parameters that can be fitted reliably is deduced. After this preprocessing, the parameter values of diffu- sion and reaction still have to be determined. Hitherto, local search algorithms like the Levenberg-Marquardt algorithm were used for inversion. However, a major drawback of such a local approach is the sensitivity to the initial param- eter settings (17,30,31). Therefore, a global search strategy that is robust against initial settings is preferable. As an alternative to the restricted local search algorithms, we used Simulated Annealing (SA) as one of the possible global approaches to infer unbiased and more reliable parameter sets (32). In summary, our new approach consists of a multiple- reaction-diffusion model with predetermination of the possible number of parameters, which then are inverted by the global search algorithm SA. To demonstrate the strength of our new approach, we applied our generalized reaction- diffusion model to artificial and real FRAP data sets and compared the results with those from previously used Submitted August 24, 2010, and accepted for publication January 21, 2011. *Correspondence: [email protected] Editor: David E Wolf. Ó 2011 by the Biophysical Society 0006-3495/11/03/1178/11 $2.00 doi: 10.1016/j.bpj.2011.01.041 1178 Biophysical Journal Volume 100 March 2011 1178–1188
11

Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

Apr 26, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

1178 Biophysical Journal Volume 100 March 2011 1178–1188

Are Assumptions about the Model Type Necessary in Reaction-DiffusionModeling? A FRAP Application

Juliane Mai,†* Saskia Trump,‡ Rizwan Ali,§ R. Louis Schiltz,{ Gordon Hager,{ Thomas Hanke,§ Irina Lehmann,‡

and Sabine Attinger†k†Department of Computational Hydrosystems and ‡Department of Environmental Immunology, UFZ - Helmholtz Centre for EnvironmentalResearch, Leipzig, Germany; §Max Bergmann Center of Biomaterials and Institute of Materials Science, Dresden University of Technology,Dresden, Germany; {Laboratory of Receptor Biology and Gene Expression, National Cancer Institute, National Institutes of Health, Bethesda,Maryland; and kInstitute for Geosciences, University of Jena, Jena, Germany

ABSTRACT At present, fluorescence recovery after photobleaching (FRAP) data are interpreted using various types of reac-tion-diffusion (RD) models: the model type is usually fixed first, and corresponding model parameters are inferred subsequently.In this article, we describe what we believe to be a novel approach for RDmodeling without using any assumptions of model typeor parameters. To the best of our knowledge, this is the first attempt to address both model-type and parameter uncertainties ininverting FRAP data. We start from the most general RD model, which accounts for a flexible number of molecular fractions, allmobile, with different diffusion coefficients. The maximal number of possible binding partners is identified and optimal parametersets for these models are determined in a global search of the parameter-space using the Simulated Annealing strategy. Thenumerical performance of the described techniques was assessed using artificial and experimental FRAP data. Our generalRD model outperformed the standard RD models used previously in modeling FRAP measurements and showed that intracel-lular molecular mobility can only be described adequately by allowing for multiple RD processes. Therefore, it is important tosearch not only for the optimal parameter set but also for the optimal model type.

INTRODUCTION

In recent years, interest in using noninvasive methods toobserve and analyze intracellular molecular mobility hasincreased dramatically (1–7). One technique widely usedfor this purpose is fluorescence recovery after photobleach-ing (FRAP) (8–16).

To study the behavior of a fluorescent molecule in a livecell by FRAP, a specific region (i.e., bleaching spot) in thecell is defined and exposed to an intense excitation pulsesufficient to irreversibly inactivate fluorescence emission(Fig. 1). The recovery of fluorescence reflects the movementof new fluorescent molecules into the photobleached region.This increase in fluorescence over time is described by a so-called recovery curve, which can be used to extract informa-tion on mobility and binding of the monitored fluorescentmolecules.

Several reaction-diffusion models have been suggestedfor the analysis of such recovery curves. However, diffusioncoefficients and reaction parameters can be deduced only ifan analytical or numerical solution for these models can bedetermined. Until now, it was possible to calculate ananalytical solution only by simplifying the model usingassumptions about particular parameter values (8,9,17–26). These assumptions are reflected in various model types,such as the reaction-dominant model, where the diffusioncoefficients of the acting molecules are fixed, or the diffu-sion-dominant model, where all reaction processes are

Submitted August 24, 2010, and accepted for publication January 21, 2011.

*Correspondence: [email protected]

Editor: David E Wolf.

� 2011 by the Biophysical Society

0006-3495/11/03/1178/11 $2.00

neglected. As implied by the model names, particular mech-anisms are supposed for the underlying biological process.To eliminate the necessity for these restrictions, we hereinintroduce the solution of the general reaction-diffusionmodel, which includes an unconfined number of reactingand diffusing compounds.

A generalization is attended by an increasing number ofmodel parameters, which raises the concern of overfitting.To address this potential problem, we adapted a methodintroduced by de Prony (27–29) to include a preprocessingstep in which the maximal number of parameters that canbe fitted reliably is deduced.

After this preprocessing, the parameter values of diffu-sion and reaction still have to be determined. Hitherto, localsearch algorithms like the Levenberg-Marquardt algorithmwere used for inversion. However, a major drawback ofsuch a local approach is the sensitivity to the initial param-eter settings (17,30,31). Therefore, a global search strategythat is robust against initial settings is preferable. As analternative to the restricted local search algorithms, weused Simulated Annealing (SA) as one of the possible globalapproaches to infer unbiased and more reliable parametersets (32).

In summary, our new approach consists of a multiple-reaction-diffusion model with predetermination of thepossible number of parameters, which then are inverted bythe global search algorithm SA. To demonstrate the strengthof our new approach, we applied our generalized reaction-diffusion model to artificial and real FRAP data sets andcompared the results with those from previously used

doi: 10.1016/j.bpj.2011.01.041

Page 2: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

FIGURE 1 Schematic representation of FRAP.

(A–D) A FRAP experiment is based on the bleach-

ing of fluorescent molecules (gray area) in a prede-

fined region of interest (ROI, black circle) and the

subsequent recovery of fluorescence intensity in

this region in a predefined region of interest.

A General RD Model for FRAP Analysis 1179

models. To our knowledge, this is the first attempt to use ananalytical approach to address both model type and param-eter uncertainties in inverting FRAP data.

MATERIALS AND METHODS

Analysis of FRAP data: A theoretical approach

In this section, we first introduce the system of reaction-diffusion equations

used to describe motion as well as chemical conversion, specify the general

initial and boundary conditions to describe FRAP experiments, and review

known solutions of special cases (17).

General model

The molecules may undergo S different binding reactions with other

substances. Each reaction is described by

Fþ Vi#koni

koffi

Bi; i ¼ 1.S: (1)

The unbound (free) fraction of molecules, F, binds with Vi vacant binding

sites to form the bound fraction of molecules, Bi. koni and koffi are the

corresponding association and dissociation rates in mol s–1 and s–1,

respectively.

The most general set of reaction-diffusion equations that describe one

unbound fraction and S bound fractions is

vcFvt

¼ DFV2cF �

XSi¼ 1

�koni cFcVi � koffi cBi

�(2a)

vcVivt

¼ DViV2cVi �

XSi¼ 1

�konicFcVi � koffi cBi

�(2b)

vcBivt

¼ DBiV2cBi þ koni cFcVi

� koffi cBi ; i ¼ 1.S; (2c)

whereV2 is the Laplacian operator, c represents the concentration of a given

molecule type at a certain time and place, and D is the diffusion coefficient.

Index i denotes the reaction as it is described in Eq. 1.

Equation 2 can be further simplified by assuming that the fluorescent

molecules are in equilibrium before photobleaching. The bleaching proce-

dure changes the number of fluorescent molecules (cF, cBi), whereas the

fraction of the nonfluorescent binding sites, cVi, is constant over time.

Therefore, we can eliminate Eq. 2b. The variable Vi in the remaining equa-

tions is constant over time. Thus, we define a new variable called the pseudo

association rate, k�oni , as

k�oni ¼ koni cVi : (3)

Thus, Eq. 2 reduces to

vcFvt

¼ DFV2cF �

XSi¼ 1

�k�oni cF � koffi cBi

�(4a)

vcBi

vt¼ DBi

V2cBi þ k�oni cF � koffi cBi ; i ¼ 1.S: (4b)

As mentioned above, the system (Eq. 2) is in steady state before

bleaching.

vcFvt

¼ vcBivt

¼ 0 if t ¼ 0: (5)

Since the total mass has to be conserved, it follows that

cF þXSi¼ 1

cBi ¼ 1: (6)

Hence, the equilibrium concentrations of free fraction, Feq, and bound frac-

tion, Beqi, are

vcBivt

¼ 0 0 Beqi ¼k�onikoffi

Feq (7a)

cF þXSi¼ 1

cBi¼ 1 0 Feq ¼ 1

1þ PSj¼ 1

k�onjkoffj

:

(7b)

To characterize the initial and boundary conditions, we divide the space

into two regions: 1), the bleaching spot area, a; and 2), the region outside

the bleaching spot, a. The initial conditions of concentrations inside the

bleaching spot, cin, are zero, whereas initial conditions of concentrations

outside this region, cout, are at equilibrium:

cinF ¼ 0 if t ¼ 0 (8a)

coutF ¼ Feq if t ¼ 0 (8b)

cinBi ¼ 0 if t ¼ 0 (8c)

coutBi¼ Beqi if t ¼ 0: (8d)

Biophysical Journal 100(5) 1178–1188

Page 3: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

1180 Mai et al.

A more general assumption is a constant initial value, q, inside the

bleaching spot whereby Eq. 8 converts to

cinF ¼ q � Feq if t ¼ 0 (9a)

coutF ¼ Feq if t ¼ 0 (9b)

cinBi ¼ q � Beqi if t ¼ 0 (9c)

coutBi¼ Beqi if t ¼ 0: (9d)

For simplicity, we assume that the bleaching spot is a radial area with

radius R, which yields the boundary conditions

vcinFvr

¼ 0 if r ¼ 0 (10a)

vcinBivr

¼ 0 i ¼ 1.S if r ¼ 0 (10b)

vcoutF

vr¼ 0 if r/N (10c)

vcoutBi

vr¼ 0 i ¼ 1.S if r/N (10d)

cinF ¼ coutF if r ¼ R (10e)

cinBi¼ coutBi

i ¼ 1.S if r ¼ R (10f)

vcinFvr

¼ vcoutF

vrif r ¼ R (10g)

vcinBivr

¼ vcoutBi

vri ¼ 1.S if r ¼ R: (10h)

Unlike the reactions (Eq. 1) where all bound states Bi depend directly on

the free fraction F, a more realistic scenario is to model reactions as a chain

Bi�1 þ Vi�1#koni

koffi

Bi ; i ¼ 1.S; (11)

where B0 represents the free molecular fraction, F.

This leads to modified differential equations and equilibrium concentra-

tions, respectively.

Beq0 ¼ Feq ¼ 1

1þPSi¼ 1

� Qij¼ 1

kon�j

koffj

� (12a)

Beqi ¼ Yi

j¼ 1

k�onjkoffj

!Feq ; i ¼ 1.S: (12b)

The initial and boundary conditions (Eqs. 8–10) are equal for both reaction

models.

Review of solutions for special cases

Different special cases for Eq. 4 with Eq. 8 and Eq. 10 have already been

solved and published (17). Since we need these results for comparison,

we briefly present the analytical results in the following sections.

Biophysical Journal 100(5) 1178–1188

Reaction Dominant Model. The first simplified scenario assumes that

diffusion is very fast compared to reaction. Thus, diffusion can be neglected,

and models that employ this mechanism are called reaction-dominant. For

the analysis of FRAP experiments, the solution of such a reaction-dominant

model was introduced by Sprague et al. (17). These authors assumed a zero

initial condition (Eq. 8) and treated models with one- and two-binding-state

solutions.Wegeneralize their approach to get the reaction-dominant solution

with S binding states.

The S reaction equations,

Fþ Vi#k�oni

koffi

Bi; i ¼ 1.S; (13)

yield the relationship

frapðtÞ ¼ 1�XSi¼ 1

Beqi e�koffi t; (14)

where frap(t) represents the recovery curve of the FRAP experiment and the

equilibrium concentration of the bound molecular fraction, Beqi, is given by

Eq. 7.

Mueller et al. (25) solved this model under the condition of a constant

initial value, q (Eq. 9). The general solution with S binding states is given by

frapðtÞ ¼ 1� ð1� qÞ �XSi¼ 1

Beqi e�koffi t: (15)

Reaction Diffusion Model with Single Diffusion. The second scenario

describes the case in which the free molecules, F, are moving with a diffu-

sion coefficient, DF, and the reaction products, Bi are immobile (DBi ¼ 0).

F|{z}DF

þVi#k�oni

koffi

Bi; i ¼ 1.S: (16)

The Laplace transformed solution dfrapðsÞ with an assumed zero initial

condition (Eq. 8) is given in Sprague et al. (17). The authors derive a solu-

tion for one- and two-binding-state models, which we extend to S binding

states:

dfrapðsÞ ¼ 1

s� Feq

sð1� 2K1ðqRÞI1ðqRÞÞ

��1þPS

i¼1

konisþ koffi

��XSi¼1

Beqi

sþ koffi; (17)

with

q2 ¼ s

DF

XSi¼ 1

konisþ koffi

!;

where R represents the radius of the circular bleaching spot and I1 and K1

are modified Bessel functions of the first and second kind. The equilibrium

concentrations of the free molecular fraction, Feq, and the bound molecular

fractions, Beqi, are given by Eq. 7.

The corresponding solution with a constant initial condition was calcu-

lated using the strategy described by Sprague et al. (17) and is given by

dfrapðsÞ ¼ 1s�PS

i¼ 1

ð1� qÞBeqi

sþ koffi

�ð1� qÞFeq

sð1� 2K1ðqRÞI1ðqRÞÞ �

XSi¼ 1

konisþ koffi

!;

(18)

with

Page 4: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

A General RD Model for FRAP Analysis 1181

q2 ¼ s

DF

XSi¼ 1

konisþ koffi

!:

The averaged fluorescence intensity within the bleaching spot (Eqs. 17

and 18) is still in the Laplace transformed form. The Laplace transform

is defined as dfrapðsÞ ¼Z N

0

e�st � frapðtÞ dt: (19)

Equations 17 and 18 have to be inverted to real time t. An analytical back

transform is not possible in closed form; therefore, the Stehfest algorithm

was used for numerical inversion (33,34).

Derivation of the general solution: Reaction Diffusion Modelwith Multiple Diffusion

After solving some simplified cases we now describe the derivation of the

general coupled reaction-diffusion equations (Eq. 4 with Eqs. 8 and 10). We

call this model full reaction-diffusion with multiple diffusion. This model

describes reactive coupled molecular fractions that could all be mobile.

We assume that both the free fraction, F, and the S bound fractions

B1.BS are moving diffusively. The motion is described by diffusion

coefficients for the free and bound molecular fractions, DF and

DB1.DBS

, respectively. The reactions are characterized by the association

rates k�oni and the dissociation rates koffi (i¼ 1,., S). We assume a chain of

reactions so that the free molecular fraction (F and B0, respectively)

converts only to B1. Molecules of B1 are able to dissociate to B0 or associate

to B2.

Bi�1|ffl{zffl}DBi�1

þVi #k�oni

koffi

Bi|{z}DBi

; i ¼ 1.S: (20)

First, we elucidate one-binding-state models and derive their solution.

The detailed derivation is given in the Supporting Material. Second, we

show that our solution is also applicable to the more general case of S

binding states.

Semi-analytical recovery function (1-binding-state). The general set of

coupled reaction-diffusion equations to describe a one-binding-state model

is

vcB0vt

¼ DB0V2cB0 � k�oncB0 þ koffcB1 (21a)

vcB1vt

¼ DB1V2cB1

þ k�oncB0 � koffcB1 ; (21b)

where B0 represents the free molecular fraction, F.

The Laplace transformed average fluorescence intensity inside the

bleaching spot is described by (for derivation, see Supporting Material)

dfrapðs�Þ ¼ q

sþ 2I1ðuÞK1ðuÞðK�s�þp�1Þðs��qÞð1�qÞ

ðK � 1Þs�ðp� qÞ�2I1ðvÞK1ðvÞðK � s� þ q� 1Þðs� � pÞð1� qÞ

ðK � 1Þs�ðp� qÞ(22)

with

u ¼ R

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffik�on � p

DF

rand v ¼ R

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffik�on � q

DF

r; (23)

where R represents the radius of the circular bleaching spot, q denotes the

initial fluorescence value inside the bleaching spot, and I1 andK1 are modi-

fied Bessel functions of the first and second kind. The original time variable

t* ¼ kon � t changed to the Laplace variable s* due to the transformation.

For the description of variables p, q, D, and K, see Supporting Material.

Since the solution is Laplace-transformed, we have to transform it back

to frap(t*) using the Stehfest algorithm (33,34).

Generalized semi-analytical recovery function (S-binding-states). In the

next step, the solution has to be generalized to S binding states. The matrix

notation of the Laplace transformed general differential equation set (Eq. 4)

is given by

�V2bcB� ¼

0BBBB@sþkon1DB0

�koff1DB0

0 . 0

�kon1DB1

sþkoff1þkon2DB1

�koff2DB1

. 0

« « « 1 «

0 0 0 .sþkoffSDBS

1CCCCA|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}

:¼A

�bcB�;

(24)

where ðV2bcBÞ and ðbcBÞ are vectors.First, the eigenvalues of A have to be derived. Let lB0

, lB1,., lBS

be the

Sþ 1 eigenvalues ofA andEV¼ {EV(lB0), EV(lB1

),., EV(lBS)} the matrix

of their eigenvectors. The Laplace transformed solution of Eq. 4 inside the

bleaching-spot area with constant initial values q (Eq. 9) is given by�bcinB� ¼ q

s

�Beq

� EV

0BB@a0 � I0

�rffiffiffiffiffiffilB0

p �þ b0 � K0

�rffiffiffiffiffiffilB0

p �a1 � I0

�rffiffiffiffiffiffilB1

p �þ b1 � K0

�rffiffiffiffiffiffilB1

p �«

aS � I0�rffiffiffiffiffiffilBS

p �þ bS � K0

�rffiffiffiffiffiffilBS

p �1CCA;

(25)

where ðbcinB Þ and (Beq) are vectors and the upper index, bcin, represents theconcentrations inside the bleaching spot.

Solving the set of equations with the constant initial values outside the

bleaching spot (Eq. 9) yields�bcoutB

�¼ 1

s

�Beq

� EV

0BB@g0 � I0

�rffiffiffiffiffiffilB0

p �þ d0 � K0

�rffiffiffiffiffiffilB0

p �g1 � I0

�rffiffiffiffiffiffilB1

p �þ d1 � K0

�rffiffiffiffiffiffilB1

p �«

gS � I0�rffiffiffiffiffiffilBS

p �þ dS � K0

�rffiffiffiffiffiffilBS

p �1CCA;

(26)

where ðbcoutB Þ and (Beq) are vectors and the upper index, bcout, represents theconcentrations outside the bleaching spot.

After determination of the coefficients ak, bk, gk, and dk by the given

boundary conditions (Eq. 10), we have to calculate the average fluorescence

of all molecular fractions inside the bleaching spot

dfrapðsÞ ¼ 1

pR2�Z2p0

ZR0

r �XSi¼ 0

bcinBidr d4: (27)

We used Eq. 22 to analyze recovery curves with one binding state and

Eq. 27 to analyze recovery curves with two binding states.

FRAP Inversion Algorithms

Our optimization problem is to find diffusion coefficients and reaction rates

within our solutions that fit a given recovery curve best.

Biophysical Journal 100(5) 1178–1188

Page 5: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

TABLE 1 Inital values and range of diffusion coefficients and

reaction rates

Initial value Minimum Maximum

DF 1.0 1.0$10�2 100.0

DBi1.0 1.0$10�2 100.0

koni 0.5 1.0$10�8 1.0

koffi 0.5 1.0$10�8 1.0

The values shown were used for parameter search by SA. The search space

is defined by the given range.

1182 Mai et al.

Preprocessing Algorithm: Adapted Prony’s method

To infer the maximal number of hidden binding sites, we used a preprocess-

ing step by adapting Prony’s algorithm (27,29,35). This preprocessing step

allowed us to limit the number of models to be fitted and to estimate the

effects of overfitting.

In the next paragraph, we describe the general strategy of Prony’s method

and how we made use of this strategy to deduce the maximal number of S.

Prony’s technique is based on exponential decay curves, in our case the loss

of bleached molecules inside the bleaching spot (loss(t) ¼ 1 – frap(t)). The

measured decay curve is approximated by

lossðtÞ ¼ f ðtÞzC1ea1t þ C2e

a2t þ.þ CSeaSt

¼ C1mt1 þ C2m

t2 þ.þ CSm

tS

; (28)

which can be rewritten as a set of equations,

C1 þ C2 þ.þ CS ¼ f0

C1m1 þ C2m2 þ.þ CSmS ¼ f1 (29)

«

C1mN�11 þ C2m

N�12 þ.þ CSm

N�1S ¼ fN�1;

if a set ofN equally spaced measurements is given. Since mi ¼ eai and Ci are

unknown, at least 2S equations are needed. The difficulty of solving the

problem is caused by the nonlinearity in mi. Therefore, the Eq. 29 set of

equations is linearized (28):

fS þ fS�1a1 þ fS�2a2 þ.þ f0aS ¼ 0

fSþ1 þ fSa1 þ fS�1a2 þ.þ f1aS ¼ 0 (30)

«

fN�1 þ fN�2a1 þ fN�3a2 þ.þ fN�S�1aS ¼ 0;

where the ai are determined and the mi are found as the roots of

mS þ a1mS�1 þ a2m

S�2 þ.þ aS�1mþ aS ¼ 0 (31)

Equation 29 then becomes a set of linear equations for Ci where the coef-

ficients are known. Ci can then be inferred by applying a least-squares

method.

Prony’s original suggestion was the least-squares algorithm

A ¼ XTX

��1XTY; (32)

where

X ¼

0BB@fS�1 fS�2 f0fS fS�1 f1« « «

fN�2 fN�3 fN�S�1

1CCAYT ¼ ð fS fSþ1 / fN�1 Þ

AT ¼ ð�a0 �a1 / �aS Þ;for the identification of ai in Eq. 30. Since simulations showed that this

strategy results in nonoptimal values of A, we used the optimized correlation

method by Sun et al. (35), introducing the following auxiliary matrix, Z:

Z ¼

0BB@fS�1þk fS�2þk fkfSþk fS�1þk f1þk

« « «fN�2þk fN�3þk fN�S�1þk

1CCA;

Biophysical Journal 100(5) 1178–1188

which is then applied to calculate the unknowns as follows

A ¼ ZTX

��1ZTY: (33)

The ai determined by this method can then be used to calculate the Ci as

described earlier.

Inversion Algorithm: Simulated Annealing

For inversion purposes, an objective function describing the difference

between the measured and the calculated data set has to be minimized.

We use the mean absolute error (MAE) as our objective function:

MAE ¼XNi¼ 1

jxi � xei j; (34)

in which N is the number of measured time points, xi represents the

measured recovery value at the ith time point, and xie is the estimated

recovery value at the ith time point.

Unlike the Gauss-Newton and Levenberg-Marquardt algorithms

commonly applied to fit data sets (17,30,31), we choose the SA technique

to solve this problem (32). SA is a heuristic optimization technique based

on the metropolis algorithm (36). The advantages of this algorithm are that

1. it can deal efficiently with cost or objective functions characterized by

quite arbitrary degrees of nonlinearities, discontinuities, and

stochasticity;

2. it can process quite arbitrary boundary conditions and constraints

imposed on these cost functions;

3. it can be implemented quite easily in comparison with other nonlinear

optimization algorithms;

4. it is independent of the initial parameter settings; and finally

5. it converges to the optimum solution, so that finding a near-optimum

solution is statistically guaranteed.

SA has the great benefit that results are not constrained by initial param-

eter settings, since it is known that initial values far from the true values

cause fitting failure by using the Gauss-Newton or the Levenberg-

Marquardt algorithm (17,37).

We take the same initial parameter settings and parameter ranges to

generate a feasible solution for every fitting (Table 1).

EXPERIMENTS

To test the reliability of our preprocessing and inversionalgorithm we created artificial data sets from our (semi)analytical solutions. The aim was to identify the correctmodel type as well as to estimate the model parameters.Subsequently, we applied our fitting algorithm to realFRAP data to prove that diffusion coefficients and reactionrates can be robustly identified from real (i.e., noisy) datasets.

Page 6: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

TABLE 3 Number of binding sites estimated by Prony’s

method

M1 M2 M3

No noise 1S 0% 0% 0%

2S 100% 100% 0%

3S 0% 0% 100%

>3S 0% 0% 0%

With noise 1S 2% 14% 0%

2S 87% 84% 56%

3S 11% 2% 44%

>3S 0% 0% 0%

A data set without noise and 100 noisy data sets (n ~ N (0.00, 0.03)) with

different underlying model types (M1, reaction-dominant (2S); M2, full

reaction-diffusion with single diffusion (2S); and M3, full reaction-

diffusion with multiple diffusion (2S)) were analyzed by Prony’s method.

Entries in bold highlight the most likely number of binding sites S.

A General RD Model for FRAP Analysis 1183

Artificial FRAP data

Preprocessing Algorithm: Setup and Results

The parameter values were chosen to correspond to theexperimental setup and the achieved values. For testingthe preprocessing step of the adapted Prony’s method, weused artificial datasets with parameter values correspondingto the experimental setup and the achieved values (Table 2).To account for realistic experimental data, we added aGaussian-distributed noise value (n ~ N (0.00, 0.03)). Insummary, we tested six different model functions: threemodel types each with and without added noise. Since noiseis a random process, we used 100 different sample data setsfor the model functions. The model types chosen were reac-tion-dominant (M1), full reaction-diffusion with singlediffusion (M2), and full reaction-diffusion with multiplediffusion (M3), all with S ¼ 2.

Prony’s original method was designed to determine thenumber of molecular fractions that are coupled by a reac-tion-dominant scheme. For the reaction-dominant modeltype with and without noise, the number of binding sites,S, was estimated correctly (M1; see Table 3). Consideringonly one diffusive fraction does not influence the correctprediction. However, increasing the number of diffusivecomponents leads to an overestimation of S. As expected,the noise level imposed on the artificial data sets influencedthe predictive power. However, in every case, the methodpredicted either the correct number or one additional S.The results for all model types with and without noise areshown in Table 3.

Inversion Algorithm: Setup and Results

To test our inversion routine, we tested the model functionsusing a SA algorithm applied to artificial datasets withsettings specified in Table 4. To allow for a more realisticdescription of measurements, different noise values wereadded. A low noise value, nL ~ N (0.00, 0.01), and a highnoise value, nH ~ N (0.00, 0.03), were taken, leading tonine artificial datasets in total (Fig. 2). They were abbrevi-ated with Mk

s, where k indicates the model type: reaction-dominant model (M1 (Eq. 14)), full reaction-diffusionmodel with single diffusion (M2 (Eq. 17)), and full reac-tion-diffusion model with multiple diffusion (M3 (Eq. 22))all with S ¼ 1. s represents the standard deviation of theadded Gaussian-distributed noise signal. Every data set

TABLE 2 Artificial data set values for testing the

preprocessing step of the adapted Prony’s method

Radius of bleaching spot R ¼ 2 (LU)

Simulated time steps N ¼ 250 (TU)

Reaction rates kon1 ¼ 0.9, kon2 ¼ 0.9

koff1 ¼ 0.9, koff2 ¼ 0.3

Diffusion coefficients DF ¼ 10.0, DB1¼ 5:0

Unit of diffusion coefficients ½LU2=TU�Unit of association and dissociation rate ½1=TU�

was fitted against the solutions of the three model typesmentioned above.

The error-function value was determined as the sum ofabsolute differences between fitted functions and data sets(Eq. 34). As expected, the error-function values for allmodel types increased along with the noise level. Evenwith a high noise level the model type was determined reli-ably (Table 5). In some cases, differentiation between modeltypes was based on marginal differences of the best error-function values. To prove that even these differences weresignificant, we investigated the distribution of the error-function values of 500 simulated-annealing runs. Plottingthe absolute frequencies of error-function values clearlydistinguished the investigated model types. The correctmodel function always shows an accumulation of the lowesterror-function values (Fig. 3). These results demonstrate thefact that even small differences in least error-function valuesare significant.

We next tested the ability of the model to correctly predictparameters. As an example, we compared the prediction ofthe reaction rates kon and koff, since these are the onlyparameters present in all three model types. Similar to themodel-type prediction, the noise level of the artificial datasets influenced the prediction power of the parameter values.Fig. 4 shows the mean and variance of these parametersdetermined by the best 100 (out of 500) SA runs with leasterror-function values. In summary, not only the correctunderlying model type but also the correct diffusion coeffi-cients and reaction rates were determined. For furtheranalysis of the artificial data set and information aboutcomputational expenses of SA, see the Supporting Material.

TABLE 4 Artificial data set used with SA algorithm to test

inversion routine

Radius of bleaching spot R ¼ 15 [LU]

Simulated time steps N ¼ 100 [TU]

Reaction rates kon ¼ 0.3, koff ¼ 0.05

Diffusion coefficients DF ¼ 10.0, DB1¼ 1:0

Unit of diffusion coefficients ½LU2=TU�Unit of association- and dissociation rate ½1=TU�

Biophysical Journal 100(5) 1178–1188

Page 7: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

FIGURE 2 Artificial data sets. Calculation of

the model function values (M1, M2, M3) with

different noise levels (s). (A) s ¼ 0.00, no noise.

(B) s ¼ 0.01, low noise level. (C) s ¼ 0.03, high

noise level.

1184 Mai et al.

Real FRAP data

Setup

The murine hepatoma cells Tao BpRc1, deficient in endog-enous aryl hydrocarbon receptor (AhR), were stably trans-fected with a green-fluorescent-protein (GFP)-labeled AhRconstruct (38) under tetracycline control. Cells were grownin phenol-free DMEM containing 10% fetal bovine serumand 2 mM L-alanyl-L-glutamine and cultured in 35 mmIbidi m-dishes. To induce expression of GFP-AhR, cellswere taken off tetracycline 24 h before exposure. To induceAhR translocation, cells were exposed to 50 nM Benzo(a)pyrene (BaP) for 15 min.

The FRAP experiments were performed on a Zeiss LSM510 META confocal microscope (Carl Zeiss, Jena,Germany) with a 100�/1.4 NA oil-immersion objective.Bleaching was performed with a circular spot (radius

TABLE 5 Error-function values of analyzed artificial data sets

Fitted model type

Model 1 Model 2 Model 3

M10.00 0.00000 1.92217 2.07131

M10.01 0.78581 2.09649 2.24620

M10.03 2.11161 3.03402 3.16262

M20.00 1.50488 0.00885 0.06056

M20.01 1.70500 0.78243 0.78933

M20.03 2.87615 2.23449 2.35603

M30.00 2.17175 0.03437 0.00092

M30.01 2.26306 0.77783 0.77610

M30.03 2.97956 2.48055 2.45850

Artificial data sets with different noise levels (s1 ¼ 0.00, s2 ¼ 0.01, and

s3 ¼ 0.03) were fitted by three model functions. Least error-function values

are printed in bold.

Biophysical Journal 100(5) 1178–1188

1.12305 mm) using the 488- and 514-nm lines from an argonlaser operating at 74% laser power. A single iteration wasused for the bleach pulse. Five prebleach images were takenand the fluorescence recovery was monitored in 83.2-msintervals. During all FRAP experiments, cells were keptat 37�C using a heated stage plate (Carl Zeiss). In total,50 FRAP experiments were performed in the nucleus. Theraw image data were used to extract the fluorescence-recovery curves. Afterward, each recovery curve wasdouble-normalized using the prebleach images as well astwo reference areas, as described by Phair et al. (8) (seealso Supporting Material). The average recovery shown inFig. 5 was calculated by taking the mean of all 50 individualrecovery curves.

To estimate the constant initial value, we extracted thebleaching profile out of the raw data and fitted a Gaussianfunction. We used this function to determine the initial valueof q as described by Hinow et al. (39) and readjusted thebleaching-spot radius according to the procedure of Muelleret al. (25).

For fitting, we used the model functions with the esti-mated constant initial value, q, and the adapted bleaching-spot radius (Eqs. 15, 18, and 22).

As a reference measurement, we performed FRAP exper-iments in Tao BpRc1 cells expressing GFP only.

Results

To infer the correct underlying model type for the movementof GFP-tagged AhR inside the nucleus, we compared allavailable model types. Data were described best by a fullreaction-diffusion model with multiple diffusion and twobinding sites (Table 6 and Fig. 5). As already describedfor the artificial data, the distribution of the error-function

Page 8: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

FIGURE 3 Error-value histogram of analyzed

artificial data sets. (A–I) Artificial data sets were

generated by three different model types to which

different noise levels were added (0.00–0.03).

The distributions of error-function values of

500 SA runs are shown for the three model types

(Model 1, black bars; Model 2, gray bars; and

Model 3, white bars).

A General RD Model for FRAP Analysis 1185

values confirmed the model type determined by the leasterror-function value (data not shown). Deduced from themodel type, three molecular fractions were predicted forthe GFP-AhR data set: one fraction is moving diffusivelywith a diffusion coefficient of DB0

¼ 5.1 mm2 s�1; this frac-tion then converts to a second fraction with a diffusion coef-ficient of DB1

¼ 3.3 mm2 s�1, and this slower fraction finally

FIGURE 4 Robustness of estimated reaction parameters. Box plots are

shown for the 100 best of the 500 SA runs (mean5 SD). Black dots repre-

sent the parameters yielding the least error-function values. The parameter

value used for artificial data set creation is represented by the dotted line.

becomes immobile (DB2¼ 0.1 mm2 s�1; due to the model

restrictions, a diffusion coefficient equal to zero is notdefined). Calculated from the fitted reaction rates, thepercentage of the fractions are Beq0 ¼ 93%, Beq1 ¼ 1%,and Beq2 ¼ 6% for the immobile fraction (Eq. 7).

Although the models became increasingly complex, alldeduced diffusion and reaction parameters were consistentamong model types (Table 6). Verifying the determinedmodel type, Prony’s method yielded a maximal number ofbinding partners, S ¼ 3. For further analysis of the realdata set, see the Supporting Material.

FIGURE 5 Graphical comparison of experimental FRAP data with

model functions. Dots represent the average GFP-AhR recovery from 50

independent FRAP experiments (nucleus, 15min, 50 nM BaP) and lines

represent model functions. It can be seen clearly that the reaction-diffusion

model with multiple diffusion (2S) with the least error-function value fits

the experimental data best.

Biophysical Journal 100(5) 1178–1188

Page 9: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

TABLE 6 Fitted parameters of models with least error

function values

DF DB1 Feq Beq1

DB2 Beq2

DB3Beq3

Full reaction-diffusion with multiple

diffusion (1S)

4.930 0.101 0.882 0.118

— —

Full reaction-diffusion with single

diffusion (2S)

4.997 0.000 0.951 0.046

0.000 0.004

Full reaction-diffusion with multiple

diffusion (2S)

5.067 3.340 0.934 0.007

0.091 0.060

— —

Analysis of nuclear FRAP data on GFP-AhR yielded three models with

comparable error-function values. The least error-function value was ob-

tained by the reaction-diffusion model with multiple diffusion (2S).

(Feq ¼ Beq0 and F ¼ B0).

1186 Mai et al.

DISCUSSION

FRAP is a powerful technique to investigate the dynamicbehavior of proteins in living cells. Mathematical modelingof FRAP data allows determination of dissociation and asso-ciation rates, distribution of mobile and immobile fractions,and corresponding diffusion coefficients. A number of sim-plified models that describe motion of reactively coupledfluorescent molecules observed by FRAP have been de-scribed (17–21). However, all of these existing approachesinclude a priori assumptions to allow for the determinationof an analytical solution. To circumvent this bias, we estab-lished a generalized reaction-diffusion model that comprisesa flexible number of reacting and diffusing fractions. Morespecifically, we impose constraints on neither model typenor parameters.

Generalizing the model type obviously yields a highernumber of parameters. To address this issue, we introducea preliminary approximation of the number of acting molec-ular fractions by applying the adapted Prony’s method(27,29,35). This method was tested using artificial as wellas real data sets and proved to identify the correct numberof binding sites, either S or S þ 1. Since the Prony’s methodis based on a reaction-dominant scheme, the introduction ofdiffusive components results in an overestimation of bindingsites. Therefore, the Prony’s method always deduces theupper limit of molecular fractions, i.e., parameters thatcan be fitted reliably. Therefore, we can rule out that thedecreasing error-function values of our more complexmodels are due to overfitting effects. In addition, for futurestudies, the deduced number S could be used to limit themodels to be fitted.

To infer optimal parameter sets, local search algorithmslike Levenberg-Marquardt are commonly used (17,30,31).However, Sprague and colleagues (17,37) showed that theLevenberg-Marquardt algorithm is very sensitive to thechoice of initial parameters. To allow for a global searchindependent of these initial settings, we employed the SA

Biophysical Journal 100(5) 1178–1188

strategy (32). The performance test of the SA algorithmusing artificial data sets demonstrated that SA indeed isable to predict the correct model type reliably. Althoughthe variance of estimated parameter values (e.g., diffusioncoefficients) increased with the complexity of the model,the prediction of the parameters was still satisfactory withrespect to their variance.

We demonstrated the consistency of our new approach byfitting existing (17) and new models to nuclear FRAPmeasurement in a murine hepatoma cell line stably trans-fected with a GFP-labeled AhR construct. AhR is a solublecytoplasmic transcription factor. After ligand binding, thereceptor-ligand complex (AhR/L) translocates into thenucleus where it associates with its cofactor, AhR nucleartranslocator (ARNT). Association of ARNT with AhR/L isnecessary for binding to so-called xenobiotic-responseelements (XREs) to regulate transcription.

For nuclear FRAP data on GFP-AhR, our model outper-formed previous models (11) and suggested the existenceof three different molecular fractions and diffusion coeffi-cients. In correspondence with common knowledge ofAhR signaling (40–43), these molecular fractions shouldrepresent AhR/L, AhR/L/ARNT, and AhR/L/ARNT/XRE.

The relation D f m–1/3 (17), where D is the diffusioncoefficient and m is the mass of a molecule, allows us todeduce the mass for a particular diffusion coefficient orvice versa, as long as a reference measurement is available.An empty vector expressing the fluorophore tagging theprotein of interest is usually used for this purpose(11,17,39). The fitted GFP diffusion coefficient of ourin vitro system correlates very well with published valuesto date (17,44–46) and helped us to identify the followingmolecular fractions: a diffusion coefficient of GFP-Ahr/L(124 kDa) equal to 4.8 mm2 s�1, and a diffusion coefficientof GFP-AhR/L/ARNT (211 kDa) equal to 4.0 mm2 s�1.Since all fitted diffusion models are restricted to D > 0,we consider the slowest fraction, with a diffusion coefficientclose to zero, as GFP-AhR/L/ARNT/XRE. These estimatedvalues are in close agreement to the fitted values (D1 ¼5.1 mm2 s�1 and D2 ¼ 3.3 mm2 s�1) inside the nucleus.

In addition, our estimated parameters are similar to thosereported for the glucocorticoid receptor, which has a masscomparable to that of AhR (25).

Our new approach shows that intracellular molecularmobility can only be described adequately by allowing formultiple reaction-diffusion processes, as shown by ourapplication to GFP-AhR FRAP data. Our general reaction-diffusion model performed significantly better than the stan-dard types of reaction-diffusion model used previously.Coming back to the question we posed at the beginning,‘‘Are assumptions of the model type necessary in reaction-diffusion modeling?’’, we give the provocative answer, itmay be not only not necessary but also too restrictive todescribe the processes sufficiently, since we may use anassumption that is not justified. Therefore, we argue that

Page 10: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

A General RD Model for FRAP Analysis 1187

optimizing parameter sets for predetermined model types istoo restrictive to describe biological processes sufficiently.

SUPPORTING MATERIAL

Derivation of new solution, further analysis of artificial and real FRAP data,

and computational expenses are available at http://www.biophysj.org/

biophysj/supplemental/S0006-3495(11)00124-X.

We thank Luis Samaniego and Rohini Kumar for helpful discussions and

advice concerning the application of the simulated annealing technique.

The authors are indebted to Anke Hildebrandt (Helmholtz Centre for Envi-

ronmental Research, Zentrum fur Umweltforschung) for helpful comments

and critical reading.

This project was supported by the Helmholtz Alliance on Systems Biology.

We are deeply indebted to JimMcNally (National Cancer Institute/National

Institutes of Health, Bethesda, Maryland) for critical comments and helpful

discussion to improve the initial conditions.

REFERENCES

1. Dundr, M., U. Hoffmann-Rohrer, ., T. Misteli. 2002. A kineticframework for a mammalian RNA polymerase in vivo. Science.298:1623–1626.

2. Aizer, A., and Y. Shav-Tal. 2008. Intracellular trafficking and dynamicsof P bodies. Prion. 2:131–134.

3. Shav-Tal, Y., X. Darzacq, ., R. H. Singer. 2004. Dynamics of singlemRNPs in nuclei of living cells. Science. 304:1797–1800.

4. Elsner, M., H. Hashimoto, ., M. Weiss. 2003. Spatiotemporaldynamics of the COPI vesicle machinery. EMBO Rep. 4:1000–1004.

5. Petrasek, Z., M. Krishnan, ., P. Schwille. 2007. Simultaneous two-photon fluorescence correlation spectroscopy and lifetime imaging ofdye molecules in submicrometer fluidic structures. Microsc. Res.Tech. 70:459–466.

6. Bacia, K., S. A. Kim, and P. Schwille. 2006. Fluorescence cross-correlation spectroscopy in living cells. Nat. Methods. 3:83–89.

7. Bacia, K., and P. Schwille. 2007. Fluorescence correlation spectros-copy. Methods Mol. Biol. 398:73–84.

8. Phair, R. D., S. A. Gorski, and T. Misteli. 2004. Measurement ofdynamic protein binding to chromatin in vivo, using photobleachingmicroscopy. Methods Enzymol. 375:393–414.

9. Phair, R. D., P. Scaffidi,., T. Misteli. 2004. Global nature of dynamicprotein-chromatin interactions in vivo: three-dimensional genomescanning and dynamic interaction networks of chromatin proteins.Mol. Cell. Biol. 24:6393–6402.

10. Koster, M., T. Frahm, and H. Hauser. 2005. Nucleocytoplasmic shut-tling revealed by FRAP and FLIP technologies. Curr. Opin. Biotechnol.16:28–34.

11. Braga, J., J. G. McNally, and M. Carmo-Fonseca. 2007. A reaction-diffusion model to study RNA motion by quantitative fluorescencerecovery after photobleaching. Biophys. J. 92:2694–2703.

12. Carrero, G., E. Crawford, ., M. J. Hendzel. 2004. Quantification ofprotein-protein and protein-DNA interactions in vivo, using fluores-cence recovery after photobleaching. Methods Enzymol. 375:415–442.

13. Stavreva, D. A., and J. G. McNally. 2004. Fluorescence recovery afterphotobleaching (FRAP) methods for visualizing protein dynamics inliving mammalian cell nuclei. Methods Enzymol. 375:443–455.

14. Giese, B., C.-K. Au-Yeung, ., G. Muller-Newen. 2003. Long termassociation of the cytokine receptor gp130 and the Janus kinase Jak1revealed by FRAP analysis. J. Biol. Chem. 278:39205–39213.

15. Howell, B. J., D. B. Hoffman,., E. D. Salmon. 2000. Visualization ofMad2 dynamics at kinetochores, along spindle fibers, and at spindlepoles in living cells. J. Cell Biol. 150:1233–1250.

16. Howell, B. J., B. Moree, ., E. D. Salmon. 2004. Spindle checkpointprotein dynamics at kinetochores in living cells. Curr. Biol. 14:953–964.

17. Sprague, B. L., R. L. Pego, ., J. G. McNally. 2004. Analysis ofbinding reactions by fluorescence recovery after photobleaching.Biophys. J. 86:3473–3495.

18. Axelrod, D., D. E. Koppel, ., W. W. Webb. 1976. Mobility measure-ment by analysis of fluorescence photobleaching recovery kinetics.Biophys. J. 16:1055–1069.

19. Bulinski, J. C., D. J. Odde, ., C. M. Waterman-Storer. 2001. Rapiddynamics of the microtubule binding of ensconsin in vivo. J. CellSci. 114:3885–3897.

20. Goldstein, B., R. G. Posner, ., B. Baird. 1989. Competition betweensolution and cell surface receptors for ligand. Dissociation of haptenbound to surface antibody in the presence of solution antibody.Biophys. J. 56:955–966.

21. Soumpasis, D. M. 1983. Theoretical analysis of fluorescence photo-bleaching recovery experiments. Biophys. J. 41:95–97.

22. Glockle, W. G., T. Mattfeld, and T. F. Nonnenmacher. 1994. Anoma-lous diffusion and angle-dependency in the theory of fluorescencerecovery after photobleaching. In Fractals in Biology and Medicine,Vol. I. T. F. Nonnenmacher, G. A. Losa, and E. R. Weibel, editors. Bir-khauser, Basel. 363–371.

23. Saxton, M. J. 2001. Anomalous subdiffusion in fluorescence photo-bleaching recovery: a Monte Carlo study. Biophys. J. 81:2226–2240.

24. Saxton, M. J. 2007. A biological interpretation of transient anomaloussubdiffusion. I. Qualitative model. Biophys. J. 92:1178–1191.

25. Mueller, F., P. Wach, and J. G. McNally. 2008. Evidence for a commonmode of transcription factor interaction with chromatin as revealed byimproved quantitative fluorescence recovery after photobleaching.Biophys. J. 94:3323–3339.

26. Sadegh Zadeh, K., and H. J. Montas. 2010. A class of exact solutionsfor biomacromolecule diffusion-reaction in live cells. J. Theor. Biol.264:914–933.

27. de Prony, R. 1795. Essai experimentale et analytique. J. Ecol. Polytech.1:24–76.

28. Hildebrand, F. B. 1974. Introduction to Numerical Analysis. Dover,New York.

29. Sun, T., Z. Y. Zhang,., A. W. Palmer. 1998. Quasidistributed fluores-cence-based optical fiber temperature sensor system. Rev. Sci. Instrum.69:146–151.

30. Elowitz, M. B., M. G. Surette,., S. Leibler. 1999. Protein mobility inthe cytoplasm of Escherichia coli. J. Bacteriol. 181:197–203.

31. Sadegh Zadeh, K., H. J. Montas, and A. Shirmohammadi. 2006.Identification of biomolecule mass transport and binding rate parame-ters in living cells by inverse modeling. Theor. Biol. Med. Model. 3:36.

32. Kirkpatrick, S., C. D. Gelatt, Jr., and M. P. Vecchi. 1983. Optimizationby simulated annealing. Science. 220:671–680.

33. Stehfest, H. 1970. Algorithm 368: numerical inversion of Laplacetransforms. Commun. ACM. 13:47–49 [D5].

34. Stehfest, H. 1970. Remark on algorithm 368: Numerical inversion ofLaplace transforms. Commun. ACM. 13:624.

35. Sun, T., Z. Y. Zhang,., A. W. Palmer. 1997. Analysis of double expo-nential fluorescence decay behavior for optical temperature sensing.Rev. Sci. Instrum. 68:58–63.

36. Metropolis, N., A. W. Rosenbluth, ., E. Teller. 1953. Equation ofstate calculations by fast computing machines. J. Chem. Phys.21:1087–1092.

37. Sprague, B. L., F. Muller,., J. G. McNally. 2006. Analysis of bindingat a single spatially localized cluster of binding sites by fluorescencerecovery after photobleaching. Biophys. J. 91:1169–1191.

38. Elbi, C., T. Misteli, and G. L. Hager. 2002. Recruitment of dioxinreceptor to active transcription sites. Mol. Biol. Cell. 13:2001–2015.

39. Hinow, P., C. E. Rogers, ., E. DiBenedetto. 2006. The DNA bindingactivityof p53displays reaction-diffusionkinetics.Biophys. J.91:330–342.

Biophysical Journal 100(5) 1178–1188

Page 11: Are Assumptions about the Model Type Necessary in Reaction-Diffusion Modeling? A FRAP Application

1188 Mai et al.

40. Hankinson, O. 1995. The aryl hydrocarbon receptor complex. Annu.Rev. Pharmacol. Toxicol. 35:307–340.

41. Holmes, J. L., and R. S. Pollenz. 1997. Determination of aryl hydro-carbon receptor nuclear translocator protein concentration and sub-cellular localization in hepatic and nonhepatic cell culture lines:development of quantitative Western blotting protocols for calculationof aryl hydrocarbon receptor and aryl hydrocarbon receptor nucleartranslocator protein in total cell lysates. Mol. Pharmacol. 52:202–211.

42. Song, Z., and R. S. Pollenz. 2002. Ligand-dependent and independentmodulation of aryl hydrocarbon receptor localization, degradation, andgene regulation. Mol. Pharmacol. 62:806–816.

43. Song, Z., and R. S. Pollenz. 2003. Functional analysis of murine arylhydrocarbon (AH) receptors defective in nuclear import: impact on

Biophysical Journal 100(5) 1178–1188

AH receptor degradation and gene regulation. Mol. Pharmacol.

63:597–606.

44. Braga, J., J. M. Desterro, and M. Carmo-Fonseca. 2004. Intracellular

macromolecular mobility measured by fluorescence recovery after

photobleaching with confocal laser scanning microscopes. Mol. Biol.

Cell. 15:4749–4760.

45. Yokoe, H., and T. Meyer. 1996. Spatial dynamics of GFP-tagged

proteins investigated by local fluorescence enhancement. Nat. Biotech-

nol. 14:1252–1256.

46. Houtsmuller, A. B., S. Rademakers,., W. Vermeulen. 1999. Action of

DNA repair endonuclease ERCC1/XPF in living cells. Science. 284:

958–961.