+ All Categories
Home > Documents > Electronic Energy Transfer within Asymmetric Pairs of

Electronic Energy Transfer within Asymmetric Pairs of

Date post: 25-Feb-2022
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
42
Electronic Energy Transfer within Asymmetric Pairs of Fluorophores: Partial Donor-Donor Energy Migration (PDDEM) Stanislav Kalinin Department of Chemistry: Biophysical Chemistry Umeå University Umeå Sweden 2004
Transcript
Page 1: Electronic Energy Transfer within Asymmetric Pairs of

Electronic Energy Transfer within Asymmetric

Pairs of Fluorophores: Partial Donor-Donor

Energy Migration (PDDEM)

Stanislav Kalinin

Department of Chemistry: Biophysical Chemistry

Umeå University

Umeå Sweden 2004

Page 2: Electronic Energy Transfer within Asymmetric Pairs of

ii

Department of Chemistry: Biophysical Chemistry

Umeå University

SE-901 87 Umeå, Sweden

Copyright © 2004 by Stanislav Kalinin

ISBN 91-7305-765-7

Printed by VMC, KBC, Umeå University, Umeå, 2004

Page 3: Electronic Energy Transfer within Asymmetric Pairs of

iii

ABSTRACT A kinetic model of electronic energy migration within pairs of photophysically non-identical fluorophores has been developed. The model applies to fluorescent groups that exhibit different photophysical and spectral properties when attached to different positions in a macromolecule. The energy migration within such asymmetric pairs is partially reversible, which leads to the case of partial donor-donor energy migration (PDDEM). The model of PDDEM is an extension of the recently developed donor-donor energy migration model (DDEM, F. Bergström et al, PNAS 96 (1999) 12477), and applies to quantitative measurements of energy migration rates and distances within macromolecules. One important distinction from the DDEM model is that the distances can be obtained from fluorescence lifetime measurements. A model of fluorescence depolarisation in the presence of PDDEM is also presented. To experimentally test the PDDEM approach, different model systems were studied. The model was applied to measure distances between rhodamine and fluorescein groups within on-purpose synthesised molecules that were solubilised in lipid bilayers. Moreover, distances were measured between BODIPY groups in mutant forms of the plasminogen activator inhibitor of type 2 (PAI-2). Measurements of both the fluorescence intensity decays and the time-resolved depolarisation were performed. The obtained distances were in good agreement with independent determinations. Finally, the PDDEM within pairs of donors is considered, for which both donors exhibit a nonexponential fluorescence decay. In this case it turns out that the fluorescence relaxation of a coupled system contains distance information even if the photophysics of the donors is identical. It is also demonstrated that the choice of relaxation model has a negligible effect on the obtained distances. The latter conclusion holds also for the case of donor-acceptor energy transfer. Keywords: fluorescence resonance energy transfer (FRET), donor-donor energy migration (DDEM), homotransfer, fluorescence relaxation, lifetimes, time-resolved fluorescence anisotropy, time-correlated single photon counting, distance measurements, protein structure.

Page 4: Electronic Energy Transfer within Asymmetric Pairs of

iv

CONTENTS

ABBREVIATIONS v

LIST OF PAPERS vi

1 INTRODUCTION 1

2 BACKGROUND 2 2.1 Introduction to fluorescence 2 2.2 Experimental techniques 4 2.3 Fluorescence anisotropy 6 2.4 Donor-acceptor energy transfer and distance measurements 7 2.5 Donor-donor energy migration (DDEM) 10

3 PARTIAL DONOR-DONOR ENERGY MIGRATION

(PDDEM) 13 3.1 The PDDEM model 13 3.2 PDDEM anisotropy 14

4 RESULTS AND DISCUSSION 16 4.1 Simulations of PDDEM 16 4.2 PDDEM within asymmetrically quenched bichromophoric molecules 18 4.3 PDDEM between ionic forms of fluorescein 19 4.4 PDDEM in proteins 23 4.5 The effect of PDDEM on nonexponential relaxation

of identical donors 27 4.6 The modelling of nonexponential fluorescence relaxation

in FRET and PDDEM 28

5 CONCLUSIONS AND FUTURE PERSPECTIVES 32

6 ACKNOWLEDGEMENTS 33

7 REFERENCES 34

Page 5: Electronic Energy Transfer within Asymmetric Pairs of

v

ABBREVIATIONS A acceptor of electronic energy a.u. arbitrary units D donor of electronic energy DDEM donor-donor energy migration EFT extended Förster theory F(t) time-dependent fluorescence intensity FRET fluorescence resonance energy transfer λ wavelength PDDEM partial donor-donor energy migration R the distance between interacting groups R0 Förster radius r(t) time-dependent fluorescence anisotropy τ fluorescence lifetime TCSPC time-correlated single-photon counting BODIPY 4,4-difluoro-4-bora-3a,4a-diaza-s-indacene and its derivatives DOPC 1,2-dioleoyl-sn-glycero-3-phosphocholine DMPC 1,2-dimyristoyl-sn-glycero-3-phosphocholine FlC16 5-(hexadecanoylaminomethyl) fluorescein FlC32Fl dotriacontanedioic acid bis(fluoresceinyl-5-methylamide) NDBY 1,3,5,7-tetramethyl-2-iodoacetamide-BODIPY PAI-2 plasminogen activator inhibitor of type 2 RhC16 Rhodamine 101 octadecyl ester RhC32Rh 1,32-dihydroxy-dotriacontane-bis-(Rhodamine 101) ester SBDY 5,7-dimethyl-3-methyliodoacetamide-BODIPY

Page 6: Electronic Energy Transfer within Asymmetric Pairs of

vi

LIST OF PAPERS This thesis is based on the following papers.

I. Kalinin, S.V., Molotkovsky, J.G., and Johansson, L.B.-Å.

Partial Donor-Donor Energy Migration (PDDEM) as a Fluorescence

Spectroscopic Tool for Measuring Distances in Biomacromolecules.

Spectrochim. Acta A, 58 (2002) 1087-1097.

II. Kalinin, S., Molotkovsky, J.G., and Johansson, L.B.-Å.

Distance Measurements Using Partial Donor-Donor Energy Migration within

Pairs of Fluorescent Groups in Lipid Bilayers.

J. Phys. Chem. B, 107 (2003) 3318-3324.

III. Isaksson, M., Kalinin, S., Lobov, S., Wang, S., Ny, T., and Johansson, L.B.-Å.

Partial Donor-Donor Energy Migration (PDDEM): A Novel Fluorescence

Method for Internal Protein Distance Measurements.

Phys. Chem. Chem. Phys., 6 (2004) 3001-3008.

IV. Kalinin, S., and Johansson, L.B.-Å.

Energy Migration and Transfer Rates are Invariant to Modeling the

Fluorescence Relaxation by Discrete and Continuous Distributions of

Lifetimes.

J. Phys. Chem. B, 108 (2004) 3092-3097.

Page 7: Electronic Energy Transfer within Asymmetric Pairs of

1

1 INTRODUCTION The first reported observation of fluorescence is dated back to 1565. Today, fluorescence spectroscopy includes several techniques, which are extensively used in biosciences. The advantage of fluorescence spectroscopy is its unmatched sensitivity, which enables experiments even at the level of single molecules. The time-resolution of modern fluorescence techniques is on the timescale of femtoseconds. In combination with advanced imaging techniques, fluorescence methods allow real-time visualisation of biological processes. An overview of the principles and applications of fluorescence is found in specialised textbooks [1, 2]. Frequently, fluorescence spectroscopy is used as a method for inter- and intramolecular distance measurements. Such experiments are most often associated with the quantitative measurements of electronic energy transfer rates between a donor and an acceptor group [3] or energy migration rates between two identical donors [4-6]. Unlike X-ray and NMR techniques, fluorescence spectroscopic methods provide distance information in the range of ~ 10 – 100 Å, which is typically comparable with the size of proteins. This thesis aims at investigating the process of energy migration within pairs of photophysically non-identical fluorophores, hereinafter referred to as donors. Typically this is the case of chemically identical fluorophores attached to different positions in a macromolecule, e.g. in a protein. Depending on the local physico-chemical properties, the donors may exhibit different photophysical and spectral properties. As a result, the energy migration within such pairs becomes only partially reversible, and therefore we refer to this case as partial donor-donor energy migration (PDDEM). Here we present a theoretical model of PDDEM, which is first of all intended for quantitative distance measurements within biomacromolecules and supramolecular assemblies. The PDDEM model is tested on several model systems and applied for distance measurements in proteins.

Page 8: Electronic Energy Transfer within Asymmetric Pairs of

2

2 BACKGROUND 2.1 Introduction to fluorescence Luminescence is the emission of photons from electronically excited states [1, 2, 7]. Emission is also referred to as fluorescence or phosphorescence, depending on whether it corresponds to a spin-allowed or a spin-forbidden transition, respectively [1, 7]. The absorption and emission of light is conveniently illustrated by the Jablonski diagram (Fig. 2.1). This diagram displays possible intramolecular processes that may take place in the singlet ground state (S0), the excited singlet states (S1, S2, …) as well as in the triplet state (T1). For organic molecules, the energies of electronic excitation are typically within a few eV. This means that radiative transitions are observed in the ultraviolet and/or visible region (~ 200–800 nm). Moreover, according to the Boltzmann distribution, practically all the molecules are in the electronic ground state at room temperature, unless excited by light absorption.

Figure 2.1. The Jablonski diagram. Absorption (S0→S1 and S0→S2), fluorescence (S1→S0) and

phosphorescence (T1→S0) are indicated by solid arrows. Radiationless transitions are

illustrated by dashed arrows.

Page 9: Electronic Energy Transfer within Asymmetric Pairs of

3

Following the excitation, several processes may occur in the excited state. As a rule, excited molecules in liquid solutions rapidly relax to the zero-vibrational level of S1 by non-radiative vibrational relaxation (VR), or by internal conversion (IC) from higher singlet states followed by VR. Then the molecules can return to the ground state S0 by fluorescence, or via internal conversion to S0 followed by VR. They can also reach S0 by intersystem crossing (ISC) to the triplet state T1, followed by phosphorescence or radiationless deactivation from T1 to S0. The radiative transition T1→S0 is forbidden and usually cannot compete with other relaxation processes. In a fluorescence experiment the intensity of emitted radiation F(t) is measured:

]S[]S[

)( *1r

r

*1 k

dt

dtF =

−∝ (2.1)

In Eq. 2.1, ]S[ *1 is the population of the excited S1 state at time t, the subscript "r"

refers to radiative pathway of deactivation from S1 to S0, and kr stands for the radiative rate constant. The proportionality factor in Eq. 2.1 depends on instrumental conditions and in the following will be omitted. The time-evolution of F(t) in response to a δ-pulse (i.e., infinitely short) excitation is referred to as the fluorescence relaxation, or the fluorescence decay. In the case where only unimolecular processes compete with the emission (as shown in Fig. 2.1), F(t) follows a first-order law

)()()(

nrr tFkkdt

tdF +−= (2.2)

Here, knr stands for the overall rate constant for all non-radiative processes, which result in relaxation of the S1 state. Integration of Eq. 2.2 leads to a single exponential expression )/exp()0()( τtFtF −= (2.3)

where 1nrr )( −+= kkτ is the fluorescence lifetime. The quantity r0 /1 k=τ is usually

called the radiative lifetime, and 0nrrr /)/( ττ=+= kkkQ is the fluorescence quantum

yield [2]. The values of 0τ are usually within 1–10 ns, while the fluorescence lifetimes

often become shorter due to non-radiative processes.

Page 10: Electronic Energy Transfer within Asymmetric Pairs of

4

2.2 Experimental techniques In a steady-state measurement, stationary excitation light is used, such as that produced by a xenon (Xe) arc lamp. The fluorescence intensity is then typically measured as a function of the emission or excitation wavelength (λem or λex, respectively). A schematic of a steady-state fluorescence spectrometer is shown in Fig. 2.2. The proper excitation wavelength is selected by means of the excitation monochromator. The emitted light is most often monitored at a right angle to the excitation beam. The wavelength of emitted light is selected by the emission monochromator, and its intensity is registered by a photomultiplier tube (PMT).

Figure 2.2. The principal set-up of a

steady-state fluorescence spectrometer.

The dynamics of excited states relaxation is measured in time-resolved fluorescence experiments. A common technique used for time-resolved measurements is called time-correlated single photon counting (TCSPC). Other methods, as well as their

advantages and disadvantages, are described elsewhere [1, 2]. In a TCSPC set-up (Fig. 2.3A) a pulsed light source is used, such as a flash lamp, a pulsed laser, or a light-emitting diode (LED). The light source activates a time-to-amplitude converter (TAC). Once the sample is excited an emitted photon can reach the PMT. This in turn sends a stop pulse to the TAC. The amplitude of the output voltage of the TAC is proportional to the time elapsed between the start and stop pulses. That time is then stored in a multi-channel analyser (MCA), in which each channel represents a short time interval. Every photon is registered in the MCA by adding one to the corresponding channel. An IBH 5000U system, which is used in our laboratory, is also shown in Fig. 2.3B.

Page 11: Electronic Energy Transfer within Asymmetric Pairs of

5

Figure 2.3. (A) Schematic diagram of a time-resolved fluorescence spectrometer and (B) an

IBH 5000U system.

In a real experiment, the convolution of the instrument response function L(t) with the δ-pulse response of the sample is observed:

∫ ′′−′≡⊗=t

tdttFtLtFtLtI0

)()()()()( (2.4)

Direct deconvolution of experimental TCSPC data is unstable. Instead, one usually fits the data by using a physically reasonable kinetic model with relatively few fitting parameters )..,( 1model pxxtF (an alternative approach will be discussed in Section 4.6).

The fitting procedure [8] involves minimisation of χ 2

−=

i i

piip

xxtItIxx

21model

12 )..,()(

)..(σ

χ (2.5)

In Eq. 2.5, )..,( 1model pi xxtI is the convolution )..,()( 1model pxxtFtL ⊗ at t = ti, and σi is

the standard deviation of the i-th data point (channel). The statistics of a TCSPC

experiment obeys a Poisson distribution, and )(2ii tI≅σ . The quality of fit is judged

by the reduced χ 2 parameter )1/(22r +−= pNχχ , where N is the number of data

points, and p is the number of fitting parameters. For a few hundred of data points,

Page 12: Electronic Energy Transfer within Asymmetric Pairs of

6

2.12r ≤χ is usually considered acceptable. In addition to the value of 2

rχ , graphical

tests are useful, such as the weighted residuals and the autocorrelation plots [8]. 2.3 Fluorescence Anisotropy Upon excitation with linearly polarised light the emission from a fluorescent sample is partly polarised. The emission polarisation is caused by the preferential excitation of a sub-population of fluorophores whose absorption transition dipole moments Ma are oriented in a direction close to that of the electric field vector of light E. The

probability of excitation is proportional to 2a )( ME • , and the phenomenon is known

as photoselection. In practice FZZ and FZX intensities are measured (Fig. 2.4). The steady-state fluorescence anisotropy can be calculated as

ZXZZ

ZXZZs 2gFF

gFFr

+−

= (2.6)

where g is the instrumental correction factor g = FYZ/FYX.

Figure 2.4. Typical configuration for measuring fluorescence anisotropy.

Page 13: Electronic Energy Transfer within Asymmetric Pairs of

7

In much the same way as with the steady-state anisotropy, the time-resolved anisotropy can be calculated by measuring polarised components. The time-resolved anisotropy is defined as

)(2)(

)()()(

ZXZZ

ZXZZ

tGFtF

tGFtFtr

+−

= (2.7)

and the correction factor G = FYZ/FYX can be also calculated as

+−

=)(

)(

21

1

ZX

ZZ

s

s

tF

tF

r

rG (2.8)

The time-resolved fluorescence anisotropy reflects changes in the direction of transition dipole moment [ ])()0()( 20 tPrtr µµ •= (2.9)

where 2/)13()( 22 −= xxP is the second-rank Legendre polynomial, µ stands for the

unit vector of the emission transition dipole moment, and r0 is the limiting anisotropy. The brackets ⟨…⟩ indicate an ensemble average over all excited molecules. A common molecular origin of time-dependent fluorescence depolarisation is rotational motions of fluorophores. The energy transfer and energy migration processes can also contribute to the depolarisation, as discussed in further sections. 2.4 Donor-acceptor energy transfer and distance measurements Donor-acceptor energy transfer, often referred to as fluorescence resonance energy transfer (FRET), is the radiationless transmission of excitation energy from a donor (D) to an acceptor (A) [1]. It can be described according to the following kinetic scheme

** ADAD +→+ (2.10) The quantum mechanical explanation to FRET was first given by Förster in 1948 [9]. Assuming weak dipole-dipole coupling between an electronically excited donor and an acceptor in its electronic ground state, the rate of energy transfer ω is given by

Page 14: Electronic Energy Transfer within Asymmetric Pairs of

8

6

0

D

2

2

3

=R

R

τ

κω (2.11)

Here Dτ , ⟨κ 2

⟩, R and R0 stand for the fluorescence lifetime of the donor, the averaged

square of the angular part of the dipole-dipole interaction, the distance between the interacting molecules and the Förster radius, respectively. The latter is defined according to

ANn

JQR

45D6

0128

)3/2)(10(ln9000

π= (2.12)

where QD, n and NA stand for the quantum yield of the donor, the refractive index of the medium and Avogadro's constant, respectively. The overlap integral J between the normalised donor fluorescence FD(λ) and the acceptor absorption εΑ(λ) can be calculated as

∫= λλλελ dFJ 4AD )()( (2.13)

Values of R0 in the range of 10–80 Å were reported [2]. At R = R0, on the average, half the excitation energy of the donor is transferred to the acceptor, while the other half is dissipated by intramolecular relaxation processes. The angular part of the dipole-dipole coupling in Eq. 2.11 is given by

222121

2 }/))((3{ RRµRµµµ ••−•=κ (2.14)

In Eq. 2.14, R is the vector connecting the interacting molecules, 1µ and 2µ denote

the unit vectors of their transition dipole moments, and R = |R|. In general, ⟨κ 2⟩ cannot

be calculated exactly, unless an additional knowledge of the system (e.g., symmetry) is available. However, the range of possible ⟨κ 2

⟩-values can be determined from time-resolved anisotropy measurements [10]. In practice ⟨κ 2

⟩ = 2/3 is often used, which is the isotropic dynamic average value. Due to the ⟨κ 2

⟩1/6 dependence of R, this

approximation is often reasonable. Sometimes it is convenient to rewrite Eqs. 2.11–2.12 as follows:

6

00D

2

2

3

=R

R

τ

κω (2.15)

Page 15: Electronic Energy Transfer within Asymmetric Pairs of

9

ANn

JR

4560

128

)3/2)(10(ln9000

π= (2.16)

0Dτ in Eq. 2.15 stands for the radiative lifetime of the donor. In this definition of R0 the

quantum yield of the donor does not appear. From Eqs. 2.15–2.16 it is also obvious that energy transfer rate is independent of intermolecular relaxation processes, such as fluorescence quenching (e.g., by oxygen). A measure of energy transfer is the transfer efficiency (E). It can be obtained from steady-state fluorescence measurements: DDA /1 QQE −= (2.17)

where the relative quantum yields of the donor in the presence and absence of energy transfer are denoted by QDA and QD, respectively. The transfer efficiency depends on the distance R between the donor and the acceptor, according to [1, 3]

[ ] 160 )/(1

−+= RRE (2.18)

The time-resolved fluorescence experiments usually provide better accuracy of distance measurements, especially for E ≈ 0 or E ≈ 1. The fluorescence relaxation of the donor in the presence of FRET [1, 3] is given by )/exp()( DDA tttF ωτ −−= (2.19)

and the fluorescence decay of the acceptor [12, 13] is )/exp()/exp()( ADAD τωτ ttttF −+−−−∝ (2.20)

The analysis of either of these decays yields the rate of energy transfer (ω). In some cases, time-resolved fluorescence measurements can also reveal a distribution of donor-acceptor distances [14]. In 1967 the Förster's theory was experimentally confirmed by Stryer and Haugland [11]. This work is sometimes considered as the starting point of FRET applications in biosciences [3]. Nowadays, FRET is widely used for distance measurements in biomacromolecules and supramolecular assemblies. Surveys of many related and

Page 16: Electronic Energy Transfer within Asymmetric Pairs of

10

relevant papers are found in books describing fluorescence spectroscopy [1, 2], as well as in textbooks specialised on energy transfer [3]. FRET does not provide complete 3D-structures and the atomic resolution comparable to that of X-ray and NMR methods. However, FRET successfully complements these techniques when sensitivity, time-resolution and non-invasiveness are of importance. 2.5 Donor-donor energy migration As a special case of FRET, energy transfer between identical chromophores is possible:

*212

*1 DDDD +↔+ (2.21)

Ideally, the energy transfer between identical fluorophores is reversible, which means that the electronic energy is exchanged back and forth within a pair. For this reason many scientists use the concept energy migration (EM) [15] in order to distinguish it from donor-acceptor energy transfer, which is irreversible. In the following, the process of energy migration within a pair of identical fluorescent groups is referred to as donor-donor energy migration (DDEM) [4-6]. The process (2.21) is sometimes also called homotransfer [2, 3]. If the donors exhibit a single exponential photophysics, the fluorescence relaxation of the D1D2 pair is invariant to the rate of DDEM. However, the energy migration causes fluorescence depolarisation, which was one of the earliest observations of energy transfer in solution [16]. The depolarisation, as a result of DDEM, indicates the donors are separated by an average distance comparable to the Förster radius for the D1D2 pair. This effect has been qualitatively used in many applications [17, 18]. On the other hand, very few studies deal with the quantitative analysis of fluorescence depolarisation in terms of energy migration rates and donor-donor distances. In fact, the idea of using DDEM for distance measurements appears to be very attractive for the following reason. The conventional studies of proteins using FRET involve specific labelling of a protein molecule with one donor and one acceptor group. This is a crucial step and in practice, it is often extremely difficult to perform. The problem with donor-acceptor labelling is circumvented by introducing one kind of specific fluorophore. This is referred to as the DDEM method. Although the problem with specific labelling is solved, the use of DDEM introduces other practical problems and theoretical demands. At first, the process of EM is usually detected by performing fluorescence depolarisation experiments, which are more complicated and time-consuming than fluorescence lifetime measurements.

Page 17: Electronic Energy Transfer within Asymmetric Pairs of

11

What is more important, energy migration and the reorienting motions of the fluorescent molecules both contribute to the depolarisation measurements, and the separation of these processes is very complex. As a result, an analytical expression for the time-resolved DDEM anisotropy can be obtained only for the case of immobile fluorophores [19-21], and in a few other simple cases [22]. Different researchers have considered the problem of separation of energy migration from the reorienting motions and come up with different models and theories. A complete theoretical description, which relates experimental data to the simultaneous motions and energy migration within a donor-donor pair, is referred to as the extended Förster's theory (EFT) [23]. A difficulty with the EFT in the analyses of TCSPC data is the lack of an explicit analytical expression that can be used in the deconvolution procedure. Recent studies show, however, that the EFT is applicable in studies of model systems [24, 25], and even more recent results demonstrate how to apply EFT to the analyses of DDEM-data obtained with proteins [26]. In addition to the EFT, several models of fluorescence depolarisation in the presence of energy migration and reorienting motions were suggested [4, 27-30]. One of these is the DDEM model, which was developed for analysing the fluorescence anisotropy obtained from experiments with singly and doubly fluorophore-labelled proteins [4-6]. Here one considers DDEM within a pair of chemically and photophysically identical fluorescent groups (denoted D1 and D2), which are covalently linked to a macromolecule, such as a protein. From depolarisation measurements the time-resolved fluorescence anisotropy is obtained for a coupled system (D1D2), as well as for the single donors: D1 and D2. These time-resolved anisotropies are denoted r(t), r1(t) and r2(t), respectively. The anisotropy contributions to r(t) from donors excited indirectly through the energy migration D1 → D2 and D2 → D1 are denoted by r12(t) and r21(t). The DDEM model reads

[ ] [ ][ ])(1)()(2

1)()()(

2

1)( 211221 tptrtrtptrtrtr −+++= (2.22)

)2,1(])()1[()( 220 =+−= jStSrtr jjjj γ (2.23)

( )

++×−== δδ γγρ SSSttSSSrtrtr 212121002112 )()(2

1)()()( (2.24)

where p(t) is the excitation probability of the initially excited donor group in absence of fluorescence relaxation

[ ])2exp(12

1)( ttp ω−+= (2.25)

Page 18: Electronic Energy Transfer within Asymmetric Pairs of

12

In Eqs. 2.23–2.24, Sj is a second rank order parameter for each of the two donor groups. The order parameter describes the orienting distribution of the transition dipoles:

][2 jjj PS zµ •= (2.26)

The angle between the symmetry axes zj of the orientational distributions of the two donors is denoted by δ and )(cos2 δδ PS = . The maximum contribution to the

anisotropy from the secondary excited fluorophores is given by 0ρ , and γj(t) describes

the reorientation dynamics of the donor molecules. In the analyses of DDEM-data ω,

0ρ and δ are fitting parameters. The DDEM model was extensively tested and used in

practice [4-6, 31-33]. In the following the anisotropy expressions (2.23–2.24) will be used for modelling the PDDEM anisotropy.

Page 19: Electronic Energy Transfer within Asymmetric Pairs of

13

3 PARTIAL DONOR-DONOR ENERGY MIGRATION (PDDEM) One of the major limitations of the DDEM method is the need to use fluorescent groups that are exceptionally insensitive to their microenvironment. Most fluorescent molecules do not fulfil this criterion, which limits the number of useful probes and consequently, the range of measurable distances. It is, however, still possible to make use of the interaction between two chemically identical but photophysically non-identical D-molecules. Here one deals with DD-pairs for which the energy migration is only partially reversible. We refer to this case as to partial donor-donor energy migration (PDDEM). Below, a theoretical model of PDDEM is presented. 3.1 The PDDEM model In the following theoretical treatment of PDDEM, we consider two D-molecules that are localised in two different positions (α and β) of a rigid macromolecule. To determine the fluorescence decays of Dα and Dβ in the absence of energy transfer, the corresponding singly labelled macromolecules that contain Dα and Dβ are studied separately. Thereby the fluorescence lifetimes τα and τβ are determined. The rates of energy transfer from Dα to Dβ and from Dβ to Dα are denoted by ωαβ and ωβα , respectively. These rates depend on the distance R between Dα and Dβ according to Eqs. 2.11–2.12. The following scheme of a coupled system summarises the kinetics: (3.1) Similar kinetic schemes are characteristic for various photophysical and photochemical processes [7, 12], and the corresponding master equations are well known. The observed fluorescence decay of the coupled system (3.1) is conveniently expressed as (Paper I):

)exp()exp()( 2em2ex1em1ex tttF tt λλαβ pCppCp += (3.2)

where

Page 20: Electronic Energy Transfer within Asymmetric Pairs of

14

]4)([2

1 211112,1 βααββααββαβααββα ωωωωττωωττλ +−+−±−−−−= −−−− (3.3)

±±

−=

βαββα

αβαβα

ωτλωωωτλ

λλ mmm

mmm

/1

/11

1,2

1,2

212,1C (3.4)

= β

α

ex

exex

p

pp and

= β

α

em

emem

p

pp (3.5)

In Eq. 3.5, αexp and β

exp denote the relative initial excitation probabilities of Dα and

Dβ, respectively. In a similar way, the components of pem are proportional to the

probabilities of registration of a photon from Dα and Dβ. The values of αemp and β

emp

are determined by the radiative rates (Eq. 2.1) as well as by the experimental set-up. Up to here, the theoretical treatment of PDDEM is very similar to a previously developed model of partly reversible FRET [34]. For real systems, the fluorescence decay for an ensemble of macromolecules labelled with fluorescent probes can be nonexponential (see, e.g., ref. [35]), even if it is single exponential for the free probe in the corresponding solvent. In this case the complex relaxation is usually modelled as a sum of discrete exponential decays:

)/exp()(

)/exp()(

j

j

j

i

i

i

tftF

tftF

βββ

ααα

τ

τ

−=

−=

∑ (3.6)

The fluorescence relaxation of the coupled system is then given by

),,()(,

jij

ji

i tFfftF βααββααβ ττ∑= (3.7)

Here ),,( jitF βααβ ττ is calculated from the above equations by using the lifetimes iατ

and jβτ instead of τα and τβ, respectively. The validity of Eq. 3.7 and the required

approximations are discussed in Papers I and IV. Eqs. 3.2–3.7 are directly applicable to the analysis of experimental data. 3.2 PDDEM anisotropy As is demonstrated above, the fluorescence relaxation of a DαDβ coupled system depends on the rates of energy migration ωαβ and ωβα . Hence, it should be possible to

Page 21: Electronic Energy Transfer within Asymmetric Pairs of

15

calculate these rates (and thus, the Dα–Dβ distance R) from the fluorescence intensity decays. However, the difference between the photophysics of Dα and Dβ could be too small for accurate distance measurements. Then, a correction to the DDEM anisotropy expression (Eq. 2.22) is needed. The fluorescence anisotropy is additive, which means that the fluorescence anisotropy of DαDβ system )(trαβ is a linear combination of the anisotropies of directly and

indirectly excited donors. The anisotropy expressions (2.23–2.24) are the same for PDDEM and DDEM, and the only difference is the weights (or contributions) of these anisotropies in the observed anisotropy decay )(trαβ . The weights are the probabilities

of the emission from directly and indirectly excited Dα and Dβ. For instance, the contribution of indirectly excited Dβ is readily calculated from Eqs. 3.2–3.7 by using

=

0ex

ex

αpp and

= β

emem

0

pp (3.8)

More details are given in Paper I, and several special cases of the PDDEM anisotropy will be discussed in the Section 4. It is worth emphasising that the fluorescence relaxation of the coupled system (3.1) contains distance information, in contrast to the "classic" DDEM case. Therefore, the rates of energy migration and the distances can be measured in fluorescence intensity decay experiments. The conditions under which the distances can be calculated with reasonable accuracy are explored in the next Section. The fluorescence anisotropy in the presence of PDDEM can also be modelled and used for distance measurements.

Page 22: Electronic Energy Transfer within Asymmetric Pairs of

16

4 RESULTS AND DISCUSSION In this Section the PDDEM model is tested by applying it to several systems of known structure. It is also demonstrated how the PDDEM method can be used for distance measurements in proteins. 4.1 Simulations of PDDEM One of the advantages of TCSPC technique is its well-defined Poissonian statistics, which makes possible a very realistic modelling of TCSPC data (cf. Eq. 2.4) for any given fluorescence relaxation F(t). Synthetically generated data can thus be re-analysed, in order to check if the initially assumed parameters are recovered within a desired accuracy. Instabilities in the analysis of the synthetic data would strongly indicate that certain difficulties should be expected in the corresponding real experiments. Here we employed this approach to explore the limitations of the PDDEM model. The range of distances that can be accurately measured was of particular interest. To generate synthetic TCSPC data, we used a method proposed by Chowdhury et al [36]. The generated decays mimicked true experimental data with respect to the statistical noise, the number of photons collected, as well as the photophysical properties of commonly used fluorescent groups (Paper I). The synthetically generated fluorescence decays of Dα, Dβ, and the coupled DαDβ pair were further reanalysed to determine the Dα–Dβ distance (R). To start, let us consider a donor-donor pair for which only the fluorescence lifetimes differ between the α and β sites. For such systems the migration rates ωωω βααβ == ,

as can be seen from Eqs. 2.15–2.16. In practice, this is the case when one of the D-groups is exposed to dynamic quenching. The relative error in distance was calculated from the reanalyses of synthetic decay data. The mean values of error are summarised in Fig. 4.1A. It appears that accurate distance measurements are difficult at high rates of energy migration (i.e., short distances). This problem also appears for the case of similar lifetimes (Fig. 4.1A). Both of these difficulties have a common explanation: for fast energy transfer, defined by

11 −− −>> αβ ττω (4.1)

it can be shown that Eq. 3.2 is approximately given by

Page 23: Electronic Energy Transfer within Asymmetric Pairs of

17

)]2/2/[exp()( 11 ttF −− +−≈ βααβ ττ (4.2)

Thus, the fluorescence relaxation of such a DαDβ pair contains very limited information about the transfer rates.

Figure 4.1. The ratios between the

distances (R) obtained from

reanalysis of the synthetic data and

the distances assumed in simulations

(Rtrue), for (A) βα0R = αβ

0R and

,ememβα pp = and (B) βα

0R =

0.9 αβ0R and β

emp = 1.2 αemp . The

fluorescence lifetimes of Dβ are

0.6τα (�), 0.8τα (∇), and τα (�).

The error bars indicate random

scatter of the results.

For different spectral shifts and shapes of the absorption and/or emission spectra of Dα

and Dβ, the Förster radii αβ0R and βα

0R may differ. In the following set of simulations,

it was assumed that βemp = 1.2 α

emp and that βα0R = 0.9 αβ

0R . The range of distances,

which is then recovered, is illustrated in Fig. 4.1B. Notice that no significant difference between τα and τβ is required for an accurate determination of R. Furthermore for equal fluorescence lifetimes, the correct rate of energy migration can be extracted. Contrary to the previous case of equal transfer rates (Fig. 4.1A, Eq. 4.2), the decay component containing energy migration rates contributes significantly to

Fαβ(t), even in the fast transfer limit given by Eq. 4.1. This is explained by βαemem pp ≠ ,

for which Eq. 4.2 is not valid. Therefore, the lower limit of recovered distances is likely dictated by the time-resolution of the experimental set-up. Similar simulations were performed to model fluorescence depolarisation experiments in the presence of PDDEM (Fig. 4 and 5 in Paper I). In general, the range of measurable distances is not very different from that of DDEM (cf. Fig. 5, Paper I).

Rtrue

0.6 0.8 1.0 1.2 1.4 1.6 1.8

R/R

true

0.8

0.9

1.0

1.1

1.2

R/R

true

0.8

0.9

1.0

1.1

1.2 A

B

Page 24: Electronic Energy Transfer within Asymmetric Pairs of

18

Moreover, the fluorescence depolarisation experiments are complementary to studies of the fluorescence intensity decays, while the latter are more suitable for measuring longer (R ≥ R0) distances. 4.2 PDDEM within asymmetrically quenched bichromophoric molecules The first experimental system to test the PDDEM model was a bichromophoric molecule RhC32Rh (1,32-dihydroxy-dotriacontane-bis-(Rhodamine 101) ester) [31, 37]. This molecule spans across the lipid bilayer of vesicles formed by 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC, see Fig. 4.2). An asymmetric system is created by adding a water-soluble fluorescence quencher (sodium iodide) to the outer bulk solution of the vesicles.

Figure 4.2. A schematic of RhC32Rh solubilised in lipid vesicles. The polar Rh101 groups are

anchored in the lipid-water interface.

For determining the fluorescence lifetimes and rotational correlation times of the rhodamine group in absence of energy migration, we used the amphiphile-like molecule RhC18 (Rhodamine 101 octadecyl ester) solubilised in DOPC vesicles. RhC18 constitutes about half a RhC32Rh molecule. In the absence of NaI the fluorescence decay of RhC18 in DOPC vesicles is nearly single exponential, with a lifetime of about 4.7 ns. In the presence of 100 mM NaI, the fluorescence relaxation of RhC18 is well described by a sum of two exponential functions, corresponding to an average lifetime of 3.5 ns. Thus, the lifetimes of rhodamine groups at the inner and

Page 25: Electronic Energy Transfer within Asymmetric Pairs of

19

outer leaflets of the vesicles are significantly different, which means that we are dealing with the PDDEM process. The attempts to calculate the Rh-Rh distance from the fluorescence decay data were not successful, which was also expected from the PDDEM simulations (Section 4.1). Indeed, the expected distance between rhodamine groups (35 – 40 Å) is much shorter than the Förster radius for the Rh-Rh pair, R0 = 59 ± 1 Å [31]. Furthermore, one can

assume that ωωω βααβ == , βαexex pp = , βα

emem pp = , and rα(t) = rβ(t) for the RhC32Rh

molecule. It means that energy transfer is very fast in comparison with the difference between relaxation rates (Eq. 4.1), which leads to the approximation given by Eq. 4.2. Thus, the distance R between the rhodamine groups can be measured only in a fluorescence depolarisation experiment. By using the PDDEM anisotropy model, we obtained R = 35.1 ± 1.4 Å, which corresponds to a migration rate of (3.7 ± 0.9) × 109 s–1. The R-value agrees very well with the value of 35.5 ± 1 Å, previously determined by using the DDEM method [31, 38]. 4.3 PDDEM between ionic forms of fluorescein With the previous test system (Section 4.2) we were unable to verify the model of fluorescence relaxation in the presence of PDDEM (Eqs. 3.2–3.5). This motivated us to look for another suitable system to explore the PDDEM model. In Paper II, we applied the model to measure the distances between fluorescein (Fl) groups that exhibit pH-sensitive fluorescence [39-41]. We studied a lipid-spanning bifluorescein molecule FlC32Fl (dotriacontanedioic acid bis fluoresceinyl-5-methylamide) incorporated into lipid vesicles of different bilayer thickness. The schematic of this model system is analogous to that shown in Fig. 4.2; the difference is that now pH in the outer (pHout) or inner (pHin) bulk of the vesicles can be varied. To study the properties of fluorescein group in lipid bilayers in the absence of energy transfer, we employed "half" a FlC32Fl molecule, FlC16 (5-hexadecanoylaminomethyl fluorescein). The equilibria between different species of fluorescein in aqueous media are rather complex [40, 41]. Here it is sufficient to consider the two fluorescent species (Fig. 4.3A), i.e., the dianion (Fl2–) and the monoanion (Fl1–). These ionic forms exist at relatively high pH, and the following equilibria are relevant for the pH range studied: (4.3) All distance information is contained in the fluorescence relaxation of Fl2–-Fl1– pairs, while the fluorescence decays of donor-donor pairs Fl2–-Fl2– and Fl1–-Fl1– are invariant to energy migration. Therefore, it is desirable to keep the fractions of these pairs

Fl Fl1- + H+ Fl2- + 2H+K1 K2

Page 26: Electronic Energy Transfer within Asymmetric Pairs of

20

relatively small, as compared to Fl2–-Fl1– pair. This corresponds to an optimum pH value of about pK2 (cf. eq. 4.3) provided that pHin = pHout. The creation of a transmembrane pH gradient provides another way to increase the fraction of the Fl2–-Fl1– pairs. Moreover, the PDDEM model must be corrected for the contributions from Fl2–-Fl2– and Fl1–-Fl1– pairs to the observed fluorescence relaxation.

Figure 4.3. (A) Mono- (Fl1–) and dianionic (Fl2–) forms of fluorescein, and (B) the excitation

and emission spectra of Fl2– (solid lines) and Fl1– (dashed lines). The spectra were obtained for

FlC16 molecule solibilised in DOPC vesicles. For this system the pKa of the monoanion-dianion

transition (pK2) is 8.0.

To calculate the excitation and emission probabilities (cf. Eq. 3.5) and the Förster radii, one needs to know the absorption and fluorescence spectra of Fl2– and Fl1–. The pK1 and pK2 values as well as the individual spectra of Fl2– and Fl1– can be calculated from spectral data [42] by using the ionisation model given by Eq. 4.3. The individual spectra (see Fig. 4.3B) were also obtained from 2-dimensional excitation-emission spectra using the DATAN algorithm [43]. The Förster radii calculated according to Eq. 2.12 together with other photophysical properties of Fl1– and Fl2– are collected in Table 4.1. In contrast to the previous test system (RhC32Rh in DOPC vesicles), the case of PDDEM between Fl2– and Fl1– is asymmetric also with respect to energy migration rates ( βααβ ωω ≠ ), as well as to the excitation and emission probabilities

( βαexex pp ≠ and βα

emem pp ≠ ).

At first we measured the distance between the fluorescein groups within FlC32Fl molecule in DOPC vesicles at pHin = pHout ≈ pK2. The fluorescence decay of FlC32Fl in this system is shown in Fig. 4.4. There is clearly an apparent shift of the FlC32Fl decay curve peak, while the curves are almost parallel then. This strongly indicates the influence of energy transfer, preferably in the direction of Fl1–→Fl2–. By fitting the

O OO

COO

CH2 NH

O OHO

COO

CH2 NH

C

O

C15H31RC

O

C

O

(CH2)30 RR

FlC16FlC32Fl

Fl2- Fl1-

or R =

A

Wavelength, nm400 450 500 550 600

Flu

ores

cenc

e, a

.u.

0.0

0.2

0.4

0.6

0.8

1.0

B

Page 27: Electronic Energy Transfer within Asymmetric Pairs of

21

PDDEM model (Eqs. 3.2–3.5) to experimental decays, we determined a distance of R = 39.2 ± 1.6 Å between the fluorescein groups (see Table 4.2). Table 4.1. The Förster radii (R0) for different combinations of mono- and dianionic forms of fluorescein. The fluorescence lifetimes (τ ), the fluorescence quantum yields (Q), the excitation (pex), and the emission (pem) probabilities are also shown.

ionic

form

R0 (Å),

acceptor = Fl2–

R0 (Å),

acceptor = Fl1– τ (ns) Q

pex (a.u.),

λex = 450nm

pem (a.u.)*,

λem = 550nm

Fl2– 48.9 ± 1.3 33.5 ± 0.9 4.1±0.1 0.88±0.02 0.38 2.3

Fl1– 43.3 ± 1.6 30.9 ± 1.1 4.7±0.3 0.41±0.03 1 1

* estimated from the emission spectra; in the analysis pem was one of the fitting parameters.

Figure 4.4. The fluorescence decay

curves obtained for FlC16 (�) and FlC32Fl

(�) in DOPC vesicles. The response

function is also shown (solid line). The

distance between the Fl fluorophores is

39.9 Å as calculated from the shown data.

The small difference between the fluorescence decays of FlC16 and FlC32Fl (Fig. 4.4) raises the question: How reliable is a distance determination? To explore the stability of the analysis, the following tests were performed. Distances R' were fixed at the values varying between 1.1R and 0.9R, while other floating parameters in the analysis were kept free. The average χ2-value corresponding to both R' = 1.1R and R' = 0.9R was 1.23, which is significantly higher than the χ2 found for the correct distance R. Some of the fitting results were also rejected because negative pem-values were extracted. In addition to the χ2-tests, we generated and reanalysed synthetic TCSPC data for distances ranging between 20 and 60 Å (Paper II). In another series of experiments FlC32Fl was studied when solubilised in DMPC vesicles. The thickness of DMPC bilayer is about 3.5 Å less than that of DOPC. Indeed this difference was detectable (Table 4.2). However in this case the stability of the analysis is worse, which is also expected for short distances (Section 4.1).

Time, ns

0 2 4 6 8 10

Num

ber

of c

ount

s, x

103

4

6

8

10

15

20

25

Page 28: Electronic Energy Transfer within Asymmetric Pairs of

22

Table 4.2. The distances (R) between fluorescein groups of FlC32Fl molecule solubilised in lipid bilayers of DOPC and DMPC.

System Method R (Å) Average χ2

DOPC, pH ≈ pK2 PDDEM 39.2 ± 1.6 1.11

DOPC, pH >> pK2 DDEM 38.6 1.05

DMPC, pH ≈ pK2 PDDEM 34.0 ± 1.0 1.06

DMPC, pH >> pK2 DDEM 34.8 1.14

DOPC, pHin < pHout PDDEM 40.2 ± 0.9 1.19

DOPC, pH ≈ pK2 PDDEM* 40.2 ± 1.3

* the steady-state model derived from the PDDEM

It is possible to increase the fraction of Fl2–-Fl1– pairs by creating a transmembrane pH-gradient (pHin < pHout) . In this case one expects to obtain the most reliable data. We therefore created a pH gradient ∆pH of ~1 unit by adding sodium hydroxide to a vesicle suspension. It turns out that, because of several problems related to H+ leakage through DOPC bilayer, the data become distorted. Despite this complication, a reasonable value of the distance was found (Table 4.2). The distances found are in a good agreement with independent determinations. The bilayer thickness determined by means of X-ray diffraction is 38 Å for DOPC [44] and 34.5 Å for DMPC [45] bilayers. Finally, we performed DDEM experiments on the same systems but at high pH values that ensure that Fl2– form was dominating. The distances obtained from these measurements (Table 4.2) are also in a good agreement with those measured by PDDEM. Interestingly, the rates of energy migration can be calculated even from the steady-state fluorescence spectra, provided that the excitation and/or emission spectra of two donors are significantly different. It is obvious that the excitation and emission spectra of a DαDβ pair can be described by a linear combination of the individual spectra of Dα and Dβ components. In the presence of energy migration, however, the fractions of these spectral components are no longer directly proportional to the concentrations, but depend on the rates ωαβ and ωβα. In Paper II, the equations for analysing the steady-state data are obtained by integration of the time-resolved relaxation described by the PDEEM model (Eqs. 3.2–3.5).

The contributions ( βexx ) of the Fl2– excitation spectrum to the spectra of FlC16 and

FlC32Fl are shown in Fig. 4.5. The data points corresponding to FlC32Fl spectra show larger fraction of the monoanion spectrum, suggesting that some fraction of Fl2– is

Page 29: Electronic Energy Transfer within Asymmetric Pairs of

23

excited by energy transfer from Fl1–. The distance of 40.2 ± 1.3 Å was determined by fitting integrated PDDEM model to the experimental data (Table 4.2). Actually this value agrees very well with the values reported above. The solid line in Fig. 4.5 represents the best fit corresponding to the distance found. The curves calculated for the wrong distances R' = 0.9R and R' = 1.1R are clearly below and above the data points, respectively (Fig. 4.5, dashed lines). The corresponding mean-square deviations are 3 to 8 times higher for the wrong distances than for R.

Figure 4.5. Contribution of Fl2– to the

excitation spectra is shown as a function

of the emission wavelength. The data for

FlC16 (�) and FlC32Fl (�) are shown.

The best fit to the FlC32Fl data (solid line)

corresponds to a distance R = 41.1 Å. The

theoretical curves calculated when using

R' = 0.9R and R' = 1.1R (dashed lines) are

below and above the best fit, respectively.

4.4 PDDEM in proteins In Paper III, the PDDEM model was applied for distance measurements between BODIPY groups, which were covalently linked to cystein residues in plasminogen activator inhibitor of type 2 (PAI-2) [46, 47]. Two sulfhydryl specific derivatives of BODIPY were used namely; N-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-yl) methyl iodoacetamide (SBDY) and N-(4,4-difluoro-1,3,5,7-tetramethyl-4-bora-3a,4a-diaza-s-indacene-2-yl) iodoacetamide (NBDY). The latter fluorophore exhibits environmental-sensitive fluorescence [48]. The PAI-2 mutants created enabled specific single or double labelling with NBDY or SBDY. The following mutants were used for single labelling: 79cys PAI-2, 171cys PAI-2 and 347cys PAI-2 (Fig. 4.6). These mutants were also combined to construct the double mutants 79cys/171cys PAI-2 and 79cys/347cys PAI-2 mutants that were used for the double labelling experiments [49, 50]. In general, the spectral and photophysical properties of BODIPY fluorophores only weakly depend on the microenvironment, and thus derivatives of BODIPY are more suitable for DDEM studies. Nevertheless, suitable quenchers (e.g., I–) can increase the rates of fluorescence relaxation. The accessibility of BODIPY groups to a quencher depends on their localisation in the protein. Upon adding quenchers one can

Emission wavelength, nm520 540 560 580 600

x β ex

0.4

0.5

0.6

0.7

0.8

Page 30: Electronic Energy Transfer within Asymmetric Pairs of

24

preferably, and to a different extent, change the fluorescence lifetimes of emitting groups localised at different positions (α or β) in the protein. Under such conditions, the chemically identical Dα and Dβ groups become photophysically different, and the fluorescence lifetime experiments on DαDβ pairs can be thus used for distance measurements.

Figure 4.6. The ribbon structure of plasminogen activator inhibitor type 2 (PAI-2). The filled

circles indicate positions of the Cα-atoms of the Cys residues that were labelled with either

NBDY or SBDY. The chemical structures of these fluorescent probes are also shown.

Dynamic quenching by iodide has no effect on the rates of energy migration (Eqs. 2.15–2.16). Also the excitation and emission probabilities (Eq. 3.5) are independent of the quenching by I–. Using these simplifying conditions, the fluorescence relaxation of a coupled DαDβ system (Eq. 3.2) is given by

)/()]exp()2()exp()2[()( 21211

1111

2 λλλττλλττλ βαβααβ −+++−−−= −−−− tttF (4.4)

For multiexponential decays of Dα and Dβ the fluorescence relaxation of the coupled system is given by Eq. 3.7. In practice the samples of double-labelled protein may contain non-negligible fractions of single labelled proteins. It may be then necessary

NB

NNH C

O

CH2I

F F

NB

N

CH2 NH C CH2I

OF F

NBDY

SBDY

Page 31: Electronic Energy Transfer within Asymmetric Pairs of

25

to account for contributions to the fluorescence from the single labelled mutants [51]. By denoting the labelling efficiencies at positions α and β by φα and φβ, respectively, one obtains )()()1()()1()( tFtFtFtF αββαβαβαβα φφφφφφ +−+−= (4.5)

The corresponding fluorescence anisotropy decay of a DαDβ pair is furthermore given by

[ ] [ ]

[ ] [ ])(1)()(4

)(1)()()()(2

1)(

11

tptrtr

tptrtptrtrtr

−−−

+−++=

−−

βααβ

αββααβ

ωττ

(4.6)

where

)()(

)]exp()[exp(21)(

21

21

tF

tttp

αβλλλλω

−−−= (4.7)

Eqs. 4.6–4.7 can be obtained from a more general expression (Paper I), when equal excitation and emission probabilities are assumed, as well as a single exponential decay for both Dα and Dβ, and ωωω βααβ == . The first and second terms in Eq. 4.6

constitute the DDEM anisotropy expression, while the last correction term vanishes in the case of fast energy transfer (Eq. 4.1). The quenching of NBDY and SBDY by I– in 171cys PAI-2 and 347cys PAI-2 is markedly higher than that of 79cys PAI-2. This fact ensures a significant difference between the lifetimes of the BODIPY groups in these positions, which is required for accurate distance measurements (Section 4.1). Overall the quenching of SBDY is more efficient than for NBDY. To determine the energy migration rates, time-resolved fluorescence relaxation for two singly labelled forms of PAI-2, as well as the corresponding doubly labelled protein were combined and analysed in a global manner [52]. Typical time-correlated single-photon data are displayed in Fig. 4.7. From the migration rate extracted, the distance between the centres of mass of each fluorophore within the interacting pair (R) was calculated using the previously determined Förster radii; i.e. 57 Å for SBDY [5] and ~ 42 Å for NBDY [48]. The distances obtained from PDDEM measurements are presented in Table 4.3. The errors in distances presented in Table 4.3 represent the

Page 32: Electronic Energy Transfer within Asymmetric Pairs of

26

standard deviations obtained for 2-3 independent measurements. In addition, several stability tests were performed (Paper III).

Figure 4.7. Time-correlated single photon

counting data showing the fluorescence

relaxation of SBDY in the singly {79cys

PAI-2 (∆) and 171cys PAI-2 (�)} as well

as doubly {79cys/171cys PAI-2 (�)}

labelled PAI-2. The quencher

concentration [I–] = 0.2 M. The

instrument response function is also

displayed (dashed line).

Table 4.3. Distances between the fluorescent groups in labelled PAI-2 mutants as determined by PDDEM and DDEM methods.

System Method R (Å)

SBDY in 79cys/171cys PAI-2 PDDEM 55.6 ± 1.0

SBDY in 79cys/171cys PAI-2 DDEM 51.0 ± 1.2*

NBDY in 79cys/171cys PAI-2 PDDEM 51.7 ± 1.8

NBDY in 79cys/171cys PAI-2 PDDEM depolarisation 48.5 ± 5.0

SBDY in 79cys/347cys PAI-2 PDDEM 47.5 ± 2.5

SBDY in 79cys/347cys PAI-2 DDEM 44.0 ± 1.3*

NBDY in 79cys/347cys PAI-2 PDDEM depolarisation 39.3 ± 1.0

* data from ref. [50]

The PDDEM-depolarisation experiments were performed on NBDY-labelled PAI-2. To hamper the influence of rotational tumbling of PAI-2, glycerol was added to the water buffer. The addition of glycerol does not influence the PAI-2 activity for concentrations used in the depolarisation experiments [50]. The fluorescence depolarisation data obtained for NBDY-labelled 79cys/171cys PAI-2 as well as the corresponding single mutants are shown in Fig. 4.8. The distances obtained from the PDDEM and the DDEM depolarisation experiments on NBDY and SBDY, respectively when attached to the same PAI-2 mutants are presented in Table 4.3. The distances are similar but systematically longer between SBDY:s. This is explained by the longer linker between the BODIPY core of SBDY and the Cα-atoms of the labelled cysteins.

Time, ns

5 10 15 20 25 30

Num

ber

of c

ount

s

101

102

103

104

105

Page 33: Electronic Energy Transfer within Asymmetric Pairs of

27

Figure 4.8. Fluorescence anisotropy

decays obtained for mono- and bis-

NBDY-labelled 79cys PAI-2 (∆),

171cys PAI-2 (�) and 79cys/171cys

PAI-2 (�).

The values of the intramolecular distances obtained from PDDEM measurements are systematically longer than those obtained from DDEM and PDDEM depolarisation measurements (Table 4.3). However, the deviations are within the experimental errors for all mutants, except for SBDY-labelled 79cys/171cys PAI-2. A difference between the systems is that glycerol (50 % w/w) was added when performing the DDEM experiments. This might have a small effect on PAI-2 structure and/or local motions of SBDY and NBDY groups attached. 4.5 The effect of PDDEM on nonexponential relaxation of identical donors The process of energy migration within pairs of chemically and photophysically identical donors, each exhibiting a nonexponential fluorescence decay, represents an interesting special case of PDDEM. It turns out that the fluorescence relaxation of the coupled system then becomes dependent on the rate of energy migration. This effect can be rationally explained by using the PDDEM model. Consider a DαDβ system for which τα ≠ τβ. For such a system it is more or less obvious (or can be easily proved using Eqs. 3.2–3.5) that the presence of coupling only increases the total relaxation rate, or decreases the average lifetime. In other words, for an interacting pair of donors, each exhibiting different rates of fluorescence relaxation, the total relaxation rate must be faster than that in absence of coupling. If the fluorescence relaxation of Dα and/or Dβ is not a single exponential function, for some

DαDβ pairs the lifetimes iατ and j

βτ (see Section 3.1) would be different. Within these

pairs energy migration would effectively increase the relaxation rate, as discussed above. Thus, even if the photophysics of Dα and Dβ is identical, but not single exponential, the average lifetime of DαDβ system can be significantly shorter in the

Time, ns

5 10 15 20 25 30 35

Ani

sotr

opy

0.0

0.1

0.2

0.3

0.4

Page 34: Electronic Energy Transfer within Asymmetric Pairs of

28

presence of coupling. Moreover, it means that the fluorescence relaxation of such a coupled system of identical donors must depend on energy migration rate, which of course cannot be observed when Dα and Dβ exhibit single exponential and the same photophysics [1]. The effect of shortening of average lifetimes in the presence of PDDEM is demonstrated in Fig. 4.9. In this particular example, both donors exhibit identical biexponential photophysics with τ2 = 0.2τ1. The results show that both kinds of lifetime averages decrease in the presence of energy migration. This is in agreement with an observation reported in Paper III. The average lifetime of e.g. SBDY in 79cys/171cys PAI-2 in the absence of quencher is 4.78 ns, which is markedly shorter that the arithmetic average of the lifetimes found for the corresponding single mutants (5.49 ns and 5.51 ns, respectively). The mechanism here proposed may also explain previous results reported in the literature [27]. Thus, the shortening of average fluorescence lifetimes in a coupled system doesn't necessary imply that a process other than energy migration is involved.

Figure 4.9. Average and amplitude-

averaged lifetimes of coupled donor-

donor systems (solid lines) plotted vs. the

fraction of the shorter lifetime a2. The

corresponding average lifetimes in

absence of energy transfer are shown as

dashed lines. The fluorescence relaxation

of both donors was assumed to be

biexponential, τ1 = 1, τ2 = 0.2, a1 = 1–a2,

and the energy migration rate ω was

taken to be 1 (=1/τ1).

4.6 The modelling of nonexponential fluorescence relaxation in FRET and PDDEM Fluorescent groups typically exhibit a complex photophysics when incorporated into biomolecular structures, e.g., proteins and lipid membranes. In practice the nonexponential fluorescence decay is most often fitted to a sum of exponential functions [1], while models that assume continuous distributions of lifetimes are also used [53, 54]. Unfortunately, both models can be statistically very well fitted to realistic experimental data, implying that one cannot definitely distinguish the nature

a2

0.0 0.2 0.4 0.6 0.8 1.0

<τ>

0.2

0.4

0.6

0.8

1.0 averagelifet imes

amplitude-averagedlifetimes

Page 35: Electronic Energy Transfer within Asymmetric Pairs of

29

of lifetime distributions [55]. In the analyses of energy transfer and PDDEM data, one frequently needs to model the nonexponential decays of the non-interacting donor and acceptor groups. In Paper IV we investigated whether the obtained energy transfer/migration rates and distances depend on the model used to describe the fluorescence relaxation. For a continuous distribution of lifetimes the fluorescence relaxation is given by (see, e.g., ref. [56])

∫∞

−=0

)/exp()()( τττ dtatF (4.8)

In practical calculations the continuous distributions of lifetimes are usually approximated by discrete distributions defined for a lifetime grid { iτ }. The lifetime

distributions in the presence of PDDEM were calculated in discrete form according to Eq. 3.7. To minimise the number of floating parameters in the analyses, the shape of the distribution )(τa can be chosen a priori. We selected a Gaussian distribution [54, 57]

−−=2

2

2

)(exp

2

1)(

σττ

σπτa (4.9)

In Eq. 4.9, τ is the average fluorescence lifetime, and σ is the standard deviation. Alternatively, the lifetime distribution can be reconstructed using the maximum entropy method (MEM) [53, 58, 59]. The MEM selects the distribution that maximises the function

2χν −= SQ (4.10)

where S is an entropy-like function [58, 59]

−−=

i i

iiii m

aamaS

)(

)(log)()()(

τττττ (4.11)

and ν is a constant [60]. At first we generated TCSPC-like data (Section 4.1) assuming that the "true" fluorescence relaxation is described by a continuous distribution of lifetimes. The synthetic data were then reanalysed using both the biexponential relaxation model and

Page 36: Electronic Energy Transfer within Asymmetric Pairs of

30

the continuous lifetime distribution model. Fig. 4.10A shows the initial lifetime distributions of the donors Dα and Dβ, as well as the calculated distribution for the DαDβ coupled system at R = R0. To mimic the case of BODIPY-labelled PAI-2 (Paper

III), we also assumed that ωωω βααβ == , βαexex pp = , and βα

emem pp = . The ratios

between the obtained distances and the values assumed in simulations (Rtrue) are presented in Fig. 4.10B. One can clearly see that differences between the distances obtained using two different models are minor over the range of distances where stable results are expected. The analogous simulations were performed to model donor-acceptor energy transfer experiments (see Paper IV). It was found that the obtained distances exhibit little or negligible dependence on the model used to describe the fluorescence relaxation.

Figure 4.10. (A) Distributions of

lifetimes for DαDβ system in the absence

(dashed lines) and in the presence (solid

lines) of energy migration. (B) The ratios

between the obtained distances R and the

true distances Rtrue. The distances were

calculated by assuming a continuous

lifetime distribution model (�) as well as

the multiexponential relaxation model

(�).

The experimental PDDEM data on BODIPY-labelled PAI-2 protein were also reanalysed by using a model that assumes a continuous distribution of lifetimes. The lifetime distributions of NBDY-labelled

79cys PAI-2 and 171cys PAI-2 single mutants reconstructed by MEM (Eqs. 4.10–4.11) are presented in Fig. 4.11. The calculated distribution of lifetimes for the double mutant 79cys/171cys PAI-2 is also shown (Fig. 4.11). The distances obtained are summarised in Table 4.4. The difference between the methods of analysis is within 1%, as can be seen from Table 4.4.

τ /τ0

0.0 0.2 0.4 0.6 0.8

a(τ )

0.0

0.1

0.2

0.3

DαDβ

Dα -DβA

Rtrue

0.6 0.8 1.0 1.2 1.4 1.6

R/R

true

0.8

0.9

1.0

1.1

1.2

B

Page 37: Electronic Energy Transfer within Asymmetric Pairs of

31

Figure 4.11. The distributions of

lifetimes for NBDY-labelled 79cys PAI-

2 and 171cys PAI-2 calculated using the

maximum entropy method (Eqs. 4.10–

4.11). The lifetime distribution of the

coupled system (NBDY-labelled

79cys/171cys PAI-2) is also shown.

Table 4.4. Distances (R) between the BODIPY groups in mutant forms of PAI-2 obtained by using various models of the nonexponential fluorescence decay.

PAI-2 mutant

R (Å),

multiexponential

model

R (Å),

Gaussian distribution

of lifetimes

R (Å),

maximum entropy

method

SBDY-79/171cys 55.6 ± 1.0 55.8 ± 0.9 55.6 ± 1.0

NBDY-79/171cys 51.7 ± 1.8 52.1 ± 2.0 52.1 ± 1.9

Finally, we reanalysed the experimental FRET data obtained for BODIPY-labelled ribosomal protein S6 from Thermus thermophilus (Paper IV). The distances between tryptophan residues and BODIPY groups can be obtained from the analysis of the fluorescence relaxation of BODIPY (cf. Eq. 2.20) [61]. In analogy to the case of PDDEM, the obtained donor-acceptor distances were almost the same irrespective of the model used for the complex fluorescence relaxation of the donor and the acceptor.

Lifetime, ns0.3 1 3 10

a(τ )

0.0

0.1

0.2

0.3

0.4

0.5

171cys

79cys

Page 38: Electronic Energy Transfer within Asymmetric Pairs of

32

5 CONCLUSIONS AND FUTURE PERSPECTIVES The PDDEM method extends the possibilities of measuring distances between chemically identical fluorophores, which exhibit different photophysical and/or spectral properties depending on their localisation in a macromolecule. Contrary to the DDEM method, the distance information is obtained from fluorescence lifetime measurements, provided that the properties of the two groups are significantly different. Furthermore, the PDDEM model can be used for quantitative analysis of time-resolved fluorescence depolarisation in terms of energy migration rates and donor-donor distances. In the present theoretical treatment, the reorientational motions of donors are not included in the model of fluorescence relaxation in the presence of PDDEM. This can be considered as a weakness of the PDDEM model, because these motions may occur on the same timescale as the EM and the fluorescence relaxation. A complete theoretical description of the PDDEM can be developed in a similar way as the extended Förster theory (EFT) [23] of the DDEM. An extended theory of PDDEM can be then employed to explore the limitations of the PDDEM model (cf. ref. [25]), or used as it stands for the analysis of experimental PDDEM data (cf. [24]). Moreover, information about the angular configuration of the donors (and consequently, the ⟨κ 2

⟩-factor) is available from the time-resolved anisotropy decays when combined with the EFT analysis [26]. In addition to the complete theoretical description of EM, several research groups are developing new methods for incorporating fluorescent probes into protein structures [62, 63]. In comparison with the standard methods of protein labelling [39], the site-specific incorporation of unnatural fluorescent amino acids into proteins seems to be a very promising idea. This approach could circumvent several difficulties, such as: low labelling efficiency, perturbation of the protein structure, and the uncertainty in distance measurements due to significant linker lengths between the polypeptide backbone and a fluorescent group. Taken together, the EFT combined with such methods may become a versatile tool that complements standard NMR and X-ray techniques, where protein sizes, concentrations and crystalline qualities are inadequate. In addition to measuring distances within pairs of fluorophores, fluorescence methods have already been shown to be very useful in studies of the aggregation of proteins. The aggregation is often connected to biological function of proteins or the related diseases [64, 65]. The PDDEM model could also be extended to quantitative studies of several interacting fluorophores, arranged in a regular way. The theory of EM in such system is currently under development [66].

Page 39: Electronic Energy Transfer within Asymmetric Pairs of

33

6 ACKNOWLEDGEMENTS First of all, I would like to thank my supervisor Lennart B.-Å. Johansson for giving me the opportunity to work on an interesting project, for his excellent scientific guidance, and for all our fruitful discussions. I hope I learned something from you. I'm proud to be a member of the "fluorescence group" at the Department of Biophysical Chemistry. I'm truly grateful to Fredrik and Peter for their help during my first year here. Mikael, Denys, Nils, Erik, and, of course, Ilya: it has been a pleasure to work together with you. I'm also grateful to everyone at the Biophysical Chemistry, for being positive and ready to help. I would like to thank some of you in particular. Anita: for taking care of all paper work. Anna-Karin: for keeping all chemical stuff in order. Gerhard: for your optimism and for most of the social activities at the department (including the "öl klockan fem" event). Per-Olof Westlund: for the "Liouville" course, which substantially improved my understanding of what is written in this thesis. Julian Georgievich Molotkovsky and the Lipid Chemistry group in Moscow – for the syntheses of bichromophoric molecules, which were essential for this work. But in particular, thank you for introducing me to fluorescence: after all, I'm not disappointed with it. I would also like to thank Prof. Tor Ny, Sergei and Shouye at the Department of Medical Biochemistry and Biophysics for their contributions to the PAI-2 and NBDY projects. The Russian-speaking community of Umeå – for a warm welcome to the north of Sweden! Thank you for all the great time that I had here (besides working on this thesis). Finally, I wish to thank my parents and friends, for your support and encouragement.

Page 40: Electronic Energy Transfer within Asymmetric Pairs of

34

7 REFERENCES 1. J.R. Lakowicz (1999). Principles of Fluorescence Spectroscopy, 2nd ed.,

Kluwer Academic/Plenum Publishers, New York. 2. B. Valeur (2002). Molecular Fluorescence. Principles and Applications,

Wiley-VCH, Weinheim, Germany. 3. B.W. Van der Meer, G. Coker III, and S.-Y.S. Chen (1994). Resonance

Energy Transfer: Theory and Data, VCH Publishers, New York. 4. L.B.-Å. Johansson, F. Bergström, P. Edman, I.V. Grechishnikova and J.G.

Molotkovsky (1996). J. Chem. Soc., Faraday Trans. 92, 1563-1567. 5. J. Karolin, M. Fa, M. Wilczynska, T. Ny and L.B.-Å. Johansson (1998).

Biophys. J. 74, 11-21. 6. F. Bergström, P. Hägglöf, J. Karolin, T. Ny and L.B.-Å. Johansson (1999).

Proc. Natl. Acad. Sci. 96, 12477-12481. 7. M. Klessinger and J. Michl (1995). Excited States and Photochemistry of

Organic Molecules, VCH Publishers, New York. 8. J.N. Demas (1983). Excited state lifetime measurements, Academic Press,

New York. 9. T. Förster (1948). Ann. Phys. 2, 55-75. 10. R.E. Dale and J. Eisinger (1974). Biopolymers 13, 1573-1605. 11. L. Stryer and R.P. Haugland (1967). Proc. Natl. Acad. Sci. USA 58, 719-726. 12. J.B. Birks (1970). Photophysics of Aromatic Molecules, John Wiley & Sons

Ltd, London, UK. 13. T. Kulinski, A.J.W.G. Visser, A. O'Kane and J. Lee (1987). Biochemistry 26,

540-549. 14. E. Haas, M. Wilchek, E. Katchalski-Katzir and I.Z. Steinberg (1975). Proc.

Nat. Acad. Sci. USA 72, 1807-1811. 15. J. Baumann and M.D. Fayer (1986). J. Chem. Phys. 85, 4087. 16. E. Gaviola and P. Pringsheim (1924). Z. Physik 24, 24. 17. D.M. Jameson, J.C. Croney and P.D.J. Moens (2003). Methods Enzymol. 360,

1-43. 18. S. Kalinin and L.B.-Å. Johansson (2004). J. Fluorescence 14, 681-691. 19. F. Tanaka and N. Mataga (1979). Photochem. Photobiol. 29, 1091-1097. 20. Y.R. Kim, P. Share, M. Pereira, M. Sarisky and R.M. Hochstrasser (1989). J.

Chem. Phys. 91, 7557-7562. 21. P.I.H. Bastiaens, A. Van Hoek, J.A.E. Benen, J.C. Brochon and A.J.W.G.

Visser (1992). Biophys. J. 63, 839-853. 22. P.-O. Westlund and H. Wennerström (1993). J. Chem. Phys. 99, 6583-6589.

Page 41: Electronic Energy Transfer within Asymmetric Pairs of

35

23. L.B.-Å. Johansson, P. Edman and P.-O. Westlund (1996). J. Chem. Phys. 105, 10896-10904.

24. P. Edman, P. Håkansson, P.-O. Westlund and L.B.-Å. Johansson (2000). Molec. Phys. 98, 1529-1537.

25. P. Edman, P.-O. Westlund and L.B.-Å. Johansson (2000). Phys. Chem. Chem. Phys. 2, 1789-1794.

26. P. Håkansson, M. Isaksson, P.-O. Westlund and L.B.-Å. Johansson (2004). J. Phys. Chem. B In Press.

27. L.W. Runnels and S.F. Scarlata (1995). Biophys. J. 69, 1569-1583. 28. M.L. Barcellona and E. Gratton (1996). Biochemistry 35, 321-333. 29. B.D. Hamman, A.V. Oleinikov, G.G. Jokhadze, R.R. Traut and D.M. Jameson

(1996). Biochemistry 35, 16680-16686. 30. M.P. Lillo, O. Canadas, R.E. Dale and A.U. Acuna (2002). Biochemistry 41,

12436-12449. 31. S.-T. Bogen, J. Karolin, J.G. Molotkovsky and L.B.-Å. Johansson (1998). J.

Chem. Soc., Faraday Trans. 94, 2435-2440. 32. M. Fa, F. Bergström, P. Hägglöf, M. Wilczynska, L.B.-Å. Johansson and T.

Ny (2000). Structure 8, 397-405. 33. I. Mikhalyov, S.-T. Bogen and L.B.-Å. Johansson (2001). Spectrochim. Acta

A 57, 1839-1845. 34. P. Woolley, K.G. Steinhauser and B. Epe (1987). Biophys. Chem. 26, 367-

374. 35. J. Karolin, L.B.-Å. Johansson, L. Strandberg and T. Ny (1994). J. Am. Chem.

Soc. 116, 7801-7806. 36. F.N. Chowdhury, Z.S. Kolber and M.D. Barkley (1991). Rev. Sci. Intrum. 62,

47. 37. J. Karolin, S.-T. Bogen, L.B.-Å. Johansson and J.G. Molotkovsky (1995). J.

Fluorescence 5, 279. 38. S.-T. Bogen, G. de Korte-Kool, G. Lindblom and L.B.-Å. Johansson (1999). J.

Phys. Chem. B 103, 8344-8352. 39. R. P. Haugland (2002). Handbook of Fluorescent Probes and Research

Products, 9th ed., Molecular Probes, Eugene, OR. 40. R. Sjöbäck, J. Nygren and M. Kubista (1995). Spectrochim. Acta A 51, L7-

L21. 41. N. Klonis and W.H. Sawyer (1996). J. Fluorescence 6, 147-157. 42. M. Kubista, R. Sjöbäck and J. Nygren (1995). Anal. Chim. Acta 302, 121-125. 43. I. Scarminio and M. Kubista (1993). Anal. Chem. 65, 409-416. 44. A.B. Bergenståhl and P. Stenius (1987). J. Phys. Chem. 81, 5944.

Page 42: Electronic Energy Transfer within Asymmetric Pairs of

36

45. L.J. Lis, M. McAlister, N. Fuller, R.P. Rand and V.A. Parsegian. (1982). Biophys. J. 37, 657-665.

46. G.A. Silverman, P.I. Bird, R.W. Carrell, F.C. Church, P.B. Coughlin, P.G. Gettins, J.A. Irving, D.A. Lomas, C.J. Luke, R.W. Moyer, P.A. Pemberton, E. Remold-O'Donnell, G.S. Salvesen, J. Travis and J.C. Whisstock (2001). J. Biol. Chem. 276, 33293-33296.

47. P.H. Jensen (1997). Inter. J. Onc. 11, 557-570. 48. M. Isaksson, S. Kalinin, S. Lobov, T. Ny and L.B.-Å. Johansson (2003). J.

Fluorescence 13, 379-384. 49. M. Wilczynska, S. Lobov and T. Ny (2003). FEBS Lett 537, 11-16. 50. S. Lobov, F. Bergström, M. Wilczynska, L.B.-Å. Johansson and T. Ny (2004).

Proc. Nat. Acad. Sci. USA Submitted. 51. J. Karolin, P. Hägglöf, T. Ny and L.B.-Å. Johansson (1997). J. Fluorescence

7, 331-339. 52. J.R. Knutson, J.M. Beechem, and L. Brand (1983). Chem. Phys. Letters 102,

501-507. 53. A.K. Livesey and J.C. Brochon (1987). Biophys. J. 52, 693-706. 54. J.R. Lakowicz, H. Cherek, I. Gryczynski, N. Joshi and M.L. Johnson (1987).

Biophys. Chem. 28, 35-50. 55. D.R. James and W.R. Ware (1985). Chem. Phys. Letters 120, 455-459. 56. D.R. James and W.R. Ware (1986). Chem. Phys. Letters 126, 7-11. 57. I. Gryczynski, M.L. Johnson and J.R. Lakowicz (1994). Biophys. Chem. 52, 1-

13. 58. A.K. Livesey and J. Skilling (1985). Acta Crystallogr. B A41, 113-122. 59. J.C. Brochon (1994). Methods Enzymol. 240, 262-311. 60. S.F. Gull and J. Skilling (1999). Quantified Maximum Entropy: MemSys5

Users' Manual, 1.2 ed., Maximum Entropy Data Consultants Ltd., Suffolk, U.K.

61. S. Kalinin, M. Olofsson, J. Zdunek, M. Oliveberg and L.B.-Å. Johansson (2004). In preparation.

62. Z.W. Zhang, B.A.C. Smith, L. Wang, A. Brock, C. Cho and P.G. Schultz (2003). Biochemistry 42, 6735-6746.

63. L. Wang, A. Brock, B. Herberich and P.G. Schultz (2001). Science 292, 498-500.

64. J.-C. Rochet and J. Lansbury, P. (2000). Curr. Opinion in Struct. Biol. 10, 60-68.

65. C.M. Dobson (2003). Nature 426, 884-890. 66. D. Marushchak and L.B.-Å. Johansson (2004). In preparation.


Recommended