You are on page 1of 10

ARTICLE IN PRESS

Journal of Luminescence 121 (2006) 431–440


www.elsevier.com/locate/jlumin

Micellar effects on the molecular aggregation and fluorescence


properties of benzazole-derived push–pull butadienes
Tarek A. Fayed, Safaa El-Din H. Etaiw, Naglaa Z. Saleh
Chemistry Department, Faculty of Science, Tanta University, 31527-Tanta, Egypt
Received 8 August 2005; accepted 18 November 2005
Available online 4 January 2006

Abstract

The absorption and fluorescence spectral behaviour of 1-(2-benzoxazolyl)-4-(p-dimethylaminophenyl)buta-1,3-diene


and its benzothiazolyl analogue (abbreviated as BODB and BTDB, respectively) have been investigated in
dioxane–water mixtures and micellar environments using steady-state techniques. In water, water-rich mixtures or
premicellar solutions, BODB and BTDB tend to form molecular aggregates labelled as H-aggregates. These aggregates
dissociate on adding surfactants forming micellized monomers. In all micellar media the fluorescence quantum yield is
greatly enhanced along with a large hypsochromic shift. Also, in TX-100, CTAB and SDS both dienes show dual
emission from the locally excited (LE) and intramolecular charge transfer (ICT) states. For BTDB in TX-100 and
DODB in all solutions, the LE fluorescence predominates, while for BTDB in CTAB and SDS, the ICT fluorescence is
the predominant. The fluorescence shifts suggest that the fluorescing molecules penetrate the core of the micellar unit in
TX-100, whereas in CTAB and SDS they occupy the interfacial region. The binding constants and the micelle properties
(such as polarity and CMC) have been determined using both dienes as probes.
r 2005 Elsevier B.V. All rights reserved.

Keywords: Diarylbutadienes; Benzazole; Dual fluorescence; Micelles; Probes; H-aggregates

1. Introduction conformationally relaxed intramolecular charge


transfer (ICT) or twisted ICT (TICT) excited
A variety of organic molecules containing states as emissive states [1–3]. The high sensitivity
donor–acceptor moieties exhibit fluorescence pro- of these states to solvation, polarity and molecular
perties that depend on the physical and chemical mobility changes render such molecules to be effi-
nature of the medium. This is due to population of cient reporters (provide a wealth of information)
of their microenvironments. Push–pull stilbenes,
Corresponding authors. Tel.: +20403120708;
i.e. stilbenes substituted with an electron donor
and an acceptor group in conjugated positions,
fax: +20403350804.
E-mail addresses: tfayed2003@yahoo.co.uk (T.A. Fayed), have attracted great interest in view of their fluore-
safaaehe@gega.net (S.H. Etaiw). scence sensing properties [2,4–7]. Donor–acceptor

0022-2313/$ - see front matter r 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.jlumin.2005.11.006
ARTICLE IN PRESS

432 T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440

styryl derivatives have been fruitfully employed in of BODB and BTDB have been investigated in 1,4-
cellular calcium sensing [8] and in the visualization dioxane–water binary mixtures and in micro-hetero-
of membrane nerve potential [9]. Also, they are geneous media of sodium dodecyl sulphate (SDS),
utilized in structural investigations of systems cetyltrimethyl ammonium bromide (CTAB) and
including micelles [10,11], vesicles [12] and poly- Triton X-100 (TX-100) as examples of anionic,
mers [13,14]. cationic and neutral surfactants, respectively. This is
Diphenylbutadienes are interesting models of to examine the effect of micelles on the diene H-
the retinyl polyenes that are involved in biological aggregates formed in aqueous media. The study aims
sensory and energy transduction [15]. Recent also to explore the usefulness of these dienes as
studies with donor–acceptor-substituted diphenyl- fluorescence probes for characterization of the
butadienes [14,16–18] have shown that these micelle properties.
compounds are also capable of exhibiting fluores-
cence from conformationally relaxed ICT excited
states. Therefore, these compounds have been used 2. Experimental
to characterize the microenvironment of micelles
[16,19,20] as well as probes for saccharides [21] and The investigated diarylbutadienes (BODB and
metal ions [22]. However, nature and dynamics of BTDB) were prepared and characterized as
excited state of diphenylpolyenes are not clearly reported in a previous study [23]. SDS and CTAB
known and the sensing applications are very were from BDH while TX-100 from Aldrich were
limited. used without purification. Double distilled water
In this respect, very recently we have synthesized was used for preparation of micelle solutions
and investigated the photophysical properties of whereas highly spectroscopic-grade ethanol and
two new diphenylbutadiene derivatives [23], viz, 1- dioxane (Merck) were used for preparation of
(2-benzoxazolyl)-4-(p-dimethylaminophenyl)buta- binary mixtures.
1,3-diene and its benzothiazolyl analogue, abbrevi- The absorption spectra were measured on a
ated as BODB and BTDB, respectively, Scheme 1. Shimadzu UV-3101PC scanning spectrophot-
The absorption and fluorescence characteristics of ometer. The steady-state fluorescence measure-
these derivatives exhibit large solvent polarity and ments were preformed using a Perkin-Elmer LS
viscosity as well as acidity-dependent changes. 50B spectrofluorimeter fitted with a temperature
Hence, they could be utilized as sensors for the controlled unit (Fischer Scientific Isotherm Refri-
microenvironment of organized assemblies. gerated Circular Model 9000). The fluorescence
Surfactant molecules aggregate in water with quantum yield (ff) was measured relative to
their polar head-groups pointing toward water to quinine sulphate (ff ¼ 0:54 in 0.1 M H2SO4) [24].
form micelles. These aggregates are interesting For all measurements, 2.0  105 M diene solu-
models of biological systems. Thus, in the present tions were used and handled under dim light to
work, the absorption and fluorescence properties avoid the trans/cis photoreaction. Since the dienes
are sparingly soluble in water, their stock solutions
were made in ethanol and added to the appro-
priate amounts of surfactant to make different
X concentrations (the ethanol content is less than
2%). The surfactant concentration was varied
N within the range 0.1–12.5 mM depending on the
nature of the used surfactant. For recording
NMe2
fluorescence spectra of BODB and BTDB in
X = O, (BODB) mixed solvents or micellar solutions, the samples
were excited at the maximum of the long-
X = S, (BTDB)
wavelength absorption band to ensure excitation
Scheme 1. of monomers.
ARTICLE IN PRESS

T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440 433

3. Results and discussion arrangement of molecules [25]. The theory predicts


that the strong coupling of the several similar
3.1. Molecular aggregation of the investigated monomers in the J- and H-aggregates results in a
dienes in aqueous solutions red and blue shift of their absorption bands,
respectively, relative to the monomer [25]. In
The structure and spectroscopy of molecular addition, the spectrum gets narrower due to the
aggregates in which the molecular arrangement is absence of vibrational coupling to the molecular
highly ordered (namely; J- and H-aggregates) are mode [25].
of much interest because of their special properties The absorption and fluorescence spectra of
and possible technological applications [25–27]. BODB in water, ethanol and dioxane are shown
Briefly, J-aggregates are one-dimensional side- in Fig. 1, while the spectral peak positions of both
by-side while the H-aggregates are face-to-face dienes in the mentioned solvents as well as in
dioxane–water mixtures are collected in Table 1.
Fluorescence intensity (a. u.) There are significant variations in the shape and
spectral peak position on going from organic
1.2 solvents to water. The broad absorption band of
DODB that appears at 417 or 411 nm in ethanol
Absorbance

0.8 and dioxane, respectively, is replaced by a narrow


band at 354 nm in water, with a broad tail
extending to over 500 nm. The spectral behaviour
0.4 of BTDB is similar except for red shifts in the
absorption maximum. Thus, based on the shape of
0.0 the absorption band, the large hypsochromic shift
300 400 500 600 700 and narrowing of the band in water, is due to
Wavelength (nm) formation of H-aggregates [25,26] while the broad
Fig. 1. Absorption and fluorescence emission spectra of BODB tail was attributed to absorption of monomers.
in: dioxane (- - -), ethanol (y.)and water (—). Excitation The very low solubility of the investigated dienes
wavelength is 415 nm. in water makes it difficult to study the effect of

Table 1
Polarity parameters of dioxane–water mixtures as well as the absorption and emission maxima (nm) of the dienes measured in these
mixtures

Dioxane % (v/v) e (D) ET(30) BODB BTDB


(kcal mol1)
la lf la lf

0 78.48 63.6 354 (417) 580 (531) 363 (422) 600 (560)
10 70.33 — 357 578 368 588
20 61.86 — 360 573 368 586
30 53.28 — 415 567 419 585
40 44.54 55.8 416 563 423 582
50 35.85 — 418 553 427 579
60 27.21 52.1 419 547 427 573
70 19.07 — 418 537 426 569
80 11.86 49.2 416 531 424 562
90 6.07 46.3 414 519 421 550
95 3.89 43.8 411 514 420 537
100 2.21 36.6 408 503 416 516

Values in parenthesis are the maxima in ethanol.


ARTICLE IN PRESS

434 T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440

concentration on the absorption spectra. To get fluorescence spectra of both dienes, recorded in
information about the effect of solvents on the water, are independent of the excitation wave-
aggregation of the used dienes, their absorption length and the intensity decreases strongly upon
spectra have been studied in aqueous binary excitation at lexc o370 nm (not shown here). Thus,
mixtures containing dioxane or ethanol. The it was concluded that the formed aggregates are
absorption spectra of BODB in various dioxane–- non-fluorescent and the emission in water is due to
water mixtures are depicted in Fig. 2a as a monomers. Consequently, the strong red shift is
representative example. Similar behaviour was due to stabilization of the charge transfer excited
observed in the case of BTDB. As the water state of BODB and BTDB in the polar media [23].
content increases, the long-wavelength absorption The fluorescence spectra of both dienes (lexc ¼
band suffers a pronounced red shift (ca. 10 nm) 410 nm) are strongly influenced by changing the
relative to that in dioxane due to increase in the water content in dioxane. The fluorescence max-
polarity of the medium. This shift is accompanied ima are highly red shifted (ca. 73 and 64 nm for
by a decrease in the absorbance and build-up of BODB and BTDB, respectively) when the solvent
the H-aggregates narrow band at the shorter is changed from dioxane to 80% water in the
wavelength side with appearance of a well-defined dioxane mixture, Table 1. In addition, the spectra
isosbestic point. The appearance of an isosbestic exhibit dual emission where two overlapping
point indicates the monomer-H-aggregate equi- bands appear in dioxane–water binary mixtures
librium. Using a protic solvent such as ethanol as a containing 20–80% water. Fig. 3 shows such
cosolvent instead of dioxane results in similar spectral changes for BTDB in dioxane–water
spectral changes, see Fig. 2b. So, attribution of the mixtures. One possible explanation for the ob-
blue-shifted absorption band to hydrogen-bonded served dual emission is the participation of a polar
complex formation (due to interaction between ICT excited state and/or formation of a hydrogen-
–NMe2 group and water molecules) was ruled out. bonded complex in addition to the local excited
Also, it indicates that the self-association of the state (LE). However, contribution of emission
present dienes takes place only in pure or water- from a hydrogen-bonded complex was ruled out
rich solvent mixtures. since both dienes exhibit dual emission also in
The fluorescence emission of both dienes was aprotic solvents, e.g. DMF [23]. The fluorescence
also studied in the mentioned solvents and maximum has been correlated with the dielectric
dioxane–water mixtures. As shown in Fig. 1, the constant and the empirical polarity parameter
fluorescence spectrum of BODB is red shifted on ET(30) of the used dioxane–water mixtures [28,29].
going from dioxane to water. However, the The ET(30) correlation plots are linear (r40:97)

Dioxane
1.2 % of H2O (v/v)
1.2 95.0 %(v/v)
70.0 0.0
100.0 50.0 50
Absorbance

Absorbance

0.8 40.0 0.8 70


30.0 75
20.0 80
10.0 90
0.4 0.4
0.0 100

0.0 0.0
300 350 400 450 500 300 400 500 600
(a) Wavelength (nm) (b) Wavelength (nm)

Fig. 2. UV–visible absorption spectra of BODB in: (a) dioxane–water and (b) ethanol–water mixtures. The composition of the
mixtures is mentioned in the figure.
ARTICLE IN PRESS

T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440 435

400
Dioxane (v%)
Fluorescence intensity (a. u.)

100 BTDB
90
80
70
50
200
30
20
0

0
450 500 550 600 650 700
Wavelength (nm)

Fig. 3. Fluorescence spectra of BTDB in dioxane–water Fig. 5. Effect of CTAB concentration on the absorption
mixtures. The composition of the mixtures is shown in the spectrum of BODB. Inset is the surfactant concentrations.
figure.

620 Table 2
Spectral maxima (nm) and fluorescence quantum yields of the
Fluorescence maximum (nm)

600 dienes in SDS, CTAB and TX-100 (12.5, 2.5 and 1.25 mM,
respectively)
580
Medium BODB BTDB
560 la lf ff la lf ff

540 SDS 418 532, 565 0.0085 422 533, 583 0.0093
CTAB 424 532, 562 0.0101 435 534, 582 0.0115
520 TX-100 421 532, 543 0.019 431 531, 567 0.013
Water 354 580 o103 363 600 o103
500
0 10 20 30 40 50 60 70 80
Dielectric constant or ET(30) solutions having different concentrations of the
three surfactants. Representative spectra are pre-
Fig. 4. Plots of the fluorescence maximum of BODB (closed
sented in Fig. 5, and the spectral data are
symbols) and BTDB (open symbols) versus dielectric constant
(squares) and ET(30), (circles) of the dioxane–water mixtures. summarized in Table 2. On adding CTAB, as an
example, the absorbance of the H-aggregates
whereas those of the dielectric constant are non- narrow band (at 354 and 363 nm for BODB
linear, Fig. 4. Both correlations show that the and BTDB, respectively) diminishes and a broad
fluorescence maximum is red shifted by increasing absorption band appears at longer wavelengths
the polarity of the medium. Therefore, the large (around 424 and 435 nm, respectively) with distinct
red-shifted fluorescence band is attributed to a isosbestic point. Similar changes were also ob-
polar ICT excited state, which is stabilized in served in SDS and TX-100. These results indicate
water-rich mixtures. that the addition of surfactants promotes disag-
gregation of BODB and BTDB, and the resulting
diene monomers are solubilized by the surfactant
3.2. Disaggregation and fluorescence spectra of the molecules to give stable structures (micellized
investigated dienes in micellar solutions monomers). At relatively higher surfactant con-
centrations (beyond critical micelle concentration,
The absorption spectra of BODB and BTDB CMC) almost all diene molecules are present as
(ca. 2  105 M) have been recorded in aqueous micellized monomers as evident from the absence
ARTICLE IN PRESS

436 T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440

400
Fluorescence intnsity (a. u.)

BODB
CTAB 200

Intensity (a. u.)


TX-100
200
100
SDS

H2O

0 0
300 350 400 450 500 550 600
450 500 550 600 650 700
Wavelength (nm)
Wavelength (nm)

400 Fig. 7. Fluorescence excitaion spectra of BTDB measured in


CTAB (- - -) and SDS (—) solutions as in Fig. 6, monitored at
Fluorescence intensity (a. u.)

BTDB
lem ¼ 530 and 580 nm, lower and higher intensities, respec-
tively.
CTAB

200 0.088 and 0.169 ns [30]. Unfortunately, the fluor-


TX-100 escence decay could not be measured in dioxane–
water mixtures or micellar solutions due to
SDS
technical reasons. Based on the insensitivity of
H2O
the excitation spectra to the monitoring wave-
0 length and the biexponential decay of fluorescence
450 500 550 600 650 700
in ethanol and DMF, it was suggested that the
Wavelength (nm)
dual emission is due to participation of an ICT
Fig. 6. Fluorescence spectra of BODB (upper panel) and excited state in addition to LE one [23]. In
BTDB (lower panel) recorded in micellar solutions with addition, the maximum of the short-wavelength
concentrations of 12.5, 2.5 and 1.25 mM for SDS, CTAB and emission is independent of the polarity of the used
TX-100, respectively.
micelle (Table 2) while that of the long wavelength
one is strongly red shifted on going from TX-100
of the narrow absorption band characteristic for to CTAB or SDS, which justifies our conclusion.
the H-aggregates. In SDS and CTAB aqueous solutions, the ICT
The fluorescence spectra of BODB and BTDB in fluorescence of BTDB (around 580 nm) predomi-
the three surfactants (above CMC) are shown in nates with the fluorescence from the LE state as a
Fig. 6. The emission profile shows two overlapping shoulder around 530 nm. In case of TX-100, the
bands whose intensities and maxima change on LE fluorescence with maximum at 530 nm is
going from water to anionic, cationic and neutral predominant while the ICT fluorescence appears
micelles. The fluorescence maxima and quantum as a shoulder around 565 nm. In contrast, the LE
yields are given in Table 2. In addition, the fluorescence of BODB predominates in all micellar
excitation spectra monitored at both the emission media, and the ICT fluorescence appears as a
maxima in CTAB and SDS for BTDB as shoulder at longer wavelengths.
examples, Fig. 7, are independent on the emission Fig. 8 illustrates the fluorescence spectra of
wavelength. This indicates that only one species is BTDB in aqueous TX-100 micellar solutions as a
responsible for the observed dual emission. The function of the surfactant concentration. Gener-
fluorescence decay profiles of BTDB in ethanol ally, gradual addition of CTAB or TX-100 is
and DMF follow biexponential fits suggest emis- associated with a hypsochromic shift of the
sion from two states. The lifetime values in ethanol emission maximum along with a great enhance-
are 0.028 and 0.102 ns while those in DMF are ment of fluorescence intensity. Both observations
ARTICLE IN PRESS

T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440 437

500
BTDB [TX-100]x10-5 ,M [SDS]x10-3, M
220

Fluorescence intensity (a.u.)


Fluorescence intensity (a.u.)

125 4.0
85 3.0
60 5.0
50 2.0
40 250 7.5
110 12.5
30
20 1.0
10 0.5
0 0.0

0 0
500 600 700 500 600 700
wavelength (n.m) Wavelength (nm)

Fig. 8. Fluorescence spectra of BTDB measured in TX-100 Fig. 9. Fluorescence spectra of BODB measured in SDS
micellar solutions, lexc ¼ 431 nm. The surfactant concentra- micellar solutions, lexc ¼ 418 nm. The concentrations of SDS
tions are shown in the figure. are shown in the figure.

reflect that the microenvironment around the small premicellar aggregates, in such a way that
dienes in micellar solutions is quite different from their exposure to water is minimized [10] hence
that in pure aqueous phase. The blue shift in the sense a less polar environment. On the other hand,
fluorescence spectra as well as the great enhance- the fluorescence quenching and the observed red
ment of the fluorescence intensity within the shift near and above CMC are due to changes
micellar solutions indicates that the polarity sensed in the shape and size of the micelle. This is
by BODB and BTDB is less than the polarity of also consistent with the fact that the micellar
bulk water. It was reported that the fluorescence environment of SDS is richer in water content
yield of the present dienes decreases with increas- than that in CTAB and TX-100 [31] and the
ing the solvent polarity [23]. In addition, the dye molecules occupy the interfacial region, vide
enhancement of fluorescence quantum yield in infra. So, the red shift of the fluorescence is due
micelles, Table 2, can be attributed to restrictions to increased polarity of the SDS micelle–water
imposed on the radiationless free rotational interface.
motions (mainly twisting around the CQC
bonds) and/or disaggregation of the non-fluores- 3.3. Probing of the micelle properties
cent H-aggregates.
Away from CTAB and TX-100, the fluorescence The micropolarity [expressed as the dielectric
spectra of both dienes show anomalous behaviour constant (e) and ET(30)] as well as CMC of the
on changing the concentration of SDS, Fig. 9. In used micelles has been determined by using BODB
solutions containing lower SDS concentrations and BTDB. The obtained values are listed in
(o4.0 mM), the fluorescence spectra suffer a Table 3. The dielectric constant and ET(30) of the
strong blue shift (ca. 58 nm) along with a great micelle–water interface in SDS, CTAB and TX-
enhancement in intensity of the LE emission. On 100 micelles were estimated by using Fig. 4 and the
adding more SDS, the fluorescence intensity measured ICT fluorescence maximum (Table 2).
decreases with developing of the ICT emission Table 3 gives also the polarity parameters of these
and a red shift in the emission maximum (ca. 18 micelles as determined by some other probes for
and 50 nm for BODB and BTDB, respectively). comparison [32,33]. The values indicate that the
The great enhancement in the fluorescence inten- dienes occupy interfacial region, which may be due
sity (more than 10 fold) as well as the blue shift, at to presence of the dimethylamino group. Also, the
lower SDS concentrations, is attributed to associa- data show that the polarity of the interface is
tion of dienes with free SDS molecules and/or dependent on the nature of the surfactant where it
ARTICLE IN PRESS

438 T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440

Table 3
Polarity (e, D, and ET(30), kcal/mol) and critical micelle concentrations (CMC, mM) of SDS, CTAB and TX-100 estimated using
BODB and BTDB as fluorescence probes

Probe parameter BODB BTDB

SDS CTAB TX-100 SDS CTAB TX-100

e 47.6 (40) 43.6 (36) 22.8. (28) 44.7 40.5 19.0


ET(30) 58.0 (57.9) 57.0 (56.3) 51.8 (55.6) 56.2 55.1 52.3
CMC 6.5, 3.95 (8.0) 0.89 (0.9) 0.28 (0.3) 5.7, 4.0 0.86 0.25

Values in parenthesis were determined by using 4-aminophthalimide as a probe [32].

increases from neutral to cationic and anionic 12


SDS
micelles. Furthermore, the values obtained by both
probes are in fair agreement.
The fluorescent dienes studied herein have been 8
also employed to determine the CMC for SDS, o
If /If
CTAB and TX-100. This was achieved by plotting
CTAB
the ratios of fluorescence intensities (in the absence
and presence of surfactants) against the concen- 4
tration of surfactants, Fig. 10. The CMC values of
CTAB and TX-100 were calculated from the break
at the abrupt increase of the fluorescence intensity, 0
and the calculated values are given in Table 3. The 0 2 4 6 8 10 12 14 16
obtained values are in good agreement with those [Surfactant], (mM)
reported in the literature [32].
Fig. 10. Variation of fluorescence intensity ratios for BODB as
On the other hand, the plots of fluorescence a function of SDS and CTAB concentrations, calculated at 522
intensity against the concentration of SDS give and 552 nm, respectively. The [CTAB] scale is multiplied by 10.
rise to two break points. The first point is at
4.0 mM and the second one is around 6.0 mM.
Multiple CMCs for SDS are already reported 3.4. Estimation of the probe-micelle binding
in a number of current scientific reports [34,35]. constant
The lower CMC is often assigned to a phase
with premicellar aggregates while the higher In order to know how strongly BODB and
one is referred to some changes in the shape and BTDB bind with the surfactants, the binding
size of the micelle. These results clearly suggest constant between each and the different micelles
that the present probes perform a better role as has been determined using the following equation
sensors for monitoring the aggregation of surfac- [35]:
tant molecules than do other probes. For example,
nile red gave only one CMC value of 8.0 mM for ðI m  I o Þ=ðI t  I o Þ ¼ 1 þ ðK½MÞ1 , (1)
SDS [36]. where Io, It and Im are the fluorescence intensities
The previous results indicate that the ICT of the probe in the absence of surfactant, at an
fluorescence of BODB and BTDB can serve as a intermediate concentration [M] and under com-
good probe for the properties of the different plete micellization, respectively. K is the binding
micelles. Furthermore, it can sense the type of constant. The micellar concentration [M] is deter-
charge on the polar head group, thus differen- mined by
tiating between anionic, cationic and neutral
surfactants. ½M ¼ ðS  CMCÞ=N, (2)
ARTICLE IN PRESS

T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440 439

CTAB References
4
[1] P. Plaza, D. Laage, M.M. Martin, V. Alain, M. Blanchard-
TX-100 Dsce, W.H. Thompson, J.T. Hynes, J. Phys. Chem. A 104
3
(Im - Io)/(It - Io)

(2000) 2396.
[2] W. Rettig, Top. Curr. Chem. 169 (1994) 253.
2 [3] M. Blanchard-Dsce, V. Alain, L. Midrier, R. Wortmann,
S. Lebus, C. Glania, P. Kramer, A. Fort, J. Muller, M.
Barzoukas, J. Photochem. Photobiol. A 105 (1997) 115.
1 [4] D. Pines, E. Pines, W. Rettig, J. Phys. Chem. A 107 (2003)
236.
[5] A.K. Singh, S. Kanvah, J. Chem. Soc. Perkin Trans. 2
0 (2001) 395.
0 2 4 6 8 10 12 [6] H. Gruen, H. Gorner, Z. Naturforsch. 38a (1983) 928.
[M]-1 x10-5 [7] N. Dicesare, J.R. Lakowicz, J. Phys. Chem. A 105 (2001)
6834.
Fig. 11. Plot of (Im–Io)/(It–Io) against [M]1 for DODB in [8] W. Rettig, K. Rurack, M. Sczepan, New trends in
CTAB and TX-100. fluorescence spectroscopy, in: B. Valur, J. Brochon
(Eds.), Applications to Chemical and Life Sciences,
Springer, Berlin, 2001, p. 125.
[9] P. Fromberz, K.H. Dambacher, H. Ephardt, A. Lamba-
Table 4 cher, C.O. Muller, R. Neigl, H. Schaden, O. Schenk, Ber.
Binding constants (K,  105 l mol1) and the free energy Bunsenges. Phys. Chem. 95 (1991) 1333.
change (DG, kJ mol1) for dienes–micelle interaction at 300 K [10] T.A. Fayed, Coll. Surf. A 236 (2004) 171.
[11] P. Martin, M.A. Martin, D. Castillo, I. Cayre, Anal. Chim.
Surfactant BODB BTDB Acta 205 (1988) 129.
[12] Y. Singh, A. Gulyani, S. Bhattacharya, FEBS Lett. 541
K DG K DG
(2003) 132.
CTAB 2.770.15 31.18 5.270.33 32.81 [13] E. Miller, B. Wandlet, S. Wysoki, D. Jozwik, A.
TX-100 2.170.076 30.55 1.670.078 29.87 Mielniczak, Biosens. Bioelectron. 20 (2004) 1196.
[14] C. Peinado, E.F. Salvador, F. Catalina, A.E. Lozano,
Polymer 42 (2001) 2815.
[15] I.M. Pepe, J. Photochem. Photobiol. B 48 (1999) 1.
[16] A.K. Singh, G.R. Mahalaxmi, Photochem. Photobiol. 71
where S represents the surfactant concentration (2000) 387.
under experimental conditions and N is the [17] H. El-Gezawy, W. Rettig, R. Lapouyade, Chem. Phys.
aggregation number of the micelles. The N values Lett. 401 (2005) 140.
[18] A.K. Singh, M. Darshi, S. Kanvah, J. Phys. Chem. A 104
were taken as 60 and 143 for CTAB and TX-100,
(2000) 464.
respectively [37]. The plots of (Im–Io)/(It–Io) vs. [19] A.K. Singh, S. Kanvah, New J. Chem. 24 (2000) 639.
[M]1 are linear (r40:96) in correspondence with [20] A.K. Singh, M. Darshi, Biochim. Biophys. Acta 1563
Eq. (1). Representative plots are presented in (2002) 35.
Fig. 11. The K values have been determined from [21] N.D. Cesare, J. Lakowicz, J. Photochem. Photobiol. A 143
the slopes of the plots, and collected in Table 4 (2001) 39.
[22] S.P. Gromov, A.I. Vendernikov, E.N. Ushakov, L.G.
with the free energy changes associated with the
Kuzmina, A.V. Feofanov, V.G. Avakyan, A.V. Churakov,
diene–micelle binding calculated at 300 K. The Y.S. Alaverdyan, E.V. Malysheva, M.V. Alfnov, J.A.K.
corresponding values for SDS could not be Howard, B. Eliasson, U.G. Edlund, Helv. Chim. Acta 85
calculated due to the complicated behaviour in (2002) 60.
this surfactant. The relatively stronger binding [23] S.H. Etaiw, T.A. Fayed, N.Z. Saleh, J. Photochem.
between both dienes and CTAB can be attributed Photobiol. A (2005), in press.
[24] J.N. Demas, G.A. Crosby, J. Phys. Chem. 75 (1971) 991.
to the polar nature of BODB and BTDB (ground-
[25] N.C. Maiti, S. Mazumdar, N. Periasamy, J. Phys. Chem. B
state dipole moment amount to 3.6) [23], which 102 (1998) 1528.
enhances their association with the charged [26] M. Mishra, G.B. Behra, M.M.G. Krishna, N. Periasamy,
surfactant molecules. J. Lumin. 92 (2001) 175.
ARTICLE IN PRESS

440 T.A. Fayed et al. / Journal of Luminescence 121 (2006) 431–440

[27] A.S. Tatikolov, G. Ponterini, Zh.A. Krasnaya, Int. J. [32] G. Saroja, A. Samanta, Chem. Phys. Lett. 246 (1995) 506.
Photoenergy 2 (2000) 17. [33] K. Kalyanasundaram, J.K. Thomas, J. Phys. Chem. 81
[28] S.K. Saha, S. Santra, S.K. Dogra, J. Mol. Struct. 478 (1977) 2176.
(1999) 199. [34] A. Mallick, B. Haldar, S. Maiti, N. Chattopadhyay,
[29] D. Mandal, S.K. Pal, D. Sukul, K. Bhattacharya, J. Phys. J. Colloid Interf. Sci. 278 (2004) 215.
Chem. A 103 (1999) 8156. [35] M. Almgren, F. Grieser, J.K. Thomas, J. Am. Chem. Soc.
[30] Unpublished results, the measurement technique was 101 (1979) 279.
described previously, see; S. Landgraf, Spectrochim. Acta [36] K. Goodling, K. Johnson, L. Lefkowitz, B.W. Williams,
A 57 (2001) 2029. J. Chem. Educ. 71 (1994) A8.
[31] S.M. Dennison, J. Guharay, P.K. Sengupta, Spectrochim. [37] G. Saroja, B. Ramachandan, S. Saha, A. Samanta, J. Phys.
Acta 55A (1999) 903. Chem. B 103 (1999) 2906.

You might also like