• Nie Znaleziono Wyników

Photobleaching kinetics of fluorescein in quantitative fluorescence microscopy

N/A
N/A
Protected

Academic year: 2021

Share "Photobleaching kinetics of fluorescein in quantitative fluorescence microscopy"

Copied!
13
0
0

Pełen tekst

(1)

Biophysical

Journal Volume 68 June 1995 2588-2600

Photobleaching

Kinetics of Fluorescein in

Quantitative

Fluorescence

Microscopy

Loling Song,*f E. J. Hennink,*§ 1. Ted

Young,*

and Hans J.

Tanke*

*DepartmentofCytochemistryandCytometry, FacultyofMedicine, LeidenUniversity,2333ALLeiden; *Departmentof Pattern

Recognition, FacultyofApplied Physics, Delft UniversityofTechnology,2628 CJDelft; and§Lambert Instruments, 9313 TH

Leutingewolde,

The Netherlands

ABSTRACT

An

investigation

on

the

photobleaching

behavior of fluorescein

in

microscopy

wascarried out

through

a

systematic

analysis of photobleaching mechanisms. The individual photochemical reactions of fluorescein

were

incorporated into

a

theo-retical analysis and mathematical simulation

to

study the photochemical

processes

leading

to

photobleaching

of fluorescein in

microscopy.

The

photobleaching

behavior of free and bound fluorescein has also been

investigated by experimental

means.

Both the theoretical simulation

and

experimental

data show that

photobleaching

of fluorescein

in

microscopy is,

in

general,

not

a

single-exponential

process.

The simulation

suggests that the

non-single-exponential

behavior is caused

by

the

oxygen-independent, proximity-induced triplet-triplet

or

triplet-ground state dye reactions of bound fluorescein in microscopy. The

single-exponential

process

is

a

special

case

of

photobleaching

behavior when the reactions

between the

triplet dye

and

mo-lecular

oxygen are

dominant.

INTRODUCTION

Photobleaching is

a

dynamic

process

in which

fluorochrome

molecules

undergo photo-induced

chemical

destruction

upon

exposure

to

excitation

light

and thus

lose their

ability

to

fluo-resce.

The

photobleaching phenomenon

has been

the basis

of

many

fluorescence measurement

techniques developed

and

successfully

applied since the

1970s. One

of the

most

widely

known

is fluorescence

photobleaching

recovery

(FPR) based

on

early work by

Peters

et

al.

(1974). Much of

the

essentials

of

current

FPR

has

come

from the

laboratories

of W. W.

Webb

and E. L.

Elson

(Axelrod

et

al., 1976; Edidin

et

al.,

1976;

Jacobson

et

al., 1976;

Schlessinger

et

al., 1976,

1977).

FPR, with

its variants

known as

fluorescence

redistribution

after

photobleaching

(Koppel

et

al.,

1986)

and

fluorescence

microphotolysis

(Peters

et

al.,

1981),

has

been

widely used

in

the

past

20 years to

study

the

rates

of

diffusion

of

fluo-rescent

molecules

inside

and

on

the surface of cells. In the

area

of

fluorescence

resonance

energy

transfer, Jovin and

others have

applied

photobleaching

to

derive the

efficiency

of

energy

transfer

between

a

fluorescent

donor

and

an

ac-ceptor

molecule

in

microscopy

(Jovin

and

Arndt-Jovin,

1989; Jovin

et

al., 1990; Kubitscheck

et

al., 1991, 1993;

Marriott

et

al., 1991; Szabo

et

al., 1992; Young

et

al.,

1994).

This method offers the

distinct

advantage that complex

cor-rections associated with

the

determination of

energy

transfer

by combined

donor and sensitized acceptor

emissions

are

avoided. The

approach is

based

on

the demonstration

of

Hir-schfeld

(1976)

that,

for

unimolecular

or

pseudo-unimolecular

reactions,

the

time-integrated

emission of a

single-exponential photobleaching

process

is

independent

of

Receivedforpublication 31 October 1994 and infinalform 19 March 1995. Addressreprint requeststoLoling Song, Dept. of Cytochemistry and Cy-tometry, Leiden University, Wassenaarseweg 72, 2333 AL Leiden, The Netherlands. Tel.: 31-71-276198; Fax: 31-71-276180; E-mail:

loling@ruly46.1eidenuniv.nl.

X) 1995bytheBiophysical Society 0006-3495/95/06/2588/13 $2.00

fluorophore

quantum

yield, excitation intensity, and

photo-bleaching lifetime. Kubitscheck

et

al.

(1991)

also

tried

to

accommodate

more

complex decay kinetics.

The

development

of the abovementioned

techniques

made

it

possible

to

derive

quantitative

measurements

in

various

applications

in

cell science.

However,

the mechanisms of

photobleaching

in

biological objects

are not

yet

well

under-stood

(for

review see

Tsien

and

Waggoner,

1989; Wells

et

al.,

1989).

Almost

all

the

existing knowledge

on

photo-bleaching

comes

from

spectroscopy

studies of

fluorochrome

molecules

in

solution.

In

spectroscopy,

free fluorochrome

molecules

are

homogeneously dissolved

in

solution, and the

chemical environment is well controlled.

A

single

phenom-enon

is studied

at a

time

and

generally single-exponential

bleaching is observed.

In

microscopy, however,

there have not been

systematic

studies

on

the

mechanisms of

photobleaching.

It

is generally

assumed that the

photobleaching analysis

of

fluorochrome

in

solution

can

be

directly applied

to

the

microscope situation.

This

assumption

does

not

take

into

account

that,

in

micros-copy,

fluorochrome molecules

are

chemically bound

to

tar-gets

of interest

(such

as

DNA, RNA,

protein,

or

other cellular

components) about which

the chemical

microenvironment

is

very

complex, often differs

from one

specimen to another,

and

is

very

difficult

to

control.

Although

a

single-exponential

(or first-order)

process

is often used as a basis for the

pho-tobleaching techniques

in

microscopy, the experimental data

from many

studies deviate

from a pure

single-exponential

function

I

=

Be-kt (Benson

et

al., 1985; Koppel et al., 1989;

Rigaut and Vassy,

1991). Benson et al. (1985) carried out an

extensive

study

on

the

heterogeneous photobleaching

rates of

different

areas

within

a

cell

using

various fluorochromes.

They described

their

experimental

bleaching curve by

a

three-parameter

exponential:

I

=

A +

Be-kt

for each pixel

in an

image,

where the

offset

A

was

attributed to the

ever-present

background

fluorescence. Koppel et al.

(1989)

used

an

identical

model in their

analysis

of

photobleaching.

2588

(2)

Rigaut and Vassy

(1991)

examined their

photobleaching

curve

by transformation

in

log (i.e., single exponential

with-out an

offset) and demonstrated

that

photobleaching

in

con-focal

microscopy

was

clearly

not a

single-exponential

pro-cess,

but

the

origin of

the

non-single-exponential

behavior

was not

discussed.

Szabo

et

al.

(1992) reported

in their

study

using photobleaching fluorescence

resonance

energy transfer

that

the

photobleaching

kinetics of FITC

(fluorescein

iso-thiocyanate) attached

to

immunoglobulin followed

a

double-exponential

process. In

their

analysis

for the

possible

ex-planation of the second

exponential

component,

they

speculated

that

the difference

in

accessibility

of

quencher

molecules to

the

fluorochrome

molecules

and

the

photo-chemical destruction

of the

triplet

FITC

molecules could

be

among

the reasons for

the

double-exponential behavior.

The experimental deviation from the

single-exponential

behavior,

as

repeatedly

observed

by different research

groups,

is

not

likely

to

be

accidental. This leads

to a

fun-damental

question of whether certain

photochemical

pro-cesses

could

give rise

to

non-single-exponential

photo-bleaching behavior under

microscope conditions.

A

better

understanding of

photobleaching

in

microscopy

will

allow

further utilization of

this

phenomenon, which is of

particular

interest

in

quantitative

analysis of fluorescently stained

tis-sues,

cells

and

cell

organelles

using confocal laser

scanning

microscopy

or

digital

imaging microscopy

(see

Carrington

et

al.

(1989) and Tsien and Waggoner

(1989)

for

review).

The

study reported

here is

aimed

at a

systematic

analysis

of

photobleaching kinetics

in

microscopy.

It

focuses

on

the

photobleaching

process

of fluorescein alone.

First,

the

pho-tochemical

and

photophysical background of

photobleaching

is

discussed with

a

detailed

analysis focusing on

the specific

mechanisms

applicable

to

fluorescein. Then,

new

experi-ments on

the

photobleaching characteristics

of free

fluores-cein

in

solution and of fluorescein covalently

attached to

nucleic

acid

probes

in

microscopy

are

presented.

Through

a

mathematical

simulation, photochemical knowledge of

fluo-rescein

from

earlier

studies

(Lindqvist, 1960; Kasche and

Lindqvist, 1964; Usui

et

al.,

1965) is used

to

study the

ki-netics

of

fluorescein from a theoretical point of view. The

results from

the simulation

are

compared with

the

experi-mental data.

Finally,

the effects of

different

photobleaching

mechanisms

on

the

behavior

of

the photobleaching process

of

fluorescein

are

discussed.

It

is

demonstrated theoretically

and

experimentally

that the

photobleaching

of

fluorescein in

microscopy is,

in

general, not a

single-exponential

process.

PHOTOCHEMICAL

AND

PHOTOPHYSICAL

BACKGROUND

Decay and photobleaching

Decay and

photobleaching are two distinct light-induced

processes,

taking place on very different time scales. Upon

ex-posure

to

excitation

light with irradiance H (W

cm-2)

and

wave-length

Aex (m),

fluorochrome molecules absorb the photon

en-ergy

(hc/Aex,

with

Planck's constant h, the speed of light c) and

make

an

electronic

transition from the low energy ground state

S to the

higher energy excited singlet state

S*.

The rate of photon

absorption ka

(s-')

by a fluorophore molecule with an absorption

cross-section

(ra

(cm2

molecule-')

is

directly proportional

to

the

photon

flux and thus the irradiance according to

ka

=

(a

[H

Ae.l(hc)]

*

(1)

For

fluorescein, Oa

is 3.06

X

10-16

cm2/molecule

at

488

nm

and

pH

>

7

(Tsien and Waggoner, 1989).

With

pulsed

excitation,

the

number

of

fluorophore

molecules that absorb the

photon

energy

and reach the excited

singlet state will attain a certain maximum,

and

then become

depopulated

via fluorescence

emission,

radia-tionless

internal

conversion,

and radiationless

intersystem

cross-ing

to

the

excited

triplet

state

T*,

at rates

kv

kc,

and

kjsc,

respec-tively

(Fig. 1).

Decay

refers

to

the

composite

effect of these three

processes

after

the

fluorophore

is

exposed

to a narrow

pulse

of

excitation

light.

The

decay

lifetime, S*,

of

a

fluorophore

is the

inverse of the

sum

of

these

three

rate constants

and

rate constants

of

any

other

processes

(e.

g.,

resonance

energy

transfer),

which

depopulate

S*.

Namely,

TS* =

Ik

=

kf

+

ki.

+

+

kother

and it is

commonly

1-10

ns.

The

decay lifetime

is intrinsic

to

a

particular

fluorophore,

and

thus

independent

of

excitation

in-tensity for levels below saturation. Molecules that

cross to

the

excited

triplet

state are

likely

to

be

trapped in that

state

for

a

duration of microseconds

to

milliseconds, since the T**>S

tran-sition is

spin-forbidden

and

therefore the

rate

k1 of the T*--->S

transition is

very

low.

The

decay

process

does

not

necessarily

involve

fluorophore

decomposition, although

the latter could

take

place.

In

contrast,

photobleaching is

the process in which

the

total number

of

ground

state

molecules is depleted via

per-S

k.

'O'

FIGURE 1 Simplified Jablonski energy diagram for a generic fluoro-chrome. S,S*, andT*represent groundsinglet, excited singlet, and excited

tripletelectronic energy levels, respectively. Fluorochrome molecules ab-sorbphotonenergyat arateka and go from the ground singlet state S up

tothe excitedsinglet state S*. Then they may return to the ground state by radiative(fluorescence)or non-radiative(internalconversion)pathway at a combinedratekd. They may also undergo non-radiative intersystem

cross-ing,at a rate

kjSC,

toT*, where they may return to the ground state at a rate

k,.

Photobleachingmaytakeplace from both S* and T* at rates kbS and

ka,

respectively. Those photobleached molecules can no longer participate in theexcitation-emission cycle.

(3)

Biophysical

Journal

manent

photochemical

destruction when molecules

are

either

in

the

singlet

or

the

triplet

excited

state.

Photobleaching

is the

cumulative

effect of

fluorophore

loss from each

excitation-emission cycle

over

time,

and the

rate

of

photobleaching

is

a

function

of

the excitation

intensity. Those bleached

mol-ecules can no

longer

participate

in the

excitation-emission

cycle.

Under

normal mercury

arc

lamps,

photobleaching

is

observed on the

time

scale of seconds

to

minutes. With

a

laser

as an

excitation

source,

photobleaching

can occur

in

as

short

as a

few microseconds.

Fig.

2

depicts the

conceptual

differ-ence

between

decay

and

photobleaching

processes.

For an

"ideal" case in the absence of

photobleaching,

the

population kinetics

of each energy

state

are

described

by:

d

-j[Ns(t)]

=

-kaNs(t)

+

kdNs.(t)

+

klcNT*(t)

d

-t

[Ns.

(t)]

=

kaNs

(t)

-

(kd

+

kisc

)Ns

(t)

(2)

d

t

[NT*(t)]

=

kiscNs5(t)

-

k,NT*(t)

where

N(t) is

the molecular

population

in

the

energy

state

designated

by the

subscript

as a

function

of

time,

and kd is

the

sum

of

kf and

kic.

Shortly after

exposure to a constant

excitation

light

intensity,

the

population

in

the

singlet

excited

state

establishes

a

steady state. For longer

times,

the

triplet

state

also

achieves

a

steady

state

((d/dt)

[Ns(t)]

=

(d/dt)

[Ns.

(t)]

(d/dt)

[NT*(t)]

=

0).

The

system

in

Eq.

2

can

then

be solved

for

Ns, Ns.,

and

NT*,

with the normalized

constraint

Ns(t)

+

Ns*(t)

+

NT*(t)

=

1:

i-

kaki

ki(ka

+ kd +

kjsc)

+

ka

kisc

NT.

=

kakisc

NT*

=

(ka + kd +

kisc)

+

kakisc

Ns

=

1-Ns*-NT*

=

1-

ka(ki+kisc

ki

(ka

+

kd

+

kic

+

kakisc'

All the

dye

molecules

are

assumed

to

reside

in the

ground

state

before the

onset

of the excitation and

are

normalized to

unity.

The

steady

state

is

maintained,

and the

population

in

each

state

is

part of the total initial

population in the

ground

state.

The relative number of molecules in each state is

a

function of the excitation rate. At low excitation rate

(ka

<<

kd

and

at

constant

k,

and

k;,

NS.

and

NT*

are

small

fractions,

and

NS

--

1,

i.e.,

most

of

the

molecules,

reside in S

(NS

>>

Ns.,

NT*).

At

high

excitation rate

(ka

>

klc),

NS

is less than both

NS.

and

NT*,

and

most

of

the

molecules are

continuously

pumped

up

to

S*,

from

which

they

cross to

T*

at a

rate

kiSC.

For

any excitation rate, if

kl

<<

kisc,

over

time

NT>>

Ns.,

and

a

large

number of molecules accumulate in the lowest

triplet

excited

state.

In

practice, however,

photobleaching

does take

place,

and

fluorophore

molecules

undergo permanent

photochemical

destruction while

they

are

in the excited

singlet or

triplet

state.

The steady

state no

longer holds [(d/dt)

[Ns(t)]

*

0; (d/dt)

[Ns.

(t)]

#

0;

(d/dt)

[NT*

(t)]

#

0.]. Instead,

within

a

time

period

governed

by

the rates of all the

reactions

in-volved, the populations of the ground state, excited

singlet,

and excited

triplet

states

will

change

in unison

according

to

a

characteristic function determined by the

reactions

in-volved in the

photochemical

destruction. Hirschfeld

(1976)

derived an

analytical

expression

where

photobleaching

is

as-sumed to take

place from the

excited

singlet

state

and

the

photochemical

reactions are of a

unimolecular

or

pseudo-unimolecular nature. He demonstrated that

photobleaching

followed a

single-exponential function.

Wells et al. (1989)

derived an analytical

expression for

photobleaching

from the

excited triplet state. In

both of

these analyses,

the

models

were

simplified

to a

single

(combined)

unimolecular or

pseudo-unimolecular

photochemical

reaction. In

micros-copy,

the

photochemical reactions

are

much more complex,

sensitive to

various

environmental factors,

and can

involve

many

bimolecular

processes. In the

following section, a

de-tailed

analysis will be carried

out

for

fluorescein

photo-bleaching

in

microscopy.

FIGURE 2 Schematicdiagram depicting the con-ceptual difference between the decay and

photo-bleaching processes. The diagrams above are the types ofexcitation,pulsed orcontinuous, and those belowarethecorresponding emission. (a) Decay re-ferstothe depopulation of the excited singlet state afterexperiencingabriefexcitation pulse. This

pro-cessdoes notnecessarily involveapermanent loss of fluorochrome molecules, although the latter often

oc-curs.Thedecay process can be resolved on the time scaleofnanoseconds.(b) Photobleaching is a process

inwhich the total amount of ground statemolecules

are depleted via permanent photochemical destruc-tion.Photobleachingisthecumulative effect of flu-orophore loss from each excitation-emission cycle duetoacontinuous exposure to the excitation light. Photobleaching in fluorescence microscopy can be observed on the time scale ofmicroseconds to minutes. on on

*...

... ...

off

e

a

pulsed

excitation

t

a

decay

photobleaching

---7

--ns

-i-o

off

continuous

excitation

b

photobleaching

inms.

to

min.

emission

2590 Volume 68 June 1995 t

emission

(4)

Photochemical

properties

and

photobleaching

mechanisms

of

fluorescein

Although

in

principle

bimolecular processes between

a

sin-glet

excited

state

dye

molecule and

a

molecule of another

species

(such

as

oxygen)

can

lead

to

photobleaching

from the

dye

singlet,

there has

not

been much evidence of this

phe-nomenon

for

fluorescein

so

far.

In

the

early

1960s,

Lindqvist

conducted

an

extensive

flash

photolysis study

on

the

pho-toreactions of fluorescein

in deaerated

water

(Lindqvist,

1960).

This

was

followed

by

studies

on

the reactions of

triplet

fluorescein

with

oxygen

(Kasche

and

Lindqvist,

1964).

Lindqvist

pointed

out

that because of its very short

decay

lifetime

(4.5

ns) in

aqueous

solution,

the

singlet

excited

state

was not

ordinarily

responsible

for the

photochemical

activity

of

fluorescein.

Imamura

and

Koizumi

(1955)

showed,

through

a

theoretical

estimation based

on

oxygen

diffusion in

aqueous

solution,

that the lifetime of

the

inter-mediate

to

be attacked

by

oxygen

must

be

at

least

10-6

s.

More

recently, Gollnick

et

al.

(1992)

found that oxygen

ex-erted

no

measurable effect

on

the

short-lived excited

singlet

state

of xanthene

dyes

even

in

oxygen-saturated

solutions.

There

was,

however, much evidence in these studies for the

existence

of

quenching

of

the

long-lived lowest triplet

ex-cited

state,

which

was

populated

exclusively by

a

transition

from

the

singlet excited

state.

Lindqvist

(1960) demonstrated

that the

triplet excited

state

fluorescein molecules

became

depopulated

via two

major pathways:

one was

the reaction

between

a

triplet

and another

triplet

or a

ground

state

dye

molecule; and

the other

was

the

reaction between

a

triplet

dye molecule

and

an

oxygen

molecule. These

two

mecha-nisms were

later studied

by

Usui

et

al.

(1965) and

called D-D

(dye-to-dye)

and

D-0

(dye-to-oxygen)

mechanisms,

respec-tively.

The

major reactions involved in these

two

mecha-nisms are

listed

in

Table

1,

along with

the

corresponding rate

constants

derived by

Lindqvist for

the pH range used

in the

present

study.

Reactions

4

(T*

+

T*

--

R

+

X)

and 5

(T*

+

S

->

R

+

X)

represent

the

occurrence of an electron

transfer

process,

which leads

to

the

formation

of

semi-reduced

(R)

and

semi-oxidized

(X) radical

forms of the dye

molecule.

These

radicals

in turn

undergo

further reaction(s)

reverting

either to the ground state dye

(Kruger

and

Mem-ming,

1974)

or

to

stable non-fluorescent

photoproduct(s).

The exact fate of these radicals was

difficult to determine in

Lindqvist's

(1960) study.

In

the present study, all X and R

molecules

are

considered

candidates for the bleached dye

molecules.

It

is

noteworthy that the electron transfer

pro-cesses

(reactions

4

and

5) compete with the two quenching

processes

(reaction

2,

T*

+

T*

--

T

+

S,

and reaction

3,

T*

+

S

--

S

+

S) respectively

with about

equal

efficiency.

Quenching

of T*

by X and R

(respectively,

T* + X

->

S

+

X

and T

+

R

->

S

+

R) leads to the return of T* back

to

S

at an

equal rate without causing a net

population change

in

X

and

R.

Quenching of T* by

02

leads to

physical

quench-ing (T*

+

02

->

S

+

02)

at a rate of

about one order of

magnitude

faster than the chemical quenching (T* +

02

->

X

+

HO2).

To

study

the

photobleaching

behavior of

fluorescein,

all

of

the

photochemical

reactions from the

studies of

Lindqvist

and

Kasche

(Lindqvist, 1960;

Kasche and

Lindqvist,

1964)

are

incorporated

into

a

model

described

by

the

following

six

coupled

differential

equations:

d

[Ns(t)]

-

[kdNs*(t)

+

klNT*(t)

+

k2N¾2(t)

+

k3NT.

(t)Ns

(t)

+

k6NT*

(t)NX(t)

+

k7NT*

(t)NR

(t)

+

k8NT. (t)No(t)]

-

[kaNS(t)

+

k5NT*

(t)NS(t)]

d

[Ns.

(t)] =

kaNs

(t) -

[kdNs*(t)

+

kiscNs*(t)]

dt

[NT*(t)]

=

kiscNs(t)

-

[klNT.(t)

+

k2NT¾(t)

+

k3NT.(t)Ns(t)

+

2k4N¾2(t)

+ k5NT.

(t)Ns(t)

(3)

+

k6NT. (t)Nx(t)

+

k7NT* (t)NR(t)

+

k8NT.

(t)NoP(t)

+

k9NT. (t)NO2(t)]

d2

dt[ NX(t)]

=

k4NT¾(t)

+

kaNT*

(t)NS(t)

+

k9NT.

(t)NO2(t)

dt

d

[NR(t)]

=

k4N2*t)

+

kNT.

(t)NS(t)

dt

dt

[No2(t)]

=-k9NT*

(t)NoP(t).

In

this system,

kl-kg

are

those derived

by

Lindqvist

and

Kasche

(Lindqvist,

1960;

Kasche

and

Lindqvist,

1964) for

the pH range in

the

current

study

(Table 1).

The nonlinearity

introduced

by the bimolecular processes makes it extremely

difficult to find an

analytical solution for the system in

Eq.

3.

Instead, an efficient iterative numerical method has to

be

used to

study

the

photobleaching kinetics of

fluorescein.

MATERIALS

AND

METHODS

Experimental

approach

Free

fluorescein in solution

Afluorescein solution of 0.01 ,uM was made by dissolving fluorescein sodium (Merck's certifiedgrade, without further purification) in phosphate-buffered saline(PBS; pH 7.6). The solution was then placed in three

cu-vettes: 1) control sample,which was not exposed to the bleaching light source, 2) air-saturated sample, and 3) deoxygenated sample, from which oxygen was purged by flushing argon gas (oxygen content < 0.5 ppm, HoekloosBV,Schiedam, The Netherlands) for 15 min. A Leitz DM epi-fluorescence microscope (Leica,Wetzlar, Germany) with a 100-W mercury

arclampand a450-490-nmexcitation filter block was used as a bleaching light source. The objective lens was removed so that a column of light was

impingedonthe cuvette. The air-saturated and argon-flushed samples were exposedtothecontinuous bleaching light source for 90 min. The

fluores-cenceintensitywasmeasured before the first exposure and again at15-min

intervalsonaspectrofluorometer (SPF-500, SLM Instruments, Inc., Urbana, IL). The cuvettes were well shaken before each measurement. The spec-trofluorometer was equipped with a xenon lamp, and the emission and

(5)

ex-TABLE 1 Photochemical reactions of fluorescein

Reaction No. Description Rate constantst

S + hv-- S*

Absorption

ka

= 3.8 X 108s-' S* S + hv' Fluorescenceemission

kd

= 2.134 X 108 S-1 S* S Internalconversion S* T*

Intersystem

crossing

k,sc

= 6.6 X 106s1 T* S 1 Radiationless deactivation

k=

50s-' T* + T*-- T* + S 2

Triplet quenching

k2

= 5 X 108 M-1

s-1

T*+S S+S 3

k3

=5 X

107

M-' s-T* + T*-* R + X 4 Electrontransfer

k4

= 6 X 108 M-1 s-T* + S R + X 5 Electrontransfer

k5

= 5 X

107

M-lS1 T*+X S +X 6 T*

quenching by

X

k6

+

k7

=1 X

109

M`

s-'

T* + R S +R 7 T*

quenching by

R

k7

T* +

02

S +

02

8

Physical

quenching by 02

k8

= 1.56 X

109

M`

s-'

T* +

02-

X +

HO2

(or

O°)

9 Chemical

quenching by 02

kg

= 1.4 X 108 M-'

s-S,groundstatedye; S*,singletexcitedstatedye; T*,tripletexcitedstate

dye;

R, semi-reducedform of the

dye; X,

semi-oxidized form of the

dye;

02,

oxygen.

tThe

rate constants

K1

to

Kg

werequotedfromLindqvist(1960)and Kascheand

Lindqvist

(1964),

and

kAz

Ad

and

ki.,

were

quoted

fromTsien and

Waggoner

(1989).

citation monochromator slitswereset to490±4nmand to 512+ 4 nm,

respectively. All the instrument settingswerekeptconstantthroughoutthe

experiment.

Bound fluorescein

in

microscopy

Fluorescein surface-labeled microspheres (Quantum Series 25, Flow

Cytometry Standards Corp.,ResearchTriangle Park, NC), suspendedata

properconcentration inPBS,werecentrifugedat1200Xgfor 30 min onto

standardmicroscope slides. Non-adherent microspheresweregently rinsed off with PBS. The slideswereair-driedinthe dark and embedded in PBS inthe absence ofanti-fadingagents.

Ficol-isolated humanlymphocytesonglass slideswerein situhybridized and fluorescein moleculesweredirectly(without antigen-antibody complex

orspacermolecules) attachedtoprobesspecificfor the centromericregion of chromosome 1. After in situhybridization the preparationwas

counter-stained with diamidinophenylindole and embeddedunder acoverslip in

PBS.

The imagingsystemconsisted ofaLeitzAristoplan fluorescence

mi-croscopeequipped witha100-Wmercuryarclamp andacharge-coupled device(CCD)camera(series CH250, Photometrics Inc., Tucson, AZ) with

aKodak(Rochester, NY) KAF-1400 chip of 1348 linesX 1035pixels by 12-bitplanes. The shutter in the CCDcameraandamechanical shutter in

the excitationlight pathwerecomputercontrolled for thedesiredon-chip integration and duration of illumination, respectively. A filter block withan

excitation bandwidth of 450-490nm, adichroic mirror of 510nm,anda

long pass emission filter of 520 nm were used for fluorescein-stained specimens.

A Macintosh IIfxcomputerservedas ahostcomputer, whichdirectly controlled the shutters andimage acquisition. For each object of interest,a

series ofimageswereacquiredovertimeunder continuoussteady illumi-nation. Forinterphase chromosomes,asequenceofimageswere continu-ously acquiredover aperiod of 120s,andeachimagewasintegrated for 1.0s.For fluorescentmicrospheres, imageswerealso continuously acquired with anintegration time of 0.7 s/image. For both preparations, a2 X 2

binningontheCCDcamera wasusedtoincreasethe signal-to-noise ratio. ThemicroscopewasadjustedtoKohlerillumination and checked for flat-field withuranyl glass. For backgroundsubtraction, thepartof thegrayvalue

intensity image, I(x,y,t), outside theareaof the objectwasused. The

back-ground maskwasfoundby first applyingagradient filtertoI(x,y,t=0) and

thenthresholding the resultant imagetolocatetheareawith the magnitude

of thegradient closetozero(i.e., the background area). Thisarea wasthen

usedas animage mask for the subsequent images in thesameseries, and

a meanvalue for eachgrayvalueintensity image under the background mask multiplied by theareaof the whole imagewas subtracted from the total integrated intensity. The final scalar value of each imagewasthe total

in-tegrated intensity after thebackgroundcorrection.Ableachingcurvewas

formed byplottingintensityvalues of all theimagesin the seriesagainst

time.

Theoretical approach

Study of kinetics by mathematical simulation

Basedonthephotochemical reactions and theirrateconstantsderived from

Lindqvist'sstudies(Lindqvist, 1960;Kasche andLindqvist,1964),asystem

ofordinarydifferentialequations(ODEs)canbe formulated fromthe

re-action scheme(see Table 1 and Eq.3).Eachdifferential equation describes thechange ofapopulation of molecules overtime. The kineticscanbe monitored by solving thissystemofODEsusing numerical methods and

integratingit from the nanosecondtothe minute scale. The numerical meth-ods madeuseof thehighly efficient FORTRAN software package

Liver-moreSolver forOrdinaryDifferentialEquations(Hindmarsh, 1983)andwas

customized for the currentstudy. Thealgorithm and theinitialstepsize

(Presset al., 1992)took into considerationthe drasticallydifferent time scalesoriginatingfrom the nanosecondsinglet excitedstatelifetime and micro-tomillisecondtripletstate lifetime(a so-called stiff problem). A strategyofadynamically adjustedstepsizewasusedtoensure an accurate

solution at theearly(fast reaction) stage andafast convergence of the solutionatthe later(slow reaction)stageof thephotobleachingprocess.All

computationswereperformed in double-precision.

Simulation of

'"ree

fluorescein

in

solution"

Simulation of thephotobleachingprocessoffree fluoresceinin solution wasaccomplished by setting the appropriate initial conditionsand

re-cordingthepopulationchange of eachenergy state overtime.The initial

conditions for simulation of air-saturated fluorescein inaqueous

solu-tionwereN,(t=0)=0.01

AM

(as usedinthe ExperimentalApproach),

and

No2

(t = 0) = 250

AM

(Usuietal., 1965).Forsimulation of the

argon-flushedtreatment,0.01% of

N02

ofthe air-saturatedsolution was

usedas anestimate basedonthe result of Vaughan andWeber(1970)

thatthe inertgasflushing method didnotcompletelyremove all

02.

Then

No2

= 0.025,uMand

Ns

= 0.01 ,uM. In the cuvette experiment,

in which the microscope objectivewasremoved and collimated light wasimpingeduponthecuvette,thepowerdensitywasmuch lower than

thatunder theobjective (see below). Then

ka

wasestimatedto be -0.038

s51.

All the otherrate constants(Table 1) usedareintrinsicto fluorescein

forthe pH used in thepresent study, and werequoted directly from

Lindqvist(Lindqvist, 1960; Kasche and Lindqvist, 1964).

Volume 68 June 1995

(6)

Simulation of "bound fluorescein in

microscopy"

When afluorochrome is boundto asurfaceorencapsulatedinavolume,

thenotionofconcentration isnolongervalid. Anexamplewould be fluo-resceincovalentlyboundtoanucleic acidsequence andtightlyboundtoa cellular target. To simulate this situation where the fluorochrome molecules can no longer diffuse freely, calibrated fluorescent beads from Flow Cytometry Standards Co.wereused.These beadswere9

gm

in diameter anddensely coated with fluoresceinequivalentto2.0 X 106 soluble fluo-rochrome moleculesonaverage(basedonthemanufacturer'sdefinition).

Assumingahomogeneous distribution offluorochrome moleculesoverthe beadsurface, the average intermolecular distance would becomparableto

that of a 10 mM fluoresceinsolution. Thus,forair-saturated aqueous

em-bedding environment, the initial conditions used in the simulation are

No2

= 250,uM and

Ns

= 10mM.

The excitation irradiance undera100-WHgarclamp, 63X/1.4oil ob-jective, and anexcitation bandwidth of 450-490 nm, was measured by

insertinganintegratingspherecontainingasiliconphotodiode(SensorHead Model350, 3MPhotodyne Inc.,Camarillo,CA)under theobjectivelens. The power measured wasconverted,accordingtoEq. 1,to a

k.

of -38s-5

forthismicroscope setup(seealso Jovinet

al.,

1990).Forconfocal laser

scanningfluorescencemicroscopyusing1mWof 488nmargonlaser and anobjective with 1.25 numerical aperture,

ka

is3.8 X 108s-5 (Tsienand Waggoner, 1989).

RESULTS

Experimental result

of free fluorescein

in

solution

The

fluorescence

intensity

measured in

the

spectrofluorom-eter

was

plotted

against time

over

the first

90

min of

bleach-ing

(Fig. 3). Purging

oxygen

by

argon

flushing reduced

the

photobleaching

rate,

but

did not

inhibit it

completely. This

indicated that photo-oxidation

was not

the

only

process that

caused

photobleaching. This experiment

was

repeated for

a

continuous

argon

flush

throughout the 90-min period

to

rule

out

the

possibility

of

02

leaking through

the rubber

stop and

diffusing

back into

the solution.

The result was the same

as

that

of the

single

argon

flush

(data

not

shown).

Both

the air-saturated and argon-flushed

bleaching

rates were

too

slow for the extraction

of a

meaningful characteristic

function.

35 -0 1:1 q) cL) 0 0 0 0 30- 25- 20- 15-

10-n_I

A 0 A 0 + + + + + + A A 0 A A A 0 0

Experimental result of bound fluorescein

in

microscopy

To ensure that the

background

correction

did not introduce

an

offset to the

bleaching

curve

and

thereby

cause a

lack of

compliance with a

single-exponential

function,

a

criterion

was

used such that the sum of the

background pixel

values

after correction

should be as close to zero as possible. With

the method described in the earlier

section, it was found that

the sum of the

background

gray values after the correction

was, on average,

<1% of the total integrated intensity of the

image.

Figs.

4

and 5

are two

typical

examples

of

bleaching

curves

derived from the

bleaching

time series of the

images

after

background

correction. The

integrated

intensity

normalized

by the maximum initial intensity was plotted against time.

The non-random distribution of the

least-square

residuals

indicated

a

deviation of the

experimental

data from a

single-exponential function for both

types of

examples.

This

ex-periment

was

repeated

for a

large

number of beads and

fluo-rescently in situ

hybridized chromosome

centromeres

in

different

embedding

media and

under

different excitation

intensities.

In

all

cases, a

behavior deviating from

a

single-exponential function

was

observed.

Simulation

result of

"free

fluorescein

in

solution"

The

initial conditions used in the simulation

were

selected

so

as

to

resemble

closely the

actual

experimental

conditions.

Fig.

6 a

shows the

simulation result

of

the

population

kinetics

of free

fluorescein

in

air-saturated solution

over a

period

of

>, 1.0-._. " 0.8-0 0 u 0.6-0 '0

E=

0.4 -m0 N Iz 0.2-A A 0 A A

I

0 0 I I I-1 I I I 0 10 20 30 40 50 60 70 80 90 Time(minutes)

FIGURE 3 Effect ofphotobleaching on free fluorescein in solution. 0.01 ,uM fluoresceinsodium/PBS (pH 7.6) was placed in three cuvettes:(+) dark control, not exposed to the bleachinglight source;(A)cappedwith a rubber stopper,argon flushed for 15 min, and exposed to the continuous bleaching

lightsource; and

(0)

air saturatedand exposed to the continuous bleaching

lightsource.Thefluorescence intensity was measured on a

spectrofluorom-eterevery 10 min over a period of 90min.

I.Jn

I I I 1 40 80 120 160 Time

(sec)

0.(

-

1113)

-0.2 0 20 40 I I I --I--- 1 60 80 100 120 140 160 Time(sec)

FIGURE 4 Thephotobleaching curve of a fluorescein surface-labeled

mi-crosphere.The microsphere was embedded in PBS (pH 7.6). The micro-scope setupwasasdescribed in the main text. The images were continuously

acquired,andeachimage was integrated for 0.7 s. In the upper panel, the solidline represents theexperimental data and the broken line represents the

single-exponentialfit. In the lower panel, thecorrespondingresidualplot is shown.

v _1

_ -II

og!:.4

(7)

Volume 68 June 1995

,>

1.0-=.o 0.8-(A 0 0 0 0.6

",

0.;*0-4O j 0.2 -0.0

X

2 _no _ 20 40 60 80 100 120 Time (sec)

'IIIIIIIIIIlI I I III IIIllIIIII ITrIIIII

0 20 40 60 80 100 120

Time(sec)

FIGURE 5 Thephotobleachingcurveof thecentromericregionof

chro-mosome 1inahumanlymphocyte. The humanlymphocyteswereinsitu hybridized and fluoresceinwasdirectly attached(without antibody-antigen

complexorspacermolecules)tothe probesspecificfor the centromeric region of chromosome1.Theinterphase chromosomeswereembedded in PBS(pH 7.6). The imageswerecontinuously acquired withanintegration time of 1.0s/image. For otherimagingparameters,refertothe maintext.

25

h.

Because

of the

very

low

ka

(or

the

low irradiance of the

bleaching light

source),

the

ground state

depletion could

last

many

hours, and the

majority of the

dye

molecules

reside in

S.

The relative S* and T*

populations

were

many

orders

of

magnitude lower

than S. However,

the

population

of T*

was

greater

than S* because of the

accumulation

over

time,

de-spite the

low

irradiance.

The

formation

of X and R

accom-panied

the

depletion

of S

and

was

almost

completely

due

to

the

D-O mechanism because of

the

high

N02

(Fig.

6

c).

The

ground state depletion of the air-saturated fluorescein

solu-tion

was

completely described by

a

single-exponential

func-tion with the least-square residual very

close

to

zero

(Fig.

6

b).

In

the

simulation

of free

fluorescein

in the

argon-flushed

solution

(Fig.

7,

a-c) where

NJ/N02

=

0.4, both D-O and

D-D

mechanisms

operated,

but

the

buildup

of the

bleached dye

population

was

largely

due

to

the

D-O

mechanism.

Because

of the

participation

of the D-D

mechanism,

the

bleaching

curve

of

the simulated

argon-flushed

fluorescein

solution

de-viated

slightly

from

a

single-exponential function. Fig.

8

shows the

comparison

of

the

bleaching

process

for free

fluo-rescein in air-saturated

and argon-flushed solutions. The

in-sert

in

Fig.

8

focuses on the first 90 min and demonstrates

that the

simulation very

closely resembles the bleaching

be-havior observed in the actual experiment

(Fig. 3).

Simulation

result of "bound

fluorescein

in

microscopy"

In

the simulated conventional fluorescence

microscopy case

shown in

Fig.

9, all the

kinetic

curves were

normalized by

the

initial ground

state

population.

The kinetic

curve

of the

ground

state

depletion deviated

significantly

from

a

single-exponential function (Fig. 9 a). This confirms the

experi-mental result

in

microscopy

and indicates a

non-single-exponential

nature

of

the

bleaching

process.

The

buildup

of

the bleached dye population was predominantly due to the

D-D

mechanism with the

D-O

mechanism

contributing

a

small

portion (Fig.

9 a,

right).

If the D-D reactions

were

artificially suppressed

(i.e., k2, k3,

k4, k5,

and

k6were

set

to

0) while other

parameters

were

held the

same as

in

Fig.

9 a,

then

the

kinetic

curve

of the

ground

state

depletion

could be

completely

described by

a

single-exponential

function

(Fig.

9

b).

If

the

D-O

reactions

were

artificially suppressed (i.e.,

k8 and

kg

were set to

0)

while other parameters

were

held the

same as

in

Fig.

9 a,

then the kinetic

curve

of

the

ground

state

depletion

deviated from

a

single-exponential

function

(Fig.

9

c).

If

photobleaching

was

suppressed

altogether

(i.e.,

k2,

k3,

k4, k5, k6, k7,

k8,

and

kg

were set to

0), then the populations

of all

three

energy

states came to a

steady

state

equilibrium

shortly

after the

onset

of

excitation

(Fig.

9

d).

In

the

simulated confocal laser

scanning microscopy,

the

high light

intensity immediately

"pumped" the ground

state

dye molecules

to

the

excited

triplet

state

(Fig.

10

a).

The

high

triplet population greatly

enhanced all the

bimolecular

pro-cesses

involving T*, particularly reactions

4

and 5. The

buildup

of the

bleached

dye in this

case

(Fig. 10 c)

was

for

the

most

part

due

to

the D-D

mechanism.

Taking

the

log

of

the

singlet excited

state

population resulted

in

a

nonlinear

curve

indicating

the

deviation

from

a

single-exponential

function

(Fig.

10

b).

DISCUSSION

Photobleaching via photooxidation

(D-O)

mechanism

Irreversible

loss of

fluorescence

caused

by the

reaction

be-tween

an

oxygen

molecule and

a

dye

molecule

has

been the

subject of

numerous

studies

in

photochemistry.

It

has

also

been

recognized

that

photooxidation

may not

be the only

process

responsible

for

the irreversible

loss of fluorescence.

In

the

present

study,

purging

oxygen

by flushing the sample

solution with argon

only

reduces the photobleaching rate.

This result

is consistent

with the findings of Johnson et al.

(1982).

It

indicates,

on

the one

hand,

that argon flushing may

not

completely

remove

all oxygen molecules and, on the

other

hand, that

there may be at least one other bleaching

process

that is

independent

of oxygen.

It

is

interesting

to note

from Table 1 that, since the

D-O

mechanism consists

of only one pseudo-unimolecular

reac-tion

(and

thus

one

exponential term)

leading to the bleached

dye molecules,

the

bleaching

behavior will be a

single-exponential

process in

the absence of all

D-D

reactions. In

the simulation of fluorescein in

solution,

where

NS

is low and

NS

<<

N,

the

probability of

a

reaction between

dye

mol-ecules

is

very

low. This

probability is

governed only by the

concentrations of

the

reacting

molecules,

since

the rate of

reaction is

constant at a

given temperature. The

photobleach-IT1Itt

(8)

FIGURE 6 Simulation for free fluores-cein in air-saturated solution. The param-etersused for this simulationare:[F]=0.01 ,uM,

[Oj

=250,LM,ka=0.038s-1,and all otherrate constants are asdefined in Table 1. (a)population kineticcurvesof fluores-cein molecules in various energy states. S and X+Rareplotted usingthe left vertical

axis;S* and T*areplottedonanother scale

ontheright.Allpopulationsarenormalized by theinitial dyepopulationattheground

state,wherealldyemoleculesareassumed toresideat t= 0. (b)Kineticcurveof the ground statedepletion ( ), the single-exponential fit (---) along with its least-squareresidual.(c)Theformation ofXand Rdue to different photobleaching mecha-nisms. Here thepopulationsof Xand Rare

grouped(added) together accordingtothe reactions from whichtheyareproduced.

X

0.8

'0. 0.6 : 0.2 - 0.0-0.4 -o fi0.6 -\

X0.4

-\

"0.2 -0

0.0- 2-la r oX. , -10' 0 5 10 15 20 25 Time(hours)

ing

in

this

case

is

primarily caused by

the

D-0

reactions and

shows

a

single-exponential

behavior

(Fig.

6

b).

In

the

simu-lation

of bound

fluorescein

in

microscopy,

if

the

D-D

reac-tions

are

assumed

not to

take

part

in

photobleaching,

then

photobleaching

is

again

a

single-exponential behavior

(Fig.

9

b)

due

solely

to

the

D-0 reactions.

These

simulation results

suggest

that

single-exponential

behavior is

only

a

special

case

of

photobleaching when D-0 reactions

are

the

predomi-nant

mechanism

leading

to

photobleaching.

A

thorough

deoxygenation

of

the

sample environment

may

not

be

completely

advantageous

with

respect to

the

pre-vention of

photobleaching. The complete removal

of oxygen

increases

the

triplet

state

lifetime

(Tsien

and Waggoner,

1989),

increases

the

accumulation of triplet

state

dye

mol-ecules,

and

promotes the bimolecular

reactions via the D-D

mechanism.

It

is important to note that of the two

reactions

involving

the

D-0

mechanism

(physical

quenching in

reac-tion

8,

T*

+

02

->

S

+

02,

and chemical

quenching

in

reaction

9,

T*

+

02

->

X

+

HO2

(Kasche

and

Lindqvist,

1964)),

the

physical

quenching in fact does not lead to

bleached molecules.

Physical quenching enhances T*-S

in-tersystem

crossing

and

returns

T*

to

S at a

higher rate than

chemical

quenching

converts

T* to X. A complete removal

of oxygen will inhibit

physical quenching

and thus fewer

4xlO-9

z 0 3 * N C-0 CD 2 0 0-o 0 M. O 0 Time(hours) 1.0 C 0.8-0.6

8 0.6-

|due

toD-O mechanism

o0.4-due toD-Dmechanism 0.2-z 0 .0 * ..-0 5 10 15 Time(hours) I 20 25

triplet

state

dye molecules

are

reverted back

to

the

ground

state

to

participate further in the excitation-emission cycle.

Photobleaching via

D-D

mechanism

Usui

et

al.

(1965)

provided evidence

of

a

switchover

between

D-0

and D-D

mechanisms depending

upon

the

relative

con-centrations

of

fluorescein

([F])

and of oxygen

([02]).

It

was

concluded

from

their

study

that:

1)

when

[F] is

very

low and

[F]<<

[021

(i.e.,

[F]/[02]

<

0.2),

the

D-0

mechanism

domi-nated,

and

the

bleaching

rate

was

simply proportional

to

the

absorption

of

light;

2) when [F]

>>

[02]

(i.e.,

[F]/[02]

>

0.2),

the

D-D

mechanism

operated

at

a

high

rate,

and the

pro-portionality

between

the

bleaching

rate

and the

absorption

of

light

no

longer

held.

The

results of

Usui

et

al.

(1965)

have an

important

im-plication

for the present

study. As stated earlier, in

spec-troscopy

experiments,

free

fluorescein

molecules are

homo-geneously distributed

in

air-saturated

aqueous

solution. At

low concentration of fluorescein in

air-saturated solution,

NS

<<

No2,

and

the

reaction

is

dominated

by

D-0

reactions.

In

microscopy,

on

the other hand, the

situation

is

different

in that fluorescein

molecules

are no

longer freely

(9)

diffus-Biophysical

Journal 1.0-x 0.8-c0 70.6 - 0-0. I) 04 4-2 0.2 -E z

FIGURE 7 Simulationfor free fluorescein in argon-flushed solution. The parameters usedfor thissimulationwere:[F]=0.01,uM,

[02]=0.025 ,uM,

ka

=0.038

s1',

and allother

rate constants are asdefinedin Table 1. All

populationsarenormalizedby the initialdye populationatthegroundstate.(a)Population kinetic curves of fluorescein molecules in

variousenergy states. SandX+Rareplotted

using theleft verticalaxis;S* and T*are on

anotherscaleontheright axis.Notethat T* is 4 orders ofmagnitude higherthanS*.(b) Kinetic curve of ground state depletion

),thesingle-exponential fit(---)and

its least-square residual.(c) Formation of the bleached dye moleculesdue to D-D and D-0

mechanisms. 15 (a) 0 5 I0 20 Time (hours) _ 1.00 -0.8 0 a.

04

-= 0.4-00 0 5 10 15 (b) 1.0 ~0 (b) CQ 0 v v 0.8 -= 0.6-v X 0.4-00 2--7---- z 20 25 0(.2-.7 r. ()2

()j

0 5 10 15 20 25 Time(hours) (c) |duetoD-0 mechanism| duetoD-Dmechanism ... 5 10 15 Time(hours) 1 20 25 160x10-2 -120 -80 -40 - 0-'\ argon-flushed 0 10 20 30 40 50 60 70 80 90 Time(minutes) Time(hours)

FIGURE 8 Comparisonbetweenthe kinetic

curves of the excited singlet population for free fluorescein in air-saturated and

argon-flushed solutions. Eachpopulationis normal-ized by the initial ground state population. Theparameters used forthis simulationare as

defined in thelegends of Figs. 6 and 7.The insertshowsthesamecurveduring thefirst 90 min forthecomparison with Fig. 3.

50 60

ible. The D-D reaction

probability is

a

function of the

intermolecular distance between

the dye molecules. It is

therefore conceivable that

with [F]

> [02],

the small

in-termolecular distance between

the

dye molecules

(rep-resented

by

a

high initial [F])

promotes

the

proximity-induced D-D

reactions. But

as

reactions proceed, the

intermolecular distance between

the

dye molecules that

can

still

react

will

become

too

large for D-D reactions.

After

that

photobleaching will be dominated by D-0

reactions.

The

switching-over of the D-D and D-0 mechanisms

is

observed in the

current

study, and the effects of each

mechanism

on

the overall

photobleaching behavior

are

also

demonstrated.

In

the simulation

of fluorescein

z - 4 El x ::L 0 -3

<CD

IV 0 0 -2 = x 0 16 -1 -x x 0 _/i _ o 25 180xl012 160

.o

140-a om 120 v 120 0. 0 80 - ~ 60-z 40-z 20 -~~ u--~~-I --- --- -~~~~~~~~ 4'w Volume 68 June 1995 2596 -.w-1

(10)

1.0-< 0.8-0. C. 06-(a) ,/ 0.4-E 0.2 -z 0.] _ 0 20 40 60 80 100 120 0.4 -7 0.0 -04 -0.44-o 1.0-a. s 0.8--a 0.6-m 0.4- 0.2-E o 00 -z due--t -Dmechani

[due

toD-D

mechanism]

Idue

toD-0 mechanism|-0 40 80 120 Time (sec) 0 20 40 60 80 1 00 120 Time (sec) 1.0 0 0.8 X 0.6-X 0.4-E 0.2 -0 z 0.1 C 0. 0' X -O0. -.Ox

O-FIGURE 9 Simulation for bound fluorescein in

con-ventional fluorescence microscopy. The parameters

used for this simulationare:[F]= 10mM,[02]=250

,uM, ka = 38 s-', and all otherrate constants are as defined inTable 1.Thehigh [F]/[O,] ratioisused to

simulate the spatially close intermolecular distance

among thedye molecules.(a) Kineticcurveofground

statedepletion,and itsresidual(left) with the solid line representingthesimulateddata and the brokenline the single exponentialfit; formationof X and R due to

dif-ferentbleaching mechanisms (right). (b)Ground state

kinetics when the D-D reactions are artificially

sup-pressedand other parametersremain thesame asin a.

The simulated dataandthesinglet exponentialfit

over-lapandthe residual isvery small.(c) Groundstate ki-netics when the D-O reactions are artificially

sup-pressedand other parametersremain thesame as in a.

(d) Simulationof an"ideal"caseofnobleaching,where bothD-0and D-Dreactionsareartificially suppressed.

1.0 -a 0 X 0.8-CL X 0.6-cn -0 X 0.4 -._ t 0.2-o z n n_ (b) ,u 0 20 40 60 80 100 120

)0

I 0 20 40 60 80 100 120 Time (sec) 0 (c) I ~~ ~ ~ 1 1 20 40 60 80 100 120 0.4--O0.4 -0.4

~ ~~~

0 20 40 60 80 100 120 Time (sec) 1.00 -.° 0.98 -0. 0.96-M 0.94-_4 I 0.92-E 0 (d) S* T* JV t ° z 0 - 10 m

_l.

- 8 X. - 6 X~ >< 2 -lp 4 cn * >< 0.90 -0 0.0 0.1 0.2 0.3 0.4 0.5 Time(sec) o, u.v 4-.l_ zC

Cytaty

Powiązane dokumenty

The Pal [16, 17] studies on induced dichroism o f starch-iodine-Methylene blue and starch-iodine- Acridine Orange complexes have led to the conclusion that starch-

7 W. Nowak: Politologie struktur edukacyjnych. Turowski: Wielkie struktury społeczne.. W ugruntowywaniu autorytaryzmu ważną rolę odgrywa przestrzeń ideo- logiczna. Wskazuje się w

The term “outcomes” of growth is associated with declining inequality in those non-income dimensions of well-being that are particularly important for promoting

Morawski jako filolog czeka na monografistę, który by opisał jego wpływ na kształt polskiej filologii, wiadomo jednak, że w pierwszym dziesięcioleciu dwudziestego

„Herby miast, gmin i powiatów województwa pom orskiego”, tom I, Oficyna Pom orska, G dańsk 2000.. Uroczystość wręczenia Nagrody odbyła się dnia 30 III

Widma UV-VIS zawiesiny nanocząstek srebra uzyska- nych metodą redukcji chemicznej przy użyciu roztworu za- rodków srebra w ilości 0,6 ml oraz szkieł po krystalizacji zawierających

Inspiracją dla współczesnej teologii przepowiadania stały się przede wszystkim dokumenty Soboru Watykańskiego II. W sposób magistralny istotą słowa Bożego i jego działaniem

Schulz jest, bądź co bądź, autorem krótkiej recenzji książki Aldousa Huxleya M uzyka nocą 9, cytowanej chęt­ nie przez Barbarę Sienkiewicz, gdzie pojawia się wizja