Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

15 POLYMER LUMINESCENCE

AND PHOTOPHYSICS
D. PHILLIPS and M. CAREY
Department of Chemistry, Imperial College, London SW7 2AY1 UK

15-1 INTRODUCTION
Ultra-violet and visible light-absorbing chromophores in synthetic polymers
may be present due to adventitious impurities such as oxidation products,
termination residues or initiator fragments (type A), or be present in the repeat
unit and thus be in high concentration (type B). Many simple synthetic polymers
such as poly(ethylene) and poly(propylene) in a pure state will exhibit only o-a*
absorptions in the high-energy UV region, where most organic molecules absorb.
Such excitations in general lead to photochemical reactions rather than lumines-
cence, and excited states will thus be very short-lived. Here we focus attention
arbitrarily on species that absorb in the spectral region from 250 nm to longer
wavelengths, where luminescence may be an additional fate of photoexcited
species, which are depicted in Figure 15.1, for a typical organic chromophore.
The many studies carried out on luminescence in synthetic polymers have been
motivated by a wide range of scientific and technological aims. Some of the more
obvious are categorized below [1,2].
(a) F undamental interests: these include studies on the nature of photoemis-
sion from polymers of type B, in which interchromophoric interactions are of
special interest.
(b) Luminescence of probe molecules: these studies permit the evaluation of
polymer properties. In particular, measurement of the relative intensities of
fluorescence of a probe molecule polarized parallel to and perpendicular to the
plane of linearly polarized exciting radiation as a function of the orientation of
a solid sample yields information concerning the ordering of polymer chains. In
soultion, similar polarization studies yield information on the rotational relax-
ation of chains and the viscosity of the microenvironment of the probe molecule.
The study of luminescence intensity of probe molecules as a function of tempera-
ture has been used as a method of studying transition temperatures and sub-
group motion in polymers.
(c) Luminescent species in polymer photooxidation: the problems associated
with establishing a mechanism for the photooxidation and weathering of syn-
Polymer Spectroscopy. Edited by Allan H. Fawcett
© 19% John Wiley & Sons Ltd
lntersystem

S-S absorption
crossing

Internal
conversion
Vibrational

T-T absorption
Internal relaxation
conversion

lntersystem
crossing
Fluorescence

Phosphorescence
lntersystem
Absorption

crossing
Vibrational Vibrational
relaxation relaxation

Figure 15.1 Jablonskii state diagram depicting the fates of photoexcited polyatomic
molecules

thetic polymers are great, and any method that provides additional information
is useful. In addition to traditional methods such as product analysis, infrared
spectroscopy (both conventional and ATR) and U V-visible absorption spectros-
copy, luminescence methods have been employed.
(d) Identification of polymers: luminescence spectroscopy can provide a con-
venient method for rapid identification of some synthetic polymers.
We will cite here a few classic examples of studies in the various categories,
using steady-state measurements.

15.2 PROBES OF ORDER IN POLYMERS


Physical properties of polymers are often altered significantly by preferential
orientation of structural units by drawing or some other means. The degree of
anisotropy thus introduced requires measurement if correlation between struc-
ture and physical properties is to be established. There are a number of methods
available for the measurement of such anisotropy, including wide-line NMR,
optical birefringence, X-ray scattering, light scattering, Raman spectroscopy,
infrared dichroism and fluorescence polarization. The methods are not all
equivalent in the type of information they provide, but when used simultaneously
on the same sample they can yield complementary data. Thus, for example,
birefringence is sensitive to the orientation of both the amorphous and the
crystalline units, whereas X-ray scattering reveals the orientation of crystallites
only. Raman methods can probe much more local order than X-ray techniques.
In principle it is desirable to have knowledge of the complete distribution
function in a sample, but X-ray diffraction is currently the only method that can
be used for this purpose. However, the majority of physical properties depend
only upon the second moment of the orientation distribution, although mechan-
ical properties such as Young's modulus depend also upon the fourth moment.
The latter information is available from both wide-line NMR and fluorescence
polarization measurements. Experimentally the fluorescence polarization

Figure 15.2 Intensity of parallel component of fluorescence, I, as function of orientation


of sample in uniaxially stretched poly(vinyl alcohol) film at draw ratios of (1) 1, (2) 1.08, (3)
1.3, (4) 1.6, (5) 2.0, and (6) 5.0 (after figure in J. Polym. ScI, Polym. Symp., 1970,31, 353
measurements are simple, in that a rigid polymer sample in which fluorescent
molecules are dispersed or chemically attached is excited in a spectrofluorimeter
with (usually) vertically plane-polarized light. Measurements of the fluorescence
intensity of the probe with the analysing polarizer parallel (J1) and crossed (J1)
with the excitation polarizer are taken as a function of the orientation of the
sample with respect to the vartically polarized excitation radiation; that is, plots
are made of the components J11 and J 1 as a function of physical ratation of
a sample through 360°. The intrinsic probe, which must be a long molecule, is
assumed to align with the fibres in a material.
A typical plot is shown in Figure 15.2, in which the anisotropy introduced by
drawing is clearly illustrated. [3] The method can be used to distinguish
orientations which correspond to different arrays which nevertheless have ident-
ical orientation functions.

15.3 PROBES OF SUB-GROUP MOTIONS


Phosphorescence may be used as a probe of sub-group motion in synthetic
polymers, particularly be studying the temperature dependence of the emission of
an intrinsic probe. Using probe naphthalenes (or ketones), a wide variety of
polymer films have been studied in which, over the temperature range 3OO-77K,
the intensity of phosphorescence from the probe was found to vary by up to four
orders of magnitude [4]. This is illustrated in Figure 15.3; for a series of styrene
polymers with emitting comonomers, the temperature dependence showed dis-
tinct linear regions with at least one common discontinuity (for each polymer
type) in the slope within a narrow temperature region. These discontinuities
coincided well with the known y-transition temperatures and secondary transi-
tion temperatures for each of the polymer types investigated. The conclusion of
this study was that the phosphorescence decrease with increasing temperature
was not due to temperature dependent intramolecular decay or 'intermolecular
deactivation' but could be best explained in terms of the increasing accessibility of
the excited chromophore to molecular oxygen. The observed temperatures of
discontinuity were explained in terms of the several possible structural relax-
ations, and in general the observed temperatures were in good agreement with
results from other relaxation measuring techniques.

15.4 PHOTOCHEMISTRY IN POLYMERS


The vast literature on photopolymerization and cross-linking makes this subject
impossible to attempt within the scope of this brief review. Photochemical effects
of formed polymers are dominated by photooxidation processes summarized in
Figure 153 Plots of In / p (J? = phosphorescence intensity) against inverse temperature
for styrene polymers. Transitions corresponding to Ty and other sub-group motions are
visible (after figure in Macromolecules, W)% 7, 233).

Figure 15.4, which also depicts the means available to protect polymers against
the effect of light. These are the use of UV absorbers, A; quenchers of excited
states, B; radical scavengers, C; singlet oxygen scavengers, D; or destroyers of
hydroperoxide, E.
QUENCHER
RADICAL SCAVENGER
UV-ABSORBER «'•

*•

RO # ROO,
MOLECULAR
DISSOCIATION

ROOH METAL
QUENCHER DEACTIVATOR
PEROXIDE
DECOMPOSER

Figure 15.4 Mechanism of photo-oxidation and stabilisation of commercial polymers

15.5 EXCIMER-FORMING POLYMERS [5]


In type B polymers, the structural constraints of the polymer chain tend to
confine the chromophores in spatial positions such that they can be expected to
exhibit strong mutual interactions. These may depend strongly upon the relative
orientation of the interacting chromophores, and the orientations themselves will
usually be dependent upon the conformation of the polymer chain. Interaction
between the excited state chromophore and a neighbouring ground state can give
rise to excimer (excited dimer) formation, which proves to be a powerful
diagnostic of interacting molecules.
The salient features of excimer formation are represented in Figure 15.5.
Aromatic molecules at large separations, that is at separations much greater than
4 A, may be considered as isolated entities. Consequently, if the aromatic
molecules are in an excited state, the fluorescence is unaffected by the presence of
other molecules. For small separations, less than 4 A, repulsive potentials R(r)
and R'(r) will exist between molecules in their ground state and between mol-
ecules in the ground and the excited state. In general, the existence of these
repulsive potentials prevents the formation of complexes. However, for the
interaction between ground and excited state molecules, an attractive potential
V(r) may be obtained, owing to configurational interaction between resonance
and exciton-resonance states. The combination of repulsive and attractive
potentials may form the excimer state shown by the potential well in Figure 15.5.
The fluorescence from the 'excimer' state will thus be unstructured (since a corre-
energy
fluorescence intensity

Figure 15.5 Energy diagram for excimer formation

sponding ground state complex does not exist) and at a lower energy than the
corresponding monomer emission. In general, excimer formation can occur
whenever aromatic chromophores adopt a face-to-face coplanar arrangement
with a separation of 0.3-0.35 nm, as shown for naphthalene in Figure 15.6.
Static measurements of intensities of monomeric fluorescence (here defined as
that from an uncomplexed chromophore attached to the polymer chain) relative
to that from the excimer can be used to yield information relating to energy
transfer and migration, rotational relaxation and segmental motion, and to the
heterogeneity of synthetic polymers and copolymers in solution and solid forms.
Results of technological importance are available. Thus, in blends of polymers,
such measurements have been used to investigate compatibility [6, 7].
Figure 15.6 Excimer formation in a naphthalene-containing
molecule

15.6 DYNAMICS OF LUMINESCENCE


The processes of depletion of excited-state population in Figure 15.1 lead fo
fluorescence decay times which may be 10 " 8 -10 " *2 s or less. Molecules may thus
provide a 'clock' over this range with which to time other processes which are

Spontaneous radiative transitions

lntersystem crossing (S 1 -T 1 )

Internal conversion (Sn - SnJ


Vibrational redistribution
and isomerization
Field-induced transitions

Coherent exciton
Exchange transfer
Resonance (Forster) transfer

Diff usional encounter (1 cp)


Rotational diffusion (1 cp)
Vibrational relaxation
Geminate recombination
Chemical reaction

Typical Q-switched laser, Typical picosecond Shortest laser pulse


excimer, N 2 pulse duration laser pulse duration yet produced

Limit of photon-counting
streak camera detection

Figure 15.7 Some physical and chemical processes which occur on the 10~ 6 -10~ l5 s
time scale
Second
harmonic generator
Second
harmonic generator
Harmonic Harmonic
separator separator
Cavity
KTP dumper KDP Sample

rfout
sync Cavity dumper rfout Fast photodipde
Mode-locker
driver out driver
Constant fraction Filter
sync out Timing filter timing discriminator
amplifier
MicroChannel
CFTD plate
CFTD
PC/AT
computer TAC/SCA
X100 Amplifier
Time-to-amplitude
converter/single
Multichannel channel analyser
analyser

Figure 15.8 Time-correlated single-photon counting spectrometer based on CW mode-


locked Nd: Y AG laser

subject to environmental influence, such as diffusion, energy transfer and migra-


tion, etc., as shown in Figure 15.7 [8,9].
For fluorescence measurements, by far the most versatile and widely used
time-resolved emission technique involves time-correlated single-photon count-
ing [8] in conjunction with mode-locked lasers, a typical modern apparatus
being shown in Figure 15.8. The instrument response time of such an apparatus
with microchannel plate detectors is of the order of 70 ps, giving an ultimate
capability of measurement of decay times in the region of « 7 ps. However, it is
the phenomenal sensitivity and accuracy which are the main attractive features of
the technique, which is widely used for time-resolved fluorescence decay, time-
resolved emission spectra, and time-resolved anisotropy measurements. Below
are described three applications of such time-resolved measurements on synthetic
polymers, derived from recent work by the author's group.

15.7 FLUORESCENCE DECAY IN VINYL AROMATIC


POLYMERS
Fluorescence in such polymers is dominated by excimer formation, the simplest
kinetics for which were described by Birks and co-workers [10,11] (Scheme 1). In
Scheme 1 Birks kinetic scheme

this treatment the influences of diffusion or energy migration are neglected, and
only the two chromophores directly involved in the excimer formation process
are considered. In Scheme 1, M refers to the ground state monomer species, 1 M*
to the monomer in its first excited singlet state and 1 D* represents the excimer; kM
is the molecular decay rate, which is the rate of depopulation of * M* by radiative
or non-radiative decay in the absence of other chromophores or intra-molecular
chemistry; kD is the rate of radiative and non-radiative decay of the excimer;fcDMis
the rate of formation of excimer from monomer, and kMD is the rate of dissociation
of the excimer to recreate the excited monomer.
Equations for the monomer and excimer population are then as follows:

[ 1 M*] = -^ ^ [ ( A 2 - J 0 e x p ( - A , t ) + (X-A 1 )exp(-A 2 t)] (D

[»D*] = kD^ll.^*\cxp(- A l t )exp(- A2t)] (2)


[A2 ~ ^V
where X9 A1 and A2 are functions of the rate parametersfcM,fcD,fcMDand /cDM, viz:
X = kM + /cDM[M] (3)
1 1/2
A1 = 1/2(Z + kD + kM - {(kD + fcM - X) + 4fcMD/cDM[M]} ) (4)
2 1/2
A2 = 1/2(Z + kD 4- /cM + {(kD + fcM - X) +4fcMD/cDM[M]} ) (5)
It can be seen from Equations (1) and (2) that the monomer and excimer decays
are both the sum of exactly two exponential decay terms, with the same lifetimes
A1 and A2 appearing in both monomer and excimer decays. In addition, the two
pre-exponential factors in the excimer decay are of equal magnitude but opposite
sign. However, except to a first approximation, neither of these characteristics is
usually seen experimentally in polymers [12-14] where, typically, the monomer
and excimer decays will give different values OfA1 and A2 and the excimer decay
will not have pre-exponential factors of equal magnitude. As real polymer decays
do not follow the Birks kinetic scheme, the scheme evidently does not take
account of all the photophysical processes which occur in polymers, and efforts to
improve the models have been made in two main directions. The first approach
has been to parameterize the deviations from Birks kinetics using a third
exponential decay term in both the monomer and excimer decays. The third term
can then be interpreted in a number of ways, such as the existence of a third
species. In fact, if all the photophysical processes occurring in the polymer are
time independent, i.e. can be expressed by a simple rate constant in a kinetic
scheme, then the existence of a third species is the only conclusion that can be
drawn from a third decay term. Some of the models proposed, and which have
been successful in explaining polymer fluorescence decays, are as follows:
(1) two monomer species, the first one able to form the second, but only the
second one able to form excimers [15,16];
(2) two monomer species, one of which can be formed only by dissociation of
excimers, and cannot reform excimers [15,16];
(3) three excimer species, two of which are in equilibrium, the third being
formed only from the monomer [17-19].
The multi-exponential approach has been criticized on the grounds that the
kinetic schemes are not unique: data which are consistent with two excimers may
also be consistent with a second type of monomer [20]. Also, there is rarely
supporting spectroscopic evidence for the presence of a third species, which
would be expected to have a different emission spectrum. However, except in the
case of poly(vinylcarbazoles) [21-24], no such evidence has been found.

15.7.1 D I F F U S I O N A L MODELS

The second approach to the study of excimer kinetics has been more theoretical.
In experiments on dilute solutions of unlinked chromophores, there has been
some success in considering the process of excimer formation as a diffusive
process [25]. Nemzek and Ware [26] used an extension of the Smoluchowski
equation [27] devised by Collins and Kimball [28], which gives for Jt(^DM

R
k(t) = 4TIDABR'N( 1+ ) (6)

The Birks kinetic scheme can then be adjusted to include k(t)DM. Because of the
complexity involved, the rate constant /cMD is usually neglected at this stage. The
population of the monomer excited state then has the time dependence of
Equation (7):

[ 1 M*] = [ 1 M^] 0 exp[ -(fcM + 4nDABR'Nt)- *' t1'2] (7)

Consider now the diffusion of excitation through an array of monomers. The


excited state moves along the polymer chain from one monomer to another,
probably by the Forster dipole-dipole mechanism, but other energy transfer
mechanisms such as the Dexter electron exchange mechanism may also play
a part, especially in solid polymers, where the chromophores are very close
together. The excitation may be trapped at any chromophore by formation of an
excimer. A number of different models have been used to consider the time
evolution of an excited state which may migrate to a trap site, and often several
different approaches are used to approximate the observable parameters for each
model.
A review of these complex mathematical models is beyong the scope of this
paper, but they can be summarized as follows.

15.7.1.1 Random Walk Migration, Evenly Spaced Chromophores


This model has been investigated ]by a number of groups and solved approxi-
mately using several different methods. Huber [29] solved the rate equations for
the donor (monomer) decay using the t-matrix approximation, resulting in
Equation (8) for the one dimensional case.
[M*] = A exp(4n2qWt)erfc(2nqW1/2tl/2) (8)
In the asymptotic limit, i.e. for long times, the decay can be approximated by
Equation (9); however, when the number of trap sites is sufficiently small, the
decay reduces to an exp — (at + btlf2) dependence similar to Equation (7)

[M
*]=(WW* (9)

In addition to the Mnatrix approximation, however, a various methods have


been used to solve the deep trapping problem which has been solved exactly in the
one dimensional case [30]. Movaghar et al. compared the coherent potential
approximation (CPA) [31] and the first passage time approach (FPT) [32]
results with the exact solution while stating that the Mnatrix approximation used
by Huber is less accurate than the CPA under all conditions. Movaghar et al.
[30] [31] found that the FPT approach is superior to the CPA at all trap
concentrations except for very high concentrations approaching 1, where all
chromophores are traps. At long times the FPT approach gives a solution which
asymptotes to exp — (ati/2)9 similar to the low trap concentration Mnatrix
derived result. By contrast, the exact result asymptotes to exp — (at1/3).

15.7.1.2 Random Walk, Random Distribution Chromophores


A second, more complex model which can approximate energy migration kinetics
involves the relaxation of the condition that there must be an even distribution
of chromophores. Such a relaxation can involve, say, a random distribution of
chromophores in three dimensions interspersed by a random distribution of
traps. The GAF [34] and LAF [35] models are of this type and, in addition,
a model has been derived for polymers (FAF) [36] which relates fluorescence
decay parameters to the radius of gyration of the polymer.

15.7.1.3 Multiple Trap Energies


A further complication is to consider the disorder of the energies of the monomer
excited states as well as positional disorder. In a polymer, the chromophores
are in a range of environments, each of which will have different energies. This
problem has been treated theoretically [37] and in a Monte-Carlo simulation
[38], both giving an approximate relationship of the form of Equation (10):
mv* = b + cf-1 (10)
More recently, the problem of energetic disorder has been considered by
Stein et al. [39], who treated the combination of energy migration and trapping
as part way between donor-donor transfer and direct trapping of the excitation.
The theory agreed well with some of their experimental polymer anisotropy
decays.

15.7.1.4 Reversible Excimer Formation


In the Birks kinetic scheme, back-transfer is considered simply by the rate
constant fcMD. Weixelbaumer et al. [40] used an approximate method to ap-
proach this, whereas Sienicki and Winnick [41] derived an exact result, and
posed the question, what happens if monomers formed by back dissociation
behave differently from those excited directly? The question was answered by
Berberan-Santos and Martinho [42], who showed that k(t)DM does not necessar-
ily decrease monotonically but can sometimes increase with time.

15.7.1.5 Diffusion of Energy and Chromophore


Baumann and Fayer [43] considered a two-body problem in which diffusion and
energy transfer occurred simultaneously. Frederickson and Frank developed
a simpler one dimensional array model [44].
The equation for the rate of monomer fluorescence in the FF model is given by
Equation (11):
W ) = 4 F M M 1 " q)2 expl(4n2q2W-kM - *rot)fj crfc(2nqW1^2) (11)
In Equation (11), iM(r) is the intensity for fluorescence from the monomer,
which is related to the monomer concentration by the monomer quantum yield of
fluorescence q^ arid the rate of decay of the monomer fluorescence in isolation
fcM; q is the pre-formed trap fraction, which is the fraction of dyads which are trap
sites at equilibrium; W is the rate of energy transfer between nearest neighbours
on the polymer, which is of course highly dependent on the distance between the
chromophores.
Tao and Frank [45] found that 2-vinylnaphthalene homopolymer fluores-
cence decays fit the FF model under conditions of relatively low temperature.
However, they noted that at higher temperatures the model breaks down,
probably because of the breakdown of one of the assumptions below:
(1) that the polymer may be considered as a one-dimensional string of equally
spaced chromophores;
(2) that the primary excimer forming step is energy migration, and not internal
rotation. This requires that there are a number of 'pre-formed trap sites' in the
ground state, which just means that there must be a number of sites where
chromophores are in high-energy configurations which are very close to the
excimer configuration, or else there must be a low-energy conformation very
close to the excimer conformation;
(3) that the number of these 'pre-formed trap sites' is low. For this concentra-
tion of trap sites, the r-matrix approximation becomes poor;
(4) that the excimer formation step is irreversible.
We have extended the FF model to high trap concentrations using the FPT
approximation. In this, the expression for the monomer fluorescence intensity is
given by Equation (12), and that for excimer fluorescence by Equation (13):

'M(0 = <ZFMM1 -<?) 2 exp(— -q ['2WU0(2W T)+ I1VWT)] exp(-2WT) <1T)


T
\ Jo /
(12)
1 2
*E(0 = <?FMkEexp(-fcEt) - ^ F E M - <?)
x e x p U ^ - g I 2Wexp(-2WT)[I0(2WT)]+ I1(HVT)IdT)

-qFEkE(kM-kE)(l-q)2[texp(-u(kE-T-1)- —
T
-q[t "2W
Jo V Jo

x exp(-2WT)lI0(2WT) +I^WTftdTjdu (13)

We have tested some of the above models using data from careful time-
resolved fluorescence measurements on 1-vinylnaphthalene homopolymer, and
copolymers with methyl methacrylate, in the following way. The FF model
appears to have five variables, the amplitude, the isolated decay ratefcM,the rate
of rotation fcrot, the rate of intramolecular energy transfer W, and the number of
trap sites q. However, some of the variables cannot be treated independently and
the FF function may actually by rewritten using only three variables. This is done
by substituting, say, r = l/(feM + krot) and Q = q W1/2 into Equation (11) to give
Equation (14), and fitting the data by varying only the amplitude, t and W. In fact,
if an attempt is made to fit the function while varying all of fcM, krov q and
W simultaneously, all solutions with the same t and Q will fit the data equally
well. So the FF model actually has only three variable parameters, which is one
fewer than the sum of two exponential decay terms.
iM(0 = A exp[(4jT2<22 - t)r] erfc(27rQt1/2) (14)
The efficacy of the FF model was investigated over the range of naphthalene
mole fractions. At 290 K, fluorescence from the 25% 1-vinylnaphthalene polymer
fits the FF model, whereas neither the 50% 1-vinylnaphthalene polymer nor the
homopolymer does so. Obviously the model fits only for low naphthalene
concentrations and low temperatures. The breakdown of the FF model at high
temperatures and high naphthalene concentrations could be explained by the
breakdown of any one of the assumptions outlined above.
Tao and Frank also found that the FF model does not adequately fit
2-vinylnaphthalene fluorescence decay profiles at high temperatures [45]. The
FPT model should be appropriate for high trap concentrations, but in Figure
15.9 the FPT model produces very similar results to the FF model and was unable
to fit any of our data which did not fit the FF model.
In the interpretation of fluorescence data, models as complex as the FF model
are seldom employed. Commercially available programs for fitting time-resolved
fluorescence data generally cover exponential decay, the exponential of a t*
function, or sums of these functions, but rarely anything more complex. It would
be useful to know when simpler approximations, for which fitting routines are

homopolymer

copolymer

Temperature /K
Figure 15.9 Comparison of fitting parameters from the FF model and the FPT model
(see text)
available, are adequate to fit data actually obeying a more complex theory, so
that information about a complex model can be inferred from the fit of the
experimental data to a simple function. It would consequently be useful to know
when the FF model can be successfully approximated by a simpler function.
If q2 W stays within certain limits, then fcDM in the FF model can be accurately
approximated by a constant term plus a term dependent on t1/2. On integration of
the rate equations, the fluorescence decay will then follow Equation (15), which is
commonly available in fluorescence decay fitting software.

[ -(kM + krJt—^-J
AQ /Wt~~\
(15)
Table 15.1 shows reduced x2 test, fcM + ferot and qW112 values obtained from fits
of some of the experimental data presented earlier to the FF model and to
Equation (15). The last two columns of Table 15.1 consists of values of
4(1 — 2/n)q2W and feM + krov The chi-squared values are equally good for both
functions, but the kM + krot and qW1/2 parameters do show some deviations
which may be not be explained by experimental error. The FF model consistently
finds a slightly less 'exponential decay', indicating that small inaccuracies in the
approximation have shown up.
Tao and Frank [45] presented data consistent with the FF model without any
reference to fitting the exp — (at + bt1/2) approximation. We tested this by
simulation; thus, Tao and Frank's data were simulated with the same amplitude
as shown in their paper, from their published parameters, and with Gaussian
noise added. When our simulated curves were analysed with our FF fitting
program, they gave chi-squared values of 1.00 ± 0.05. They were subsequently
fitted to Equation (15). The x2 values from the FF fit were then subtracted from
the x2 values from the t1'2 fit to give a measure of the difference in the quality of
the fit. These results are presented in Table 15.2, along with parameters extracted
from the paper.
At low temperatures, Equation (15) is well satisfied, and #2(*1/2) —X2(FT) is
also very low. As the temperature rises, however, qW1/2 increases faster than

Table 15.1 Quality of fit and some fitting parameters for 27% l-vinylnaphthalene/72%
methyl methacrylate copolymer
Tw Tw kM +1/2kTOt 4(1-2/7T)- /cM + /crot
Temp (FF)/ (r 1/2 )/ (r )/ q2W/ (FF)/
(K) *2(FF) 2 112
x (t ) 10 7 S" 1 10 7 S- 1 (xlO" 4 ) 10 7 S" 1 10 7 S" 1
290 U5 1.11 O20 019 115 O30 236
270 1.11 1.05 0.12 0.12 116 0.18 130
250 1.09 1.09 0.049 0.047 114 0.07 118
230 0.99 1.06 0.041 0.038 1.95 0.06 1.99
210 1.26 1.30 0.053 0.046 1.71 0.08 1.76
Table 15.2 Fitting parameters for actual and synthesized data of Tao and Frank;
2-vinylnaphthalene homopolymer
Temperature/K * 2 (FF) * 2 (' 1/2 )-X 2 (FF) 4(1-2/Tr)^2WyIO7S-1 ikM +Jk n ^lO 7 S" 1
293 L29 017 15 11
273 1.18 0.11 1.9 3.1
253 1.10 0.05 1.0 2.8
233 1.10 0.03 0.62 2.4
213 1.08 0.02 0.44 1.9
193 1.05 0.01 0.24 1.6
173 1.06 <0.01 0.09 1.5
153 1.05 <0.01 0.08 1.5
133 1.02 <0.01 0.02 1.4
113 1.03 <0.01 0.02 1.4

fcM + kTOV until the condition that 4(1 — 2/^)9 W < WkM + fcrot is no longer satis-
fied above 193 K. At the same time, x2(f1/2) - X2(ff) increases until, at « 293 K,
the two models should easily be differentiated. However, by 293 K, the experi-
mental x2(FF) value has also increased to a stage where the data no longer fit the
FF model. So at 293 K the exp - (at + bt1/2) function may possibly fit the data
better than the FF model. In fact, nothing so far has contradicted the premise that
Tao and Frank's data can fit Equation (15) as well as the FF model. This means
that the polymer could actually be undergoing any set of processes which approxi-
mates sufficiently well to an exp — (at + bt1/2) function.

15.7.1.6 Fluorescence Anisotropy Measurements


The fluorescence decay times of excited states are such that the fluorescence
depolarization technique may only be used to examine relatively high frequency
relaxation processes of polymers. Consequently fluorescence depolarization has
been primarily limited to the study of relaxation processes of polymers in
solution. The anisotropy of a system, r(t\ is derived from measurements of the
fluorescence decays with polarizations parallel and perpendicular to the polariz-
ation of excitation:
Ht) = [Z1W - IAt)WiIt) + 2Z1(O] = D(t)/S(t)
Time-resolved fluorescence anisotropy measurements [47] can provide de-
tailed information on the reorientation dynamics of molecules in solution. Until
recently, however, this information has been limited to single rotational correla-
tion times, which are only strictly appropriate for the diffusion of spherically
symmetric molecules. Improvements in instrumentation and data analysis tech-
niques during the last decade have led to increasingly accurate measurements of
fluorescence lifetimes, with parallel improvements in determinations of fluores-
cence anisotropies.
The advances in time-resolved techniques have fostered a reexamination of
theories of the rotational motions of molecules in liquids. Models considered
include the anisotropic motion of unsymmetrical fluorophores; the internal
motions of probes relative to the overall movement with respect to their
surroundings, the restricted motion of molecules within membranes (e.g., wobbl-
ing within a cone), and the segmental motion of synthetic macromolecules [8].
Analyses of these models point to experimental situations in which the anisot-
ropy can show both multi-exponential and none-exponential decay. Current
experimental techniques are capable in principle of distinguishing between these
different models. It should be emphasized, however, that to extract a single
average rotational correlation time demands the same precision of data and
analysis as fluorescence decay experiments which exhibit dual exponential
decays. Multiple or non-exponential anisotropy experiments are thus near the
limits of present capabilities, and generally demand favourable combinations of
fluorescence and rotational diffusion times [48].
An example is cited below of study on the copolymers (a) methyl methacrylate/
acenaphthylene (PMMA/ACE), (b) methyl methacrylate/1-vinylnaphthalene
(PMMA/1-VN), (c) methyl acrylate/acenaphthylene (PM A/ACE), and (d) methyl
aery late/1-vinylnaphthalene (PMA/l-VN). The results are summarized in
Table 15.3.
Averaging all the determinations for the initial anisotropy for each polymer
sample leads to the following values for excitation at 300 nm: PMMA/ACE,

Table 153 Fluorescence anisotropy parameters for labelled acrylic polymers


T/K Tf/ns To TR
298 ±2 17.4 ±0.2 0.10 ±0.01 0.8 ±0.3
PMA/ACE 260±2 17.4±0.3 0.10±0.01 1.3±0.2
245±2 17.4±0.3 0.11 ±0.01 1.8±0.3
230±2 17.5±0.3 0.12±0.02 2.5±0.3
289±2 15.1 ±0.1 0.13±0.01 0.5±0.1
PMA/VN 275±2 14.9±0.1 0.13±0.01 0.8±0.1
260±2 14.8±0.1 0.14±0.01 1.0±0.2
245±2 14.9±0.1 0.14±0.01 1.3±0.3
230±2 14.9±0.1 0.15±0.01 1.7±0.3
298±2 15.5±O.l 0.13±0.01 1.3±0.1
PMMA/ACE 275±2 15.7±O.l 0.13±0.01 2.2±0.2
260±2 15.4±0.2 0.13 ±0.01 3.2 ±0.5
245±2 15.5±0.2 0.13±0.01 4.5±0.7
23O±2 15.6±0.1 0.11 ±0.02 5.6±0.7
298±2 15.9±0.02 0.15±0.01 1.3±0.2
PMMA/VN 275±2 15.6±0.2 0.16±0.01 2.2±0.5
260±2 15.5±O.l 0.14±0.01 2.7±0.3
245±2 15.4±0.1 0.15±0.01 3.6±0.5
23O±2 15.4±0.1 0.16±0.01 4.9±0.7
Direction of
independent motion
Polymer
backbone backbone

Figure 15.10 Alignment of the transition dipoles and the direction of the independent
motion of the 1-vinylnaphthalene chromophore relative to a polymer backbone

r0 = 0.13 ±0.01; PMA/ACE, r0 = 0.11 ±0.01; PMMA/1-VN, r0 = 0.15 ±0.01;


PMA/l-VN, r0 = 0.14 ±0.02. These results are in excellent agreement with
values obtained for polymers with similar compositions. Initial anisotropies are
expected to have the value of 0.4. However, the first and second excited states of
naphthalene and its derivatives are, in the Platt notation, designated 1L1, and 1 L 3
respectively. The transition dipole moments for absorption into these bands are
directed along the long (1L1,) and short ( 1 LJ axes of the aromatic rings. Irradi-
ation at 300 nm produces excitation of both absorption bands, and so naphtha-
lene, when excited at this wavelength, can be considered to have a planar rather
than a linear absorption oscillator.
The 1-vinylnaphthalene chromophore, unlike the acenaphthylene chromo-
phore, would appear to be capable of motion independent of the polymer back-
bone by rotation about the single bond [Fig. 15.10]. However, such rotation
cannot lead to depolarization. Consequently for the 1-vinylnaphthalene labelled
polymers, as with the acenaphthylene labelled polymers, it is only segmental
motions which lead to depolarization.
For the poly(methacrylates) and poly(acrylates), the a and /? relaxations are
associated with segmental motions of the polymer and independent motions of
the ester substituents respectively. The merging of these transitions at high
frequencies or temperatures corresponds, at the molecular level, to the incidence
of co-operative motion between the substituent and the polymer backbone.
Consequently, it is to be expected that, in solution, the high frequency motions of
both polymer chain and fluorescent label will assume a co-operative form
characterized by a single relaxation process/time.
The activation energies derived from the results at different temperatures
(Table 15.3) show that in poly(methyl methacrylate) and poly(methyl acrylate)
the segmental motions are largely controlled by solvent flow.

15.8 CONCLUSION
Time-resolved luminescence measurements have still unrealized potential for the
study of energy migration, rotational motion and surface effects in polymers in
solution and in the solid state.
15.9 ACKNOWLEDGEMENTS
This paper has drawn upon the work of AJ. Roberts, G. Rumbles, R.C. Drake,
C.F.C. Porter and R.L. Christensen, of Imperial College, London, and of
Professor Ian Soutar and his group at the University of Lancaster. All are
thanked for their contributions. Financial support from SERC and The Royal
Society is gratefully acknowledged.

15.10 REFERENCES
[1] D. Philips (Ed.), Polymer Photophysics: Luminescence, Energy Migration and Mol-
ecular Motion in Synthetic Polymers, Chapman Hall, London, 1985.
[2] S.W. Beavan, J.S. Hargreaves and D. Phillips, Adv. Photochem. 1979,11, 207.
[3] Y. Nishijima, J. Polym. ScL Polym. Symp., 1970, 31, 353.
[4] A.C. Somersall, E. Dan and J.E. Guillet, Macromolecules, 1974,7, 233.
[5] D. Phillips, Br. Polym. J., 1987,19,135.
[6] W.C. Tao and CW. Frank, in J.-P. Fouassier and J.F. Rabek (Eds.), Lasers in
Polymer Science and Technology: Applications, Vol. 1, CRC Press, Boca Raton, 1990
p. 161.
[7] M.A. Winnick, in J.-P. Fouassier and J.F. Rabek (Eds.), Lasers in Polymer Science
and Technology: Applications, Vol. 1, CRC Press, Boca Raton, 1990, p. 197.
[8] G. Rumbles and D. Phillips, in J.-P. Fouassier and J.F. Rabek (Eds.), Lasers in
Polymer Science and Technology: Applications, Vol. 1, CRC Press, Boca Raton, 199
p. 91.
[9] D. Phillips, in CE. Hoyle and J.M. Torkelsen (Eds.), Photophysics of Polymers, ACS
Symp. Ser., 1987, (358), 308.
[10] J.B. Birks, Photophysics of Aromatic Molecules, Wiley-Interscience, London, 1970,
pp. 322-335.
[11] J.B. Birks, DJ. Dyson and T.A. King, Proc. R. Soc. London, Ser. A, 1964, 277, 270.
[12] AJ. Roberts, D.V. O'Connor and D. Phillips, Ann. N.Y. Acad. ScL, 1981,366,109.
[13] D. Phillips, AJ. Roberts and I. Soutar, Polymer, 1981, 22, 293.
[14] D. Phillips, AJ. Roberts and I. Soutar, J. Polym. ScL, Polym. Phys. Ed., 1982,20,411.
[15] D. Phillips, AJ. Roberts and I. Soutar, Polymer, 1981, 22,427.
[16] D. Phillips, AJ. Roberts and I. Soutar, J. Polym. ScL, Polym. Phys. Ed., 1980, 18,
2401.
[17] D.A. Holden, P.Y.K. Wang and J.E. Guillet, Macromolecules, 1981,14,405.
[18] F.C. DeSchryver, K. Demayer, M. Van der Anweraer and E. Quanten, Ann. N.Y.
Acad. ScL, 1981, 109.
[19] AJ. Roberts, D. Phillips, F. Aboul-Rasoul and A. Ledwith, J. Chem. Soc, Faraday
Trans. 7,1981,77,2725.
[20] K. Sienicki and G. Durocher, Macromolecules, 1991, 24,1102.
[21] C. David, M. Piens and G. Geuskens, Eur. Polym. J., 1972,8,1019.
[22] C David, M. Piens and G. Geuskens, Eur. Polym. J., 1972,8, 1291.
[23] CE. Hoyle, T.L. Nemzek, A. Mar and J.E. Guillet, Macromolecules, 1978,11, 429.
[24] G.E. Johnson, J. Chem. Phys., 1975,62,4697.
[25] J.C. Andre, F. Baros and M.A. Winnick, J. Phys. Chem., 1990, 94, 2942.
[26] T.C. Nemzek and W.R. Ware, J. Chem. Phys., 1975, 62,477.
[27] M.V. Smoluchowski, Z. Phys. Chem., 1917,92, 129.
[28] F.C. Collins and GE. Kimball, J. Colloid ScL, 1949,4,425.
[29] D.L. Huber, Phys. Rev. B, 1979, 20, 2307.
[30] B. Movaghar, G.W. Sauer and D. Wurtz, J. Slat. Phys., 1982, 27, 473.
[31] B. Movaghar, J. Phys. C (Solid State), 1980,13,4915.
[32] E.W. Montroll, J. Math. Phys., 1969,10, 753.
[33] J. Klafter and R. Silbey, J. Chem. Phys., 1981,74, 3510.
[34] CR. Gochanour, H.C. Andersen and M.D. Fayer, J. Chem. Phys., 1970,70,4254.
[35] R.F. Loring, H.C. Andersen and M.D. Fayer, J. Chem. Phys., 1982,76, 2015.
[36] G.H. Fredrickson and H.C. Andersen, Macromolecules, 1984,17, 54.
[37] M. Griinewald, B. Pohlmann, B. Movaghar and D. Wurtz, Philos. Mag. B, 1984,49,
341.
[38] R. Richert, B. Ries and H. Bassler, Philos. Mag. B, 1984, 49, L25.
[39] A.D. Stein, K.A. Petersen and M.D. Fayer, J. Chem. Phys., 1990,92, 5622.
[40] W. Weixelbaumer, J. Burbaumer and H.F. Kaufmann, J. Chem. Phys., 1985,83,1980.
[41] K. Sienicki and M.A. Winnick, J. Chem. Phys., 1987,87, 2766.
[42] M.N. Berberan-Santos and J.M.G. Martinho, J. Chem. Phys., 1991,95,1817.
[43] J. Baumann and M.D. Fayer, J. Chem. Phys., 1986,85,4087.
[44] G.H. Frederickson and CW. Frank, Macromolecules, 1983,16, 572.
[45] W.C Tao and CW. Frank, J. Phys. Chem., 1989,93, 776.
[46] R. Gelles and CW. Frank, Macromolecules, 1982,15, 747.
[47] D. V. O'Connor and D. Phillips, Time-Correlated Single Photon Counting, Academic
Press, London, 1984.
[48] R.L. Christensen, R.C Drake and D. Phillips, J. Phys. Chem., 1986,90, 5960.
[49] R.C Drake, Ph.D. Thesis, University of London, 1986.

You might also like