Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Molecular Fluorescence

.pdf
Скачиваний:
72
Добавлен:
15.08.2013
Размер:
7.26 Mб
Скачать

8.4 Methods based on intramolecular excimer formation 235

Fig. 8.3. Excimer-forming bifluorophores for the study of fluidity. 1: a,o-di-(1-pyrenyl)- propane; 2: a,o-di-(1-pyrenyl)-

methylether; 3: 10,100-diphenyl- bis-9-anthrylmethyloxide (DIPHANT); 4: meso-2,4-di- (N-carbazolyl)pentane.

formation on phospholipid bilayers and in particular on phase transition, lateral di usion and lateral distribution (Vanderkoi and Callis, 1974; Hresko et al., 1986).

Molecular mobility in polymer films and bulk polymers can also be probed by excimer formation of pyrene (Chu and Thomas, 1990).

8.4

Methods based on intramolecular excimer formation

Bifluorophoric molecules consisting of two identical fluorophores linked by a short flexible chain may form an excimer. Examples of such bifluorophores currently used in investigations of fluidity are given in Figure 8.3.

In contrast to intermolecular excimer formation, this process is not translational but requires close approach of the two moieties through internal rotations during the lifetime of the excited state. Information on fluidity is thus obtained without the di culty of possible perturbation of the di usion process by microheterogeneity of the medium, as mentioned above for intermolecular excimer formation.

236 8 Microviscosity, fluidity, molecular mobility. Estimation by means of fluorescent probes

Moreover, the e ciency of excimer formation does not depend on the concentration of fluorophores, so that Eq. (8.21) should be rewritten as

 

IE

 

kr0

 

 

 

¼

 

tEk1

ð8:22Þ

IM

kr

 

 

 

 

 

 

As in the case of intermolecular excimer formation, it should be recalled that difficulties may arise from the possible temperature dependence of the excimer lifetime, when e ects of temperature on fluidity are investigated. It is then recommended that time-resolved fluorescence experiments are performed. The relevant equations established in Chapter 4 (Eqs 4.43–4.47) must be used after replacing k1[M] by k1.

The viscosity dependence of intramolecular excimer formation is complex. As in the case of molecular rotors (Section 8.2), most of the experimental observations can be interpreted in terms of free volume. However, compared to molecular rotors, the free volume fraction measured by intramolecular excimers is smaller. The volume swept out during the conformational change required for excimer formation is in fact larger, and consequently these probes do not respond in frozen media or polymers below the glass transition temperature.

Using again the Doolittle equation (8.13) and assuming that the rate constant for excimer formation is given by

 

0

exp x

V0

 

 

ð8:23Þ

 

k1 ¼ k1

 

 

 

Vf

 

(note the analogy with Eq. 8.14), we get

 

 

 

 

 

 

 

 

 

k1 ¼ ah x

 

 

 

 

ð8:24Þ

It is important to note that the response of a probe is not the same in two solvents of identical viscosity but of di erent chemical nature. For instance, the variations in IE=IM for a;o-di-(1-pyrenyl)propane in ethanol/glycerol mixtures and in hexadecane/para n mixtures at 25 C (either degassed or undegassed) show significant di erences between IE=IM values for the same viscosity but with di erent solvent mixtures. A dependence on the chain length was also found. This e ect of chain length and solvent can be explained in terms of the internal rotations involved in excimer formation. These rotations depend on the torsional potential of the various bonds and on the solvent nature; the solvent intervenes not only by the viscous drag, but also by its microstructure and possible interactions with the probe. The distribution of distances between the two fluorophores at the instant of excitation evolves during the excited-state lifetime of the monomer. Some of the bifluorophores with one excited fluorophore form an excimer because of favorable initial interchromophoric distance and favorable time evolution. Once the two chromophores are close to each other, the formation of the excited dimer (excimer)

8.5 Fluorescence polarization method 237

is very rapid, so that the overall process is di usion-dependent. The initial distribution and its further evolution depends on chain length, solvent nature, viscosity and temperature (Viriot et al., 1983).

It should be noted that in organized assemblies, the local order may a ect the internal rotations and the distribution of interchromophoric distances. Therefore, it is not surprising that the values of the equivalent viscosity may depend on the probe and, in particular, on the chain length. A comparison between degassed and undegassed solutions showed a strong e ect of oxygen at low viscosities. The e ect of oxygen leading to smaller values of the ratio IE=IM is simple to understand on the basis of Eq. (8.22): the ratio IE=IM is proportional to the excited-state lifetime of the excimer, which is reduced by oxygen quenching. The rate of oxygen quenching decreases with increasing viscosity, and its e ect on excimer formation in dipyrenylalkanes becomes negligible at viscosities higher than 100 cP.

Intramolecular excimers have been used for probing bulk polymers, micelles, vesicles and biological membranes (Bokobza and Monnerie, 1986; Bokobza, 1990; Georgescauld et al., 1980; Vauhkonen et al., 1990, Viriot et al., 1983; Zachariasse et al., 1983). In particular, this method provides useful information on the local dynamics of polymer chains in the bulk (see Box 8.2).

In conclusion, the method of intramolecular excimer formation is rapid and convenient, but the above discussion has shown that great care is needed for a reliable interpretation of the experimental results. In some cases it has been demonstrated that the results in terms of equivalent microviscosity are consistent with those obtained by the fluorescence polarization method (described in Section 8.5), but this is not a general rule. Nevertheless, the relative changes in fluidity and local dynamics upon an external perturbation are less dependent on the probe, and useful applications to the study of temperature or pressure e ects have been reported.

8.5

Fluorescence polarization method

In Chapter 5, devoted to fluorescence polarization, it was shown that information on the rotational motions of a fluorophore can be obtained from emission anisotropy measurements. Application to the evaluation of the fluidity of a medium, or molecular mobility, is presented below.

8.5.1

Choice of probes

The probes to be used in fluorescence polarization experiments should fulfil as far as possible the following requirements:

. minimum disturbance of the medium to be probed;

.symmetry and directions of the transition moments such that rotations can be considered as isotropic;

238 8 Microviscosity, fluidity, molecular mobility. Estimation by means of fluorescent probes

Box 8.2 Intramolecular excimer fluorescence for probing the mobility of bulk polymers

The method is based on the fact that the rate of conformational change required for excimer formation depends on the free volume induced by the segmental motions of the polymer occurring above the glass transition. DIPHANT (compound 3 in Figure 8.3) was used as an excimer-forming probe of three polymer samples consisting of polybutadiene, polyisoprene and poly(dimethylsiloxane).a)

Figure B8.2.1 shows the fluorescence spectra of DIPHANT in a polybutadiene matrix. The IE=IM ratios turned out to be significantly lower than in solution, which means that the internal rotation of the probe is restricted in such a relatively rigid polymer matrix. The fluorescence intensity of the monomer is approximately constant at temperatures ranging from 100 to 20 C, which indicates that the probe motions are hindered, and then decreases with a concomitant increase in the excimer fluorescence. The onset of probe mobility, detected by the start of the decrease in the monomer intensity and lifetime occurs at about 20 C, i.e. well above the low-frequency static reference temperature Tg (glass transition temperature) of the polybutadiene sample, which is 91 C (measured at 1 Hz). This temperature shift shows the strong dependence of the apparent polymer flexibility on the characteristic frequency of the experimental technique. This frequency is the reciprocal of the monomer excited-state

Fig. B8.2.1. Temperature dependence of the fluorescence spectrum of DIPHANT dispersed in polybutadiene. Insert: temperature evolution of the fluorescence intensity of monomer and excimer (reproduced with permission from Bokobza and Monneriea)).

8.5 Fluorescence polarization method 239

lifetime (experimental window for excimer formation): @1:5 108 s 1 for DIPHANT, and @5 106 s 1 for pyrene-based bifluorophores (whose lifetimes are of the order of 200 ns).

The correlation time tc of the motions involved in intramolecular excimer formation is defined as the reciprocal of the rate constant k1 for this process. Its temperature dependence can be interpreted in terms of the WLF equationb) for polymers at temperatures ranging from the glass transition temperature Tg to roughly Tg þ 100 :

ln

tcðTÞ

 

C1ðT TgÞ

tcðTgÞ

¼ C2 þ ðT TgÞ

 

where C1 and C2 are two constants that depend on the chemical structure of the polymer. Their values were selected to be those given by Ferryc) for elastomers. The values of the rate constant k1 were determined by time-resolved fluorescence measurements, and the values of log tc were plotted as a function of T Tg (Figure B8.2.2). The data reveal an important e ect of the chemical structure of the matrix and a good agreement with the WLF equation in all cases.

Fig. B8.2.2. Logarithmic plot of the correlation time versus T Tg for DIPHANT dispersed in polybutadiene (PB), polyisoprene (PI) and poly(dimethylsiloxane) (PDMS). The broken lines are the best fits with the WLF equation (reproduced with permission from Bokobza and Monneriea)).

240 8 Microviscosity, fluidity, molecular mobility. Estimation by means of fluorescent probes

This investigation shows that it is indeed possible to study the flexibility of polymer chains in polymer matrices by means of excimer-forming probes and that the rotational mobility of these probes reflect the glass transition relaxation phenomena of the polymer host matrix, in agreement with the appropriate WLF equation.

a)Bokobza L. and Monnerie L. (1986), in: Winnik M. A. (Ed.), Photophysical and Photochemical Tools in Polymer Science, D. Reidel Publishing Company, New York, pp. 449–66.

b)Williams M., Landel R. F. and Ferry J. D.

(1955) J. Am. Chem. Soc. 89, 962. The WLF equation is commonly used for mechanical relaxation data analysis at low frequency (< 106 Hz).

c)Ferry J. D. (1970) Viscoelastic Properties of Polymers, Wiley, New York.

. minimum specific interactions with the surrounding molecules;

.minimum sensitivity of the excited-state lifetime to the microenvironment when only steady-state anisotropy is measured.

Figure 8.4 shows examples of some probes. Regarding disturbance of the medium, small molecules are to be preferred, but the measurement is then often sensitive to specific interactions with the surrounding molecules. Note that lipid-like probes such as parinaric acids o er minimum disturbance for the investigation of lipid bilayers. With large probes, these interactions have minor e ects, as demonstrated in the case of the large probe BTBP [N,N 0-bis(2,5-di-ter-butylpheny)- 3,4,9,10-perylenetetracarboximide] by the linear relationship observed between rotational correlation time and bulk viscosity of the solvent (n-alkanes, n-alcohols, ethanol/glycerol mixtures, para n oil/dodecane mixtures) in the 0.5–150 cP range. However, such a large probe is not suitable for the study of confined media of small dimensions.

8.5.2

Homogeneous isotropic media

The case of probes undergoing isotropic rotations in a homogeneous isotropic medium will be examined first. Rotations are isotropic when the probe has a spherical shape, but it is di cult to find such probes1) because most fluorescent molecules are aromatic and thus more or less planar. Nevertheless, when a probe interacts with the solvent molecules through hydrogen bonds, experiments have shown that in some cases the observed rotational behaviour can approach that of a sphere. In the case of rod-like probes whose direction of absorption and emission transition moments coincide with the long molecular axis (e.g. diphenylhexatriene; Figure 8.4), the rotations can be considered as isotropic because any rotation about this long axis has no e ect on the emission anisotropy.

1) Fullerene C60 is perfectly spherical but for

of rotational motions. Moreover, its

symmetry reasons, the emitted fluorescence

fluorescence quantum yield is very low

is completely depolarized even in the absence

ð@

10 4

Þ

.

2

 

8.5 Fluorescence polarization method 241

Fig. 8.4. Examples of probes used in fluorescence polarization experiments to evaluate fluidity and molecular mobility. 1: 1,6-diphenyl-1,2,5-hexatriene (DPH). 2: 1-(4-trimethyl- ammoniumphenyl)-6-phenyl 1,3,5-hexatriene, p-toluene sulfonate (TMA-DPH). 3: cisparinaric acid. 4: trans-parinaric acid. 5. BTBP [N,N0-bis(2,5-di- ter-butylphenyl)-3,4,9,10- perylenetetracarboximide].

For isotropic motions in an isotropic medium, the values of the instantaneous and steady-state emission anisotropies are linked to the rotational di usion coe - cient Dr by the following relations (see Chapter 5):

 

rðtÞ ¼ r0 expð 6Dr

 

ð8:25Þ

 

 

 

 

 

 

 

 

 

 

 

1

1

ð1

þ 6DrtÞ

 

 

ð8:26Þ

 

 

 

¼

 

 

 

 

r

r0

 

 

 

 

 

 

 

 

 

 

 

 

which

allow

determination of Dr. In principle, a value of the

viscosity h

ð¼ kT=6VDrÞ could be calculated from the Stokes–Einstein relation (8.3) provided that the hydrodynamic volume V is known. In addition to the problem of validity of the Stokes–Einstein relation (see above), the hydrodynamic volume cannot be calculated on a simple geometrical basis but must also take into account the solvation shell.

Introducing the rotational correlation time tc ¼ ð6DrÞ 1, Eqs (8.25) and (8.26)

can be rewritten as

 

rðtÞ ¼ r0 expð t=tcÞ

ð8:27Þ

242 8 Microviscosity, fluidity, molecular mobility. Estimation by means of fluorescent probes

1

1

1 þ

t

 

ð8:28Þ

 

¼

 

 

r

r0

tc

The changes in correlation time upon an external perturbation (e.g. temperature, pressure, additive, etc.) reflects well the changes in fluidity of a medium. It should again be emphasized that any ‘microviscosity’ value that could be calculated from the Stokes–Einstein relation would be questionable and thus useless.

8.5.3

Ordered systems

Equations (8.25) to (8.28) are no longer valid in the case of hindered rotations occurring in anisotropic media such as lipid bilayers and liquid crystals. In these media, the rotational motions of the probe are hindered and the emission anisotropy does not decay to zero but to a steady value ry (see Chapter 5). For isotropic rotations (rod-like probe), assuming a single correlation time, the emission anisotropy can be written in the following form:

rðtÞ ¼ ðr0 ryÞ expð t=tcÞ þ ry

ð8:29Þ

Time-resolved emission anisotropy experiments provide information not only on the fluidity via the correlation time tc, but also on the order of the medium via the ratio ry=r0. The theoretical aspects are presented in Section 5.5.2, with special attention to the wobble-in-cone model (Kinosita et al., 1977; Lipari and Szabo, 1980). Phospholipid vesicles and natural membranes have been extensively studied by time-resolved fluorescence anisotropy. An illustration is given in Box 8.3.

8.5.4

Practical aspects

From a practical point of view, the steady-state technique (continuous illumination) is far simpler than the time-resolved technique, but it can only be used in the case of isotropic rotations in isotropic media (Eqs 8.26 and 8.28) provided that the probe lifetime is known. Attention should be paid to the fact that the variations in steadystate anisotropy resulting from an external perturbation (e.g. temperature) may not be due only to changes in rotational rate, because this perturbation may also a ect the lifetime.

The time-resolved technique is much more powerful but requires expensive instrumentation.

It is worth pointing out that many artifacts can alter the measurements of emission anisotropy. It is necessary to control the instrument with a scattering non-fluorescent solution (r close to 1) and with a solution of a fluorophore with a long lifetime in a solvent of low viscosity (r A0). It is also recommended that the probe concentration is kept low enough to avoid interaction between probes.

8.5 Fluorescence polarization method 243

Box 8.3 Investigation of the dynamics and molecular order of phosphatidylinositol incorporated into artificial and natural membranesa)

Lipid–protein interactions are of major importance in the structural and dynamic properties of biological membranes. Fluorescent probes can provide much information on these interactions. For example, van Paridon et al.a) used a synthetic derivative of phosphatidylinositol (PI) with a cis-parinaric acid (see formula in Figure 8.4) covalently linked on the sn-2 position for probing phospholipid vesicles and biological membranes. The emission anisotropy decays of this 2-parinaroyl-phosphatidylinositol (PPI) probe incorporated into vesicles consisting of phosphatidylcholine (PC) (with a fraction of 5 mol % of PI) and into acetylcholine receptor rich membranes from Torpedo marmorata are shown in Figure B8.3.1.

The curves were fitted using the following decay function, which must be considered as a purely mathematical model:

rðtÞ ¼ a1 expð t=tc1Þ þ a2 expð t=tc2Þ þ ry

where tc1 and tc2 are the correlation times for rotational di usion of the probe and ry is the residual anisotropy at long time with respect to the correlation times.

The parameters a1; a2; tc1; tc2, and ry are obtained from the best fit of IkðtÞ and I?ðtÞ given by Eqs (5.7) and (5.8) of Chapter 5:

IðtÞ

IkðtÞ ¼ 3 ½1 þ 2rðtÞ&

IðtÞ

I?ðtÞ ¼ 3 ½1 rðtÞ&

where IðtÞ is the total fluorescence intensity decay ½¼ IkðtÞ þ 2I?ðtÞ& that can be satisfactorily fitted in this case by a sum of two exponentials.

Then, according to the wobble-in-cone model (see Section 5.6.2), the order parameter that is related to the half angle of the cone, and the wobbling di usion constant (reflecting the chain mobility) can be determined.

For comparison, 2-parinaroyl-phosphatidylcholine (PPC) was also incorporated into the membrane preparations. For the Torpedo membranes, the acyl chain order measured by PPI was found to be lower than that by PPC, whereas the opposite was true for the vesicles. This inversion strongly suggests that PI has di erent interactions with certain membrane components compared to PC. In contrast, the correlation times of PPI and PPC were only slightly di erent in Torpedo membranes, and the values showed little di erence with those measured in vesicles.

244 8 Microviscosity, fluidity, molecular mobility. Estimation by means of fluorescent probes

Fig. B8.3.1. Fluorescence anisotropy decays at

half angles of the cone are found to be 61

4 C of PPI. A: in phospholipid vesicles (PC:PI,

(A) and 45 (B), and the wobbling di usion

95:5 mol %). B: in Torpedo membranes. From

constants are 0.041 ns 1 (A) and 0.020 ns 1

the best fit of the I t

and I

 

(reproduced with permission from van

ð Þ

?ðtÞ components,

(B)

 

a)

).

and by using the wobble-in-cone model, the

Paridon et al.

 

a)van Paridon P. A., Shute J. K., Wirtz K. W. A. and Visser A. J. W. G. (1988) Eur. Biophys. J. 16, 53–63.

Соседние файлы в предмете Химия