This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
, where P(t) is the dipole moment of a whole chain at time t. Jernigan deduced the dielectric properties of oc,co dibromide chains and found that
was given by a weighted sum of exponential decay terms plus a constant value. The latter quantity has a value that is dependent on the choice of reference coordinates, and this is a basic problem with this theory. Jernigan [42] multiplied
by a correlation function changed when the reference coordinates were changed, demonstrating the inadequacy of this approach. The origin of the problem is clear: Equation (11) is a scalar equation but measured properties such as dielectric permittivity, Kerr constant, nuclear magnetization and fluorescence emission are related to time-correlation functions for the motions of molecular axes—which are vectorial or tensorial quantities. Although the Jernigan approach is inapplicable to the motions of chains in the bulk amorphous state or in solution, it would be applicable where the reference coordinates are well-defined, e.g. for a chain tethered to a surface.
11.3 CRYSTALLINEPOLYMERS Numerous account of the dielectric properties of partially crystalline polymers are available [3,12,14,17,44,45]. Two classes of partially crystalline polymers are important, those of high crystallinity, such as polyethylene, i-polypropylene and polyoxymethylene, and those having only a medium degree of crystallinity, such as the nylons and polyethylene terephthalate (up to « 50% crystallinity). Multiple relaxations are observed, e.g. lightly oxidized and lightly chlorinated polyethylenes have, in descending order of temperature, ac, /?a and yc relaxations. These have been documented by Ashcraft and Boyd [46] and others [3,4,5]. The <xc process in polyethylenes was first explained by Frohlich [47] using a chaintwist-assisted rotational model in an alkane crystal. Subsequently Hoffman et al. [44] and Williams et al. [48] extended the theoretical model and applied it successfully to polyethylenes and alkanes of different chain lengths. Further development of the chain-twist-assisted rotation and model was made by Mans-
field and Boyd [49], who carried out a computer simulation for a realistic model of a chain moving in the crystal. In all cases it is predicted that the average relaxation time for the ac process increases linearly with chain length for short chains, and that a plateau level is reached for long chains when chaintwisting becomes an essential part of the chain rotation mechanism. The /?a absorption in polyethylenes is generally accepted as being due to large-scale motions in the disordered phase [44,45,46], while the yc process is thought to be due to local motions in the amorphous phase [46] or to local motions in both amorphous and crystalline phases [44]. While many dielectric studies have been made of oxidized and chlorinated polyethylenes, we note that pure polyethylene would not give any dipole relaxation owing the low polarity of the methylene group. For polymers of medium degree of crystallinity, again motions in both amorphous and crystalline phases are observed [3,12,14,17, 23]. Polyethylene terephthalate (PET) is an interesting case, as samples can be obtained in the amorphous state by quenching from the melt or in the partially crystalline state by melt crystallization or quench annealing. Since partially crystalline samples are entirely composed of spherulites, it follows that the amorphous regions (up to 50% of polymer) are contained within the spherulites. Thus this is an 'abnormal amorphous phase', whose relaxation behaviour would be expected to be qualitatively different from that for a wholly amorphous polymer. This has been demonstrated to be the case from the dielectric measurements of Ishida (see [3,45] and subsequently of Tidy and Williams (see data reported in [12]). The amorphous PET exhibits a well-defined a dielectric relaxation, but the partially crystalline sample exhibits a broad a' relaxation whose frequency location is removed to lower frequencies when compared with that for the oc relaxation. Tidy and Williams followed the evolution of the a and a' relaxations in time as an amorphous material was annealed, leading to crystallization, above its T%. As the normal a process disappeared the a' process emerged, and grew as crystallization proceeded. This demonstrates that the 'amorphous' regions within the spherulites suffer a range of constraints imposed by the crystalline regions, giving a slower, broader a process than that for a normal amorphous phase. Thus real time dielectric studies are able to give information on the dynamics of the amorphous regions within spherulites that cannot be readily obtained through NMR and dynamic mechanical relaxation studies. For accounts of DRS studies of other crystalline polymers, including polyoxymethylene, polyvinylidene difluoride, polyvinyl fluoride and the nylons, the reader is referred to the text by McCrum et al. [3], the reviews [12-14,17,23,44, 45] and references therein. In all cases multiple dielectric relaxations are observed, arising from motions within crystals, on crystal surfaces and in the constrained amorphous regions within crystals. These processes are also observed in NMR and mechanical relaxation studies of such polymers.
11.4 LIQUID CRYSTALLINE (LC) POLYMERS Liquid-crystal-forming (mesogenic) groups may be incorporated into main chain, side chain or main chain and side chain, giving MCLC, SCLC and MCSCLC polymers respectively [50]. MCLC polymers show promise as high modulus, high melting thermoplastics, whereas SCLC polymers show promise as electroactive and electrooptical materials for optical data storage and non-linear optics [51]. For MCLC polymer the long stiff chains have only slight reorientational freedom in the LC or 'glassy' LC states, as has been shown from DRS studies [52,53]. Araki et al. [53] studied the following MCLC polymer where m is 2 or 3. For m = 3 a well-defined dielectric a process was observed having an apparent activation energy of 290 kJ mol ~ *. This material could not be aligned in directing electric fields [53].
Dielectric studies of SCLC polymers are more numerous (see [14], [54-60] and references therein). The dielectric behaviour of unaligned SCLC polymers gives little information on the underlying motions since the observed loss curves correspond to a superposition of several components. Alignment may be achieved using directing electric (E) or magnetic (B) fields or by surface forces.The alignment process in directing E fields is a dielectric phenomenon [58] and depends on the dielectric anisotropy Ae{a)) at the frequency / = a>/2n at which the E field is applied. Homeotropic (n\\Z) and planar ( n i Z ) alignments may be obtained by choice of the frequency of the directing E field. Here n is the LC director axis and Z is the laboratory axis defined as the normal to the parallel plates that confine the LC material. The two-frequency-addressing principle that leads to homeotropic (H) and planar (P) alignments for LC polymers has been reviewed [58]. Studies have been made of LCSC polymers of the following generic structures
where m denotes a spacer group; m typically lies in the range 2 < m < 12. R1 is H (for acrylates) or CH 3 (for methacrylates) and R2 is typically an alkylcyanobiphenyl group or an aromatic ester group, as follows
For such materials, which may be smectic or nematic liquid crystals, the dielectric properties of the LC phase are anisotropic. For a uniaxial LC phase, the dielectric tensor is diagonal such that
e((o) = diag [E1(O)), E1(G)), E1(O))]
where, for a material for which Ae(co) is positive at low frequencies, we find that S11(O)) and E1(O)) are measured for H-aligned and P-aligned samples respectively [58]. The permittivity E'(CO) and loss factor E"(O)) change markedly when a SCLC polymer is aligned in directing E fields or B fields [54-60]. As one example, Figure 11.1 shows plots of our recent results [61] for a carbon chain polymer having m = 2 and an alkyl cyanobiphenyl mesogenic head group in the side chain. The plot shows dielectric loss G/co = E"C0, where G is the equivalent parallel conductance of the sample and C 0 is its geometrical electrode capacitance, as a function of log(//Hz) and temperature for an unaligned sample (Figure 1 l.l(a)) and for the same sample that was aligned homeotropically using a low frequency E field (30Hz, 50 V across a 70 urn thick sample) (Figure ll.l(b)). For the unaligned sample one broad loss peak is observed, which moves rapidly to higher freqencies with increasing temperature. Only a slight change in property is observed when the material transforms from the LC state to the isotropic liquid at 89 0 C. Figure 1 l.l(b) shows data for the H-aligned sample. The loss peak in the LC state is nearly twice the height of that for the unaligned sample and much narrower, being only slightly broader than that for a single relaxation time process. As the clearing temperature Tc = 89 0 C is approached in the LC state there is a marked fall in the peak height to the level of that for the isotropic liquid. A part of the fall is due to the decrease in local order parameter S(T) as Tc is approached, and the remainder is due to the onset of the biphasic region, which in this case is restricted to « 2 0 C. These data serve to illustrate the anisotropic nature of molecular motions in LCSC and show (compare Figures 11. l(a) and (b)) that it is necessary to align samples macroscopically in order to reveal this property. DRS provides a particularly useful means of monitoring the nature and extent of macroscopic alignment in SCLC samples that have been subjected to E fields, B fields, surface forces or are aligning/disaligning after electrical and/or thermal treatments. As we have shown [56], the complex permittivity of a uniaxial sample of intermediate alignment is given, to a good approximation, by the linearaddition relationship E{O)) = (1 4- 2S d ) £|) M/3 + 2(1 - S6)E1(O))P
(12)
Here, S6 is a macroscopic director order parameter Sd = O c O S 2 ^ 2 - l > / 2
(13)
where 6nZ is the angle between a local director n and the laboratory Z axis (Z is
(GZo))ZpF
UNSUBTRACTED DATA FOR UNALIGNED LCP95
(GZ(O)/pF
UNSUBTRACTED DATA FOR HOMEOTROPIC LCP95
Figure 11.1 Plots of G/co = e"C0 as a function of log frequency/Hz and temperature for (a) unaligned and (b) homeotropically aligned SCLC polymer. Note the marked change in loss on melting the H-aligned material (Tc« 89 0C) and the lack of change on melting the unaligned material [61]
normal to the plane of the parallel electrodes, as described above). Thus Sd = 1,0, - 0 . 5 for H-aligned, unaligned and planarly aligned samples respectively. Application of Equation (12) using both its real part e'(ca) and/or its imaginary part s"(co) allows Sd to be determined for a sample of intermediate alignment if e\(co), s'^G)), e'±((o)9 and e'[(a>) are known. Two crossover frequencies occur at /', say, when s[((o) = ei(co), and at /", say when ej[(co) = el(cw). (Insertion of these conditions in Equation (12) show that e'((o) is independent of Sd for s'^co) = e'L(a>) and s"(co) is independent of Sd for ej,'(co) = £^(co)). The accuracy o the method for determining Sd can be checked via the consistency of Sd values determined at different frequencies through the spectral range, and this has been shown to be very successful in practice for siloxane polymers [54, 56, 57, 62]. Thus DRS provides a direct unambiguous means of determining the extent of macroscopic order, through Sd, in SCLC samples. We note that optical microscopy and infrared and Raman spectroscopy may not be used easily to monitor alignment in SCLC samples owing to the scattering of light by LC materials, but NMR provides a further method. Furthermore, DRS may be used to monitor the kinetics of alignment of SCLC polymers, as we have described [62, 63]. In our studies of a chiral nematic LC polymer, the changes of loss spectra with time as a sample realigned from P to H alignment in the presence of a steady d.c. E field were monitored, and were fitted using a continuum theory first described by Martins et al. [64] and further developed by Esnault et al. [65]. An important consequence of this theory is that it predicts that Sd reaches a plateau determined by the balance between dielectric forces (involving Ae £ 2 ) and elastic forces (involving elastic constants of the LC phase). It has been shown [58] that the ease of alignment in SCLC polymers is strongly dependent on chemical structure and the thermal/electrical treatments given to samples. In most cases it is difficult to align SCLC polymers in the LC state using directing E fields [56-58,66], so cooling from the melt with an a.c. E field of chosen frequency and amplitude may provide an alternative route— although dielectric breakdown is then a problem because E fields of 100 V/50 ^m are required, typically, in order to achieve full H of P alignment. An 'electrical cleaning' method may be used to reduce the extrinsic conduction of melt samples and hence to allow the sample to sustain higher aligning E fields in the melt before breakdown occurs (see [58] and references therein for a review). In addition to providing a method for determining the alignment of SCLC samples, DRS data also give information on the anisotropic reorientational dynamics of the dipolar mesogenic groups in a LC polymer. As we have shown [57,67], the generalization of the earlier theories [68,69] of dielectric relaxation of low molar mass liquid crystals can be achieved in the following way. The field-free orientation distribution function /°(Q 0 ) a f l d the field-perturbed orientation distribution function fE(Cl0) of mesogenic groups may be written as
expansions involving the Wigner rotation matrix elements DQ 0 as follows QO / 9 J
- L i X
-
AQ 0 ) = A Z - J V pJooDoo("o) j=0\
™
(14)
/
AOo) = *(l+^+-)AQo)
(15)
where A and B are normalization constants, D 0 0 are order parameters and D 00 (Q 0 ) describes the orientations of mesogenic dipolar groups (each of dipole moment fi) in Euler space with respect to the laboratory frame. When the dipolar units reorientate in the LC potential, the conditional probability of finding the dipole group in the orientation around Q at time t given that it was around the orientation Q0 at t = O is given formally by the further expansion
f(0,t/Qo,0) = S I Z DUO 0 )DL(O)GLW
<16)
J mn
where the G^n(O are time-correlation functions of the motion of the group, and are defined by the relationship <„(')=[
f/(Q,t/Q 0 ,0)D^ n (Q 0 )DU")dQ o dQ
(17)
The dipole moment in the laboratory frame (X, Y, Z) following the step withdrawal of the measuring electric field is then determined from the relationship [57]
W W ) = f f /E(Qo)/(ar/Qo,0)^P)dQodQ
(18)
JnoJn where P = O, ± 1 and relates to the laboratory axes. Insertion of Equations (14)—(16) into Equation (18) and using standard relations for integrals of triple products of the rotation matrix elements leads to expressions for (fifab(t)} and <ji£b(f) > ( = <^£b(f) >) in terms of weighted sums of Clebsch-Gordan coefficients. Relating these quantities to dipole polarization and hence to permittivities sfah(t) and s^h(t), then Fourier transformation into the frequency domain, gives the important result that [57] e,(co) = S0011 + ^ [ ( 1 + 2S)tiF{(co) + (1 - S)Ai12FJ(CO)] « » = Sooi + 3^r [(I + S)^1(W)
+ (1 + SWrfF^o)-]
(19a) (19b)
where G is a constant involving internal field factors, S0011 and S001 are limiting high frequency permittivities, S is the local order parameter in the LC and /z, and /*t are
the longitudinal and transverse components of the dipole moment of the mesogenic group. The different F)(O)) are given by F)(Co)- 1 — io>3[£j(r)] where 3 indicates a one-sided Fourier transform and the Cj(O are (real) time-correlation functions which correspond to certain linear combinations of the different O^in(0 as follows
^W = Oj0W; CJ(O = Oj1(O + 0^ 1 (O CiW = OL10W + *io(0 CiW = ol. x _ , w + ol. xx(t) + 0 ^ 1 W + Oi1W
(20a-d)
Thus, four active dielectric relaxation 'modes' are predicted, corresponding to the reorientational motions of /i( = (/^,/if)) with respect to the reference frame defined by the LC potential, giving two modes for e^o) and two for E1(O)). The strengths of these modes are governed by \i{j\iK and the magnitude of the local order parameter S (and hence depend on sample temperature). When a SCLC polymer is aligned macroscopically, then e^co) and e±(o)) are unchanged, as these are quantities that relate to local regions in the LC material and are defined in terms of molecular properties (see Equation (19)). However, the macroscopic dielectric properties change with macroscopic alignment as Sd is changed, as we have discussed above in connection with Equation (12) and Figure 11.1. For H alignment (Sd = 1), E,,(CO) is measured, and is usually dominated by F^1(O)), which is the S process (or W mode), and this is the case in Figure 11.1. The correlation function for this process is given by OJ0(O =
(21)
where O is the polar angle between the molecular z axis and the laboratory Z axis. In many cases samples initially in an unaligned state may be aligned homeotropically but not planarly, owing to experimental difficulties of two kinds: (i) that the frequency required to prepare P alignment is too high for practical generators in the temperature range where P alignment is possible, and (ii) that Ae is too small to allow realignment to the P state to occur in practical time-scales for applied fields below breakdown levels. In these cases, if data for unaligned (U) and H-aligned samples have been obtained, then ^1(a>) may be calculated as follows. For the unaligned sample Sd = O, so Equation (12) gives £uM
= [IS1(Oi) + 2 e » ] / 3
(22)
£xM = [3%M-e (l (a>)]/2
(23)
thus Thus, the permittivity and loss spectra for P-aligned material may be calculated from the permittivity and loss spectra of U- and H-aligned material. We note that the above treatments consider only the anisotropic dielectric properties arising from the anisotropic motions of dipolar mesogenic groups in
the LC phase. However, it has been emphasized by Haase that isotropic motions, e.g. of a dipole unit in the main chain removed from the dipolar mesogenic head group will give an additional term in each of Equations (19) which is independent of S, the local order parameter. This is important for acrylate and methacrylate polymers with mildy polar mesogenic groups but of lesser importance for siloxane polymers with strongly polar mesogenic groups of the kind we have investigated [54-58].
11.5 REAL-TIMESTUDIESOFCHEMICAL AND PHYSICAL CHANGES The speed of operation of modern measuring equipment now allows measurements of e(co) to be made in real time as a system undergoes chemical or physical transformations such as polymerization or crystallization, respectively. This capability greatly extends the applications of DRS in polymer science and poses new challenges for theoretical interpretations of the observed dielectric phenomena, as we shall mention below. One subject presently receiving much attention is that of the bulk step polymerization of epoxides with amines. Dielectric spectroscopy of such systems has been studied for many years, notably by Lane and Seferis [70], and has been reviewed by Senturia and Sheppard [71,72]. In a series of recent papers, Mangion and Johari [73-78] have given extensive data for permittivity e'(co,t,TK) and loss factor e"(a),t, TR) for the diglycidyl ether of bisphenol-A (DGEBA) reacting with diaminodiphenylmethane (DDM) and/or diaminodiphenylsulphone (DDS) as a function of reaction time t for fixed measuring frequencies ( / = l/27ico) and fixed reaction temperatures TR. Samples of different compositions of DGEBA/DDM and DGEBA/DDS and DGEBA/DDM/DDS were investigated in real time; also, post-cured materials were studied over a wide range of temperature, including the glassy state at low temperatures, in order to detect changes in sub-Tg processes resulting from cure and post-cure. It was found that, as a reacting mixture transformed from a liquid to a (thermoset) glass, e'((o, t> TK) showed dispersion and s"((o, f, TR) showed an absorption peak. Thus, the DRS method detects 'vitrification' of the thermosetting material, but this is a relaxation phenomenon whose interpretation relates to the measuring frequencies used. As one example of the dielctric behaviour of a thermosetting system Figure 11.2 shows real-time e\ a" data obtained recently by us [79] for a 1:2 molar mixture of DGEBA with an alicyclic diamine diaminodicyclohexylmethane (DDCM)
In Figure 11.2(a), s'(co,t, TR) is plotted against time for different measuring
(a) Variation of e' as a function of time for the thermoset (DGEBA/DDCM) at Tcure= 60 0C
e1
note: the frequency step between the curves is 0.4 in Iog(f/Hz)
Log(F)=5
Log(F)=3
time / s (b) Variation of e" as a function of time for the thermoset (DGEBA/DDCM) at 1=60 0C
e"
Log(f)=5
Log(f)=3
time / s
Figure 11.2 Plots of e' and e" against reaction time at different measuring frequencies for a 1:2 molar mixture of DGEBA with DDCM at 60 0C [79] frequencies, and dielectric dispersion is seen to occur and move to longer times as the frequency is lowered. Figure 11.2(b) shows the complementary data for fi"(co, U 7R)> and a well-defined loss peak is seen to occur that moves to longer times as / is lowered. Such data were obtained by continual measurements of e\e" at chosen frequencies and recording the time of each measurement. Post-processing
of these data allows values of s\ e" to be calculated for chosen times, thus enabling plots of e\e" to be made as a function of frequency, as shown Figure 11.3, where dispersion and absorption are now seen in a familiar representation. An instructive representation of these data is provided by the 3D plots of permittivity and loss as a function of both log / and t, as shown in Figure 11.4. These data show that an a process, far broader than the 'normal' a process described above for conventional amorphous polymers, is observed, which moves rapidly to lower frequencies as reaction time increases. Evidently, when the frequency of maximum loss e'n occurs below, say, 10" 2 Hz, we would regard the material to be a glass operationally; extrapolation of such dielectric data to lower frequencies thus provides a method of dielectrically monitoring the 'effective cure-time' of a given reaction mixture, and this is useful for practical thermosetting systems. Mangion and Johari [73-78] have made quantitative analyses of their data. For example, they found that plots of logr cure vs 1/TR for a fixed measuring frequency were linear. Here, fcure is defined as the time at which s" = e^ for the given measurement frequency and given reaction temperature. They derived an apparent activation energy which was 47 kJ mol" 1 and 44.51CmOl"1 for their DGEBA/DDM and DGEBA/DDS mixtures, respectively. With regard to the time-temperature variation of e' and s", they write ^cure) = £oo(^cure)+[^cure)~£ao(tc«re)](
eXP - k J
^
~
d
* ( 24 )
where rcurc is the reaction time, £(*curc) = S(Q)9 tcurc). A difficulty with this equation is that a Fourier transform relationship between the frequency-dependent permittivity e((o) and a relaxation function 4>(t) is valid only for a stationary system: i.e. one whose thermodynamic properties are independent of time.
e1
(a) Variation of e' with the frequency of measurement for the thermoset (DGEBA/DDCM) at 7 = 60° C measurements every 6 mins, 4 ^ 0 , « 25 min
Logf/Hz
eM
(b) Variation of e" with the frequency of measurement for the thermoset (DGEBA/DDCM) at 7 = 60° C
Logf/Hz
Figure 113 Plots of ef and e" against log frequency/Hz for different reaction times for a 1:2 molar mixture of DGEBA with DDCM at 600C [79]. Curves 1-10 correspond to 25 min to 225 min in increments of 6 min
e'
(a) Evolution of the permittivity 7 = 6 0 0C
e"
(b)Evdlution of the loss factor T=60°C
Figure 11.4 3Dplotsofe' ande" against reaction time for a 1:2 molar mixture of DGEB A with DDCM at 600C [79] are fitted approximately by the KWW function. Thus, fixing the frequency f=co/2n and constructing the Argand diagram for e' and e" variations as t changes would lead to a skewed arc of KWW type. However, the variations of the KWW parameter with reaction time need to be investigated.
In many studies of polymerizing systems, low frequency conductivity-related processes are large and may obscure the dipole relaxation process. Some evidence of these processes is seen in Figures 11.2-11.4 at low reaction times and low frequencies, but in other systems they may dominate the overall behaviour, e.g. as in the case of phase-separating elastomer-epoxy resins described by Pethrick and coworkers [80, 81] and by Maistros et al. [82]. In such cases the apparently dominating conductivity processes may be represented alternatively by the modulus representation [73] M = [1/e], which gives M' = ^
;
A*" = ^
(25a,b)
Such a duality of representations is well-known in dynamic mechanical relaxation [3] for compliance and modulus. The apparent correlation times Compliance a n d Modulus a r e related through the ratios (E0/EJ = [JJJ0), where subscripts 0, oo refer to limiting low and high frequencies respectively and E and J are modulus and compliance respectively. In the view of the author, the permittivity representation should be used for the dipolar relaxations in cure-monitoring data. The M representation may be useful where it is known that conductivity processes are involved, but in this numerical representation space-charge and electrode processes appear to have less importance than their true contributions. Finally, we note that present applications of DRS to polymerizing systems have emphasized epoxy-amine thermosetting mixtures. Clearly much information could be obtained for photosetting and thermosetting addition polymerizations. We have conducted such studies recently using dimethacrylate monomers, which are photopolymerized using blue light (480 nm) and a camphoroquinone-amine initiator [83]. The dielectric data obtained during photoreaction are qualitatively different from those shown in Figures 11.2-11.4 for a thermosetting system. Analysis of our data show that the a process in the unreacted liquid monomer transforms into two processes as reaction proceeds, and that a /? process remains at low frequencies when polymerization is essentially completed. It is anticipated that DRS will be applied to real-time studies of phaseseparating polymer-polymer mixtures and to crystallizing polymers (see the data of Tidy and Williams in [12] for early results for isothermal crystallization of amorphous polyethylene terephthalate). Although such measurements are entirely feasible in the usual low frequency region 10-10 7 Hz, it would be highly desirable to develop experimental dielectric cells and fast measuring methods for the high frequency range 10 8 -10 1 0 Hz, thus providing the wide frequency range necessary to document and define the multiple dielectric relaxations that arise from dipole motions in the different phases.
11.6 CONCLUSIONS AND FUTURE PROSPECTS It is apparent that DRS provides an important method for studying the reorientational dynamics of polymer chains in bulk amorphous, crystalline and liquid-
crystalline polymers and in polymerizing, phase-separating and crystallizing systems. The merits of DRS include small sample size (typically 1 cm 2 x 50 jim), wide frequency range (typically 10"3 to 107 Hz) and its sound theoretical basis both in phenomenological and molecular terms. Difficulties with DRS include (i) low frequency conductivity-related processes, which may obscure the dipole relaxation processes, and (ii) the present limited access to high frequency (10 8 10 10 Hz) techniques. It is anticipated that DRS studies of polymers will be extended to more complicated systems, e.g. where two phases may be present and electrical-field-induced processes occur. One example is polymer-dispersed liquid crystals (PDLC), which may be formed in several ways, in which liquid-crystal droplets are dispersed in a polymer matrix. Such PDLC materials as films 50-500 ^m in thickness may be switched optically using moderate E fields, and the required voltage depends both on the response of the liquid crystal and the dielectric properties of the host polymer. In such cases a knowledge and control of the dielectric properties of guest and host materials will enable the optical switching properties to be optimized, especially in regard to the lowering of the E field required for switching. A further example where DRS properties are useful is non-linear-optical (NLO) films, which show promise for second- and thirdharmonic generation of laser light [84]. It is generally observed that the nonlinear susceptibility for second-harmonic generation of E-poled films decays with time, and this may be due partly to the intrinsic motions of the electroactive molecules in the glassy amorphous or glassy LC state of the poled material. Such motions can be characterized using DRS, and this provides direct information on such electro-optical phenomena which may not be readily obtained by other methods such as NMR and quasi-elastic light scattering.
11.7 ACKNOWLEDGEMENTS The author acknowledges the support of the SERC and AFOSR and the Erasmus scheme for studies of the dielectric properties of polymers, and thanks Prof. CIi ve Bucknall for information regarding the 'Axum' programs used for the 3D plots shown in this paper.
11.8 REFERENCES [1] CP. Smyth, Dielectric Behaviour and Structure, McGraw-Hill, New York, 1955. [2] N.E. Hill, W.E. Vaughan, A.H. Price and M. Davies, Dielectric Properties and Molecular Behaviour, Van Nostrand, London, 1969. [3] N.G. McCrum, B.E. Read and G. Williams, Anelastic and Dielectric Effects in Polymeric Solids, Wiley, London, 1967, Dover, New York, 1991. [4] C.J.F. Bottcher and P. Bordewijk, Theory of Electric Polarization, Vol. 2, Elsevier, Amsterdam, 1978.
[5] A.K. Jonscher, Dielectric Relaxation in Solids, Chelsea Dielectrics Press, London, 1983. [6] J. Wong and CA. Angell, Glass; Structure by Spectroscopy, Marcel Dekker, Basel, 1976. [7] G. Allen and J.C. Bevington (Eds.), Comprehensive Polymer Science, Vol. 1, Polymer Characterization, Pergamon Press, Oxford, 1989. [8] E. Riande and E. Saiz, Dipole Moments and Birefringence of Polymers, Prentice Hall, New Jersey, 1992. [9] F.I. Mopsik, Rev. ScL Instrum., 1984,55, 79. [10] J.M. Pochan, JJ. Fitzgerald and G. Williams, in B.W. Rossiter and R.C. Baetzold (Eds.), Determination of Electronic and Optical Properties, Physical Methods of Chemistry Series, 2nd edn., Vol. VIII, Wiley, New York, 1993, Ch. 6, p. 392. [11] G. Williams, Chem. Rev., 1972,72, 55. [12] G. Williams, Adv. Polym. ScL, 1979,33,60. [13] G. Williams and D.C. Watts, in NMR Basic Principles and Progress, Vol. 4, NMR of Polymers, Springer Verlag, Heidelberg, 1971, p. 271. [14] G. Williams, in G. Allen and J.C. Bevington (Eds.), Comprehensive Polymer Sciences, Vol. 2, Polymer Properties, Pergamon Press, Oxford, 1989, p. 601. [15] M. Cook, D.C. Watts and G. Williams, Trans. Faraday Soc, 1970,66, 2503. [16] G. Williams, M. Cook and PJ. Hains, J. Chem. Soc, Faraday Trans, 2,1972,69,1045. [17] G. Williams, in R. Pethrick and R.W. Richards (Eds.), Dynamic Properties of Solid Polymers, NATO ASI, Reidel, Dordrecht, 1982. [18] G. Williams and D.A. Edwards, Trans. Faraday Soc, 1966,62,1329. [19] G. Williams, Trans. Faraday Soc, 1964,60,1548,1556. [20] G. Williams, Trans. Faraday Soc, 1966,62, 2091. [21] G. Williams, D.C. Watts, Trans. Faraday Soc, 1970, 66, 80. [22] G. Williams, D.C. Watts, S.B. Dev and A.M. North, Trans. Faraday Soc, 1971,67, 1323. [23] G. Williams, IEEE Trans. Electr. Insul., 1982, El-17, 469. [24] N. Koizumi and Y. Kita, Bull Inst. Chem. Res. Kyoto Univ., 1978,56, 300. [25] G. Williams in M. Davies (Ed.), Dielectric and Related Molecular Processes, Vol. 2, Chemical Society, 1975, p. 151. [26] G. Williams, IEEE Trans. Electr. Insul., 1985, El-20, 843. [27] K.L. Ngai, in T.V. Ramakrishnan and M. Raj Lakshmi (Eds.), Non-Debye Relaxation in Condensed Matter, World Scientific, Singapore, 1987, p. 1. [28] W. Goetze, Rep. Progr. Phys., 1992,55, 241. [29] W. Goetze, Ferroelectrics, 1992,128, 307. [30] G.F. Mazenko, J. Non-Cryst. Solids, 1991,131-133,120. [31] A.K. Rajagopal, K.L. Ngai, and S. Teitler, J. Non-Cryst. Solids, 1991,131-133,282. [32] K.L. Ngai, Commun. Solid. State Phys., 1979,9,127, 141. [33] K.L. Ngai, A.K. Rajagopal and S. Teitler, J. Chem. Phys., 1988,88, 5086. [34] G. Williams and M. Cook Trans. Faraday Soc, 1971,67,990. [35] R.H. Cole, J. Chem. Phys., 1965, 42, 637. [36] U.M. Tituiaer and J.M. Deutch, J. Chem. Phys., 1974,60,1502. [37] G. Williams in E. Wyn-Jones (Ed.), Chemical and Biological Applications of Relaxation Spectroscopy, Reidel, Dordrecht, 1975, p. 515. [38] G. Williams and D.C. Watts, in Dielectric Properties of Polymers, Plenum Press, 1971, p. 17. [39] G. Williams and D.C. Watts, Trans. Faraday Soc, 1971, 67, 2793. [40] G.D. Smith and R.H. Boyd, Macromolecules, 1991, 24, 2725, 2731. [41] V. Rosato and G. Williams, Adv. MoL Relax. Processes, 1981, 20, 233.
[42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [81] [82] [83] [84]
R.L. Jernigan, in Dielectric Properties of Polymers, Plenum Press, 1971, p. 99. M.S. Beevers and G. Williams, Adv. MoL Relax. Processes, 1975, 7, 237. J.D. Hoffman, G. Williams and E. Passaglia, J. Polym. ScU Part C, 1966,173. Y. Ishida, J. Polym. ScU 1969,7,1835. CR. Ashcraft and R.H. Boyd, J. Polym. ScU Polym. Phys. Ed., 1976,14, 2153. H. Frohlich, Proc. Phys. Soc. (London), 1942,54,422. G. Williams, J.D. Hoffman and J.I. Lauritzen, J. Appl. Phys., 1967,38,2503. M. Mansfield and R.H. Boyd, J. Polym. ScU Polym. Phys. Ed., 1978,16,1227. Adv. Polym. ScL, 1984,60/61. CB. McArdle, (Ed.), Side Chain Liquid Crystal Polymers, Blackie, Glasgow, 1989. F. Laupretre, C Noel, W.N. Jenkins and G. Williams, J. Chem. Soc, Faraday Discuss, 1985, 79,191. K. Araki, M. Aoshima, N. Namiki, S. Ujiie, N. Koide, K. Imamura and G. Williams, Makromol. Chem. Rapid Commun., 1989,10, 265. G.S. Attard, K. Araki, JJ. Moura-Ramos and G. Williams, Liq. Cryst., 1988,3,861. A. Kozak, JJ. Moura-Ramos, G.P. Simon and G. Williams, Makromol. Chem., 1989, 190, 2463. G.S. Attard, K. Araki and G. Williams, Br. Polym. J., 1987,19,119. K. Araki, G.S. Attard, A. Kozak, G. Williams, G.W. Gray, D. Lacey and G. Nestor, J. Chem. Soc, Faraday Trans. 2,1988,84, 1067. A. Nazemi, G. Williams, G.S. Attard and F.E. Karasz, Polym. Adv. TechnoL, 1992,3, 157. W. Haase, H. Pranoto and FJ. Bormuth, Macromolecules, 1985,18,960. FJ. Bormuth and W. Haase, MoL Cryst. Liq. Cryst., 1987,148,1. G. Williams and J. Hayden, manuscript in preparation. A. Kozak, G.P. Simon and G. Williams, Polym. Commun., 1989, 30,102. A. Kozak, G.P. Simon, J.K. Moscicki and G. Williams, MoL Cryst. Liq. Cryst., 1990, 193,155. A.F. Martins, P. Esnault and F. Volino, Phys. Rev. Lett., 1986, 57,1745. P. Esnault, J.P. Casquilho, F. Volino, A.F. Martins and A. Blumstein, Liq. Cryst., 1990,7,607. G.S. Attard, G. Williams and A.H. Fawcett, Polymer, 1990,31, 928. G.S. Attard, MoL Phys., 1986,58,1087. W. Maier and G. Meier, Z. Naturforsch., 1961,162,1961. P.L. Nordio, G. Rigatti and U. Segre, MoL Phys., 1973,25,129. J.W. Lane and J.C Seferis, J. Appl. Polym. ScL, 1986,31,1155. S.D. Senturia and N.F. Sheppard Jr., Adv. Polym. ScL, 1986,80,1. N.F. Sheppard, Jr. and S.D. Senturia, Polym. Eng. ScL, 1986, 26, 354. M.B.M. Mangion and G.P. Johari, J. Polym. ScL, Polym. Phys., 1990,28, 71. M.B.M. Mangion and G.P. Johari, J. Polym. ScL, Polym. Phys., 1990, 28,1621. M.B.M. Mangion and G.P. Johari, J. Polym. ScL, Polym. Phys., 1991, 29,1117. M.B.M. Mangion and G.P. Johari, J. Polym. ScL, Polym. Phys., 1991,29, 1127. M.B.M. Mangion and G.P. Johari, Polymer, 1991, 32, 2747. M.B.M. Mangion and G.P. Johari, Macromolecules, 1990,23, 3867. C Duch, J. Fournier and G. Williams, manuscript in preparation. AJ. MacKinnon and R.A. Pethrick, Macromolecules, 1992,25, 3492. D. Lairez and R.A. Pethrick, Macromolecules, 1992, 25, 7208. G. Maistros, H. Block, CB. Bucknall and LK. Partridge, Polymer, 1992, 33,4470. G. Williams and J. Fournier, manuscript in preparation. D.S. Chemla and J. Zyss (Eds.), Nonlinear Optical Properties of Organic Molecules and Crystals, VoIs. 1 and 2, Academic Press, Orlando, 1987.
12
LIGHTSCATTERING FROM POLYMER SYSTEMS R, W. RICHARDS Department of Chemistry, University of Durham, Durham DHl 3LE, UK
12.1 INTRODUCTION Light scattering from dilute polymer solutions has long been used by polymer scientists to determine the molecular weights, the radii of gyration and the second virial coefficients of polymers [1,2]. Such information has been instrumental in the development of two-parameter theories [3] of polymer solutions, and classical intensity light scattering has been exhaustively discussed. In addition to two of the parameters mentioned (mean molecular weight and second virial coefficient), classical intensity light scattering has also been used to obtain information on the compositional dispersity in copolymers, and depolarised light scattering can provide information on molecular anisotropy in rod-like polymerss such as the main chain liquid crystal polymers [4,5]. Both of these applications have attendant difficulties. For copolymers, light scattering measurements have to be made in at least three solvents with different refractive indices to obtain compositional heterogeneity, although the judicious choice of solvents can considerably simplify the calculation process [6]. Depolarised light scattering intensity is often low and the signal-to-noise ratio can also be low. These aspects of light scattering have been plentifully discussed and will not be reviewed here. A more recent development has been quasi-elastic light scattering [7,8], which is often used to obtain diffusion coefficients of polymers in dilute solution. Quasi-elastic light scattering has been applied to solvent-swollen cross-linked networks and to semi-dilute solutions of polymers in an effort to investigate scaling theories and reptation theories. Additionally, the possibility of obtaining the viscoelastic properties of spread films of polymers at the air-water interface by quasi-elastic light scattering has also been discussed. Both of these aspects will be reviewed here. Light scattering from solid polymer films and melts was first reported some 40 years ago [9]. The experimental difficulties have been considerably simplified by newer technology, and the increased power of computing has eased the process of Polymer Spectroscopy. Edited by Allan H. Fawcett © 1996 John Wiley & Sons Ltd
data analysis. Much simpler experiments than implied in the original theory are capable of producing information on the kinetics of phase-separating blends and the thermodynamics of the systems. An overview of small angle light scattering applied to semi-crystalline polymers and phase-separating blends will also be given here. All of these different types of experiment are united by a common source for the light scattering observed, that is, the existence of fluctuations in polarisability, and hence refractive index, at microscopic length scales in the material upon which the light is incident. The cause of these fluctuations may differ markedly from thermal fluctuations (surface quasi-elastic light scattering), concentration gradients (quasi-elastic light scattering), density variations due to packing (crystallinity) etc., but such fluctuations scatter light efficiently, so that light scattering provides a convenient non-perturbative probe of the structure and dynamics of polymer systems.
112 SMALL ANGLE LIGHT SCATTERING (SALS) 12.2.1 SEMI-CRYSTALLINEPOLYMERS The pioneering work in SALS was done by Stein and his collaborators [10] some 30-40 years ago, and it is a relatively simple technique for investigating spherulite growth and size in nearly transparent polymers. The original equations derived by Stein and Rhodes [10] sought to explain the 'four leaf clover' pattern of scattered light intensity observed in the Hv scattering experimental set-up shown schematically in Figure 12.1. (Hv implies vertically polarised incident light, horizontally polarised scattered light.) Spherulites are optically anisotropic, and originally Stein and Rhodes explained the disposition of the scattered light as lobes along the azimuthal angle of n/4 as being due to the orientation of dipoles in the spherulite with respect to the plane of polarisation of the incident light. Much interesting work has been analysed on the basis of the original equations (e.g. influence of deformation [12], size distribution of spherulites [13], influence of chain branching [14]). However, this explanation has been shown to be fundamentally incorrect by Meeten and co-workers [15-17], although the extraction of parameters such as spherulite size etc. from SALS data using the original equations is not altered. Meeten and Navard [16] used Mie theory, RayleighGans-Debye theory and anomalous diffraction theory on isotropic spheres and the latter two theories on anisotropic spherical scatterers. For all theories, both types of particles produced 'four leaf clover' scattering patterns for Hv scattering. The dependence of scattered intensity on azimuthal angle (f> and scattering angle 0 can be calculated from the dimensionless angular gaing (which is approximately the ratio of the intensity of the scattered light to the incident
Potariser
Laser
Sample Analyser
Detector Plane
Figure 12.1 Schematic of small angle light scattering experiment: scattering angle = 0; the azimuthal angle <j> is measured from the plane of polarisation of the incident light light intensity). G^ = ( I A V ) I S 1 - S 2 I 2 sin 2 20 GVv = (4AV)IS 1 Sm 2
5 2 = S1COsO V = wm/ns with nm and ns being the refractive index of the matrix and the sphere respectively.
Anisotropic sphere 2ik3r3 51 =
3
{3(/x— l)(sinw —MCOSW) + A / * [ u c o s K - 4 s i n u + 3Si(u)]}
2ifc3r3 52 =-^~3-{3(/I- I)(sinu-Mcosw)cos0- A^(I + cos2(0/2))(wcosw-4sinwSi(u))} where jx = (nr + 2nt)/3nm, Afi = (nr — nt)/nm, nT is the radial refractive index, nt the tangential refractive index, and Si(u) is the sine integral of u. For both cases r is the radius of the sphere and u = An/X r sin(Q/2), with A being the wavelength of light in the scattering polymer film. Figure 12.2 shows the form of the scattering for both isotropic and anisotropic spheres and Hv scattering. Note that both show a maximum scattering disposed in lobes at azimuthal angles of n/4. Figure 12.3 shows the intensity variation with u along one such lobe; again both have the same qualitative features, i.e. a maximum at a defined value of u. For
Figure 12.2 (Continued)
Figure 12.2 Contour plots of Hv scattered light intensity from (a) optically anisotropic spheres (b) optically isotropic spheres; x = r sin <£, y = r cos <£, where r is the radius of the sphere anisotropic spheres wmax = 4.09, for isotropic spheres wmax = 2.74 for the first order maximum. Over the years the method of detecting the scattered light has improved; originally photographic methods were used, followed by high speed cameras, vidicons and optical multichannel analysers. Nowadays CCD cameras are able to record scattering patterns digitally, and fast shutter speeds mean that the data can be recorded in real time. A typical apparatus as constructed at Durham [18] is shown in Figure 12.4; a description of similar equipment has recently appeared [19]. Typically, the fastest shutter speed may be «20/xs and the time needed to refresh the detector area and store 4K pixels each with 18 bit dynamic range is « 2 s. One system on which this apparatus has been used is the crystallisation
10 Intensity
Intensity 4
Figure 12.3 Variation of Hv scattered light intensity at <j> = 45° for (a) optically anisotropic spheres, (b) optically isotropic spheres; in each case u = (4nnr/Ao)sin(0/2). (a) Reprinted with permission from Macromolecules, 1982, 15, 1004; (b) reprinted with permission from [53]. Copyright 1982 and 1993 American Chemical Society kinetics of linear diblock copolymer of methyl methacrylate and ethylene oxide [20]. A copolymer with 76 mol % of ethylene oxide was quenched to a temperature of 308 K from 400 K and the Hv SALS pattern was recorded as a function of time; the variation of intensity obtained is shown in Figure 12.5. From the maxima in such curves, the radius of the spherulite was obtained, and was used in an Avrami analysis of the crystallisation kinetics [21]. This provides a parameter proportional to the rate of crystallisation and the Avrami exponent. From the latter parameter it is sometimes possible to infer something about the growth mechanism and geometry (Table 12.1). For this block copolymer the Avrami
Fibre optic link
CCD Camera
Marata plate
analysor
hot stage+ sample
polariser
ND filters
beam expander
Figure 12.4 Block diagram of a small angle light scattering instrument exponent obtained was 1.8; the equivalent homopolymer blend had an Avrami exponent of 1.5. A word of caution is appropriate here. Meeten's analysis shows that the value OfU1114x depends on the anisotropy of the spherulite; thus in the early stages the observed growth rate may appear to be smaller than the actual growth rate.
Intensity
PEO-b-PMMA (76% w/w PEO) Block Copolymer, 7C=35 0C
Polar scattering angle
Figure 123 Variation of Hv scattered light intensity at <j> = 45° for a linear diblock copolymer of methyl methacrylate and ethylene oxide. The copolymer was quenched from 1000C to a crystallisation temperature of 350C and the time lapse between each curve is 2.1s
Table 12.1 Avrami exponents predicted for a variety of growth geometries and growth control mechanisms Exponent Nucleation" Growth geometry Growth control* 0.5 Instantaneous Rod Diffusion 1 Instantaneous Rod Interface 1 Instantaneous Disc Diffusion 1.5 Instantaneous Sphere Diffusion 1.5 Homogeneous Rod Diffusion 2 Homogeneous Disc Interface 2 Homogeneous Disc Diffusion 2 Homogeneous Rod Interface 2.5 Homogeneous Sphere Diffusion 3 Instantaneous Sphere Interface 3 Homogeneous Disc Interface 4 Homogeneous Sphere Interface 'Instantaneous: nuclcation on existing heterogeneities. Nuclei form simultaneously at beginning. Homogeneous: sporadic formation of nuclei. Nucleation continuous in the untransformed material. ^Diffusion controlled: kinetics are controlled by the rate of diffusion of molecules to the nuclei. Interface controlled: kinetics controlled by rate of attachment of molecules to the nuclei.
12.2.2 PHASE-SEPARATING POLYMER MIXTURES Generally, compatible polymer mixtures display a lower critical consolute (or coexistence curve), and the variation of the Gibbs free energy change on mixing with composition shows a pair of inflection points above a particular temperature defined by the thermodynamics of the system and the mixture composition. At these inflection points (d2AGJd<j>2)TtP = 0, and they define the locus of the spinodal curve. This curve is the limit of stability of the mixture, i.e. within the curve the mixture is unstable to any fluctuation and demixing (spinodal decomposition) takes place spontaneously. Between the coexistence curve and the spinodal curve there is a metastable region wherein large fluctuations are necessary to initiate demixing, usually via a nucleation and growth process (Figure 12.6). Both curves meet at the critical temperature where (d2AG/d
Temperature (K)
binodal spinodal
Figure 12.6 Schematic phase diagram for polymer-polymer mixtures
gradient due to composition gradients. The solution to this equation is a Fourier series describing the compositional fluctuations in the system, i.e. the local compositional deviations from the average value, and these fluctuations are the source of the scattered light intensity. Since light scattering is described in Fourier space terms, the solution to the Cahn-Hilliard equation is also needed in Fourier space, Le. in terms of a wave vector, where the wave vector is (In/X) and X is the concentration fluctuation wavelength. The intensity of light scattered from the phase-separating mixture is proportional to the square of the fluctuation amplitudes and is given by: / ( a 0 = /(Q^ = 0)exp[2R(Q)r] where Q is the scattering vector = Ann sin(0)/Ao, with n the refractive index of the sample, 20 the scattering angle and X0 the wavelength of light in vacuo. The term R(Q) is known as the amplification factor and R(Q) = - M(d2A/d
10- 3 Q(Cm- 1 )
Figure 12.7 Scattered light intensity (Vv conditions) as a function of angle for different times for a phase-separating mixture of polystyrene and polyvinyl methyl ether. Times after start of phase separation (seconds) A 2 +20 D40 O60 V 10 x 30 O 50 • 70
10ln(I(Q max )
Time (S)
Figure 12.8 Exponential dependence of scattered light intensity for a phase separating mixture of polystyrene and poly vinyl methyl ether at Qmmx fluctuation wavelength which will grow preferentially as phase separation proceeds. This leads to a maximum in the observed scattered intensity at a finite value of Q. The position of this intensity maximum does not alter as in the early stages of phase separation but increases in intensity as phase separation proceeds. At late stages in the phase separation, there is a coalescence of particles via Ostwald ripening and the maximum will shift to lower Q values. During the early stages, at a fixed value of Q, the scattered light intensity from a spinodally decomposing system should increase exponentially with time, and from this relationship the amplification factor can be obtained. A set of light scattering data for a demixing polystyrene/polyvinyl methyl ether (PVME) mixture collected at discrete time intervals is shown in Figure 12.7 [23]; the dependence of the scattered intensity on time at the Q value (Qmax) where the maximum intensity is seen is shown in Figure 12.8. From values of R(Qn^x) the effective diffusion coefficient De, can be obtained, as Dt = 2R(QmAX)/Qliax. At the spinodal curve Dt = 0, and thus if Dc is obtained for a series of composition and over a range of temperatures, the spinodal curve can be obtained. Figure 12.9 shows values of Dc as a function of temperature obtained for the mixture of polystyrene and PVME referred to earlier, and the spinodal curve predicted from these data is given in Figure 12.10. Light scattering investigations of other demixing polymers have been reported elsewhere [24, 25] and recently a very sophisticated instrument for such studies has been described [26].
1O10C-D9)Cm2S'1
PS(2)/PVME
Temperature (K) Figure 12,9 Effective diffusion coefficient as a function of temperature
B
PS(2)/PVME
TEMPERATURE (K)
cloud point curve
WEIGHT FRACTION PS
Figure 12.10 Spinodal curve (•) predicted from temperature dependence of Dc for polystyrene/polyvinyl methyl ether mixtures
123 QUASI-ELASTIC LIGHT SCATTERING (QELS) 12.3.1 DILUTE POLYMER SOLUTIONS Light scattering by polymers in solution is not a perfectly elastic process, small amounts of energy being transferred between molecules and photons. This energy transfer leads to a broadening of the frequency of the scattered light relative to the incident light, and the intensity variation of the scattered light over a frequency range from — oo to + oo is the spectral density or power spectrum, which is given by I((o) = 1/2« I °° < E*{t)E{t + T) > exp iojTdt where <£*(r)£(t+T)> is the electric field autocorrelation function gt(t). In quasi-elastic light scattering (QELS) what is actually obtained as the output from the photomultiplier tube is the unnormalised intensity autocorrelation function G2(t\ and G2(t) = A + [Bg 1 (O] 2 (homodyne) where A is a constant background intensity to which the correlation function decays after a suitably long delay time f, and B is a constant close to unity. If we have a single species in the solution, e.g. a monodisperse polymer, and there are only concentration gradient relaxation processes, then ^1(O = exp(-Ff) and F l is the relaxation time of the diffusive process of the polymer down the concentration gradients; F = DQ2 with Q = (4nn/Xo)sin(0/2) and D is the translational diffusion coefficient. For polymer solutions, D is concentration dependent D = D 0 (l + fcDc) where D 0 is the infinite dilution value of D and c is the concentration of polymer. The term kD is composed of thermodynamic and factional parameters for the polymer in the particular solvent conditions investigated. Polymers are not often monodisperse, and each different relaxation time will make a contribution to the observed average F. A popular method of obtaining the diffusion coefficient is to use the cumulants approach outlined by Koppel [27] and the algorithm of Pusey et al. [28] In Q1(Z) = - T11 + (F2/2!)r2 - (F3/3!)r3 + • • • Generally only the first two cumulants can be extracted from the correlation functions with any confidence, and TJQ2 = D29 the z-average diffusion coefficient. About 12 years ago, Burchard et al. [29] showed that r JQ2 = D(I+
CR2Q2)
Hq 2 XiO 8 (Cm 2 S 1 )
Cj 2 XiO- 10 ^kC(Cm 2 )
Figure 12.11 Dynamic Zimm plot for polystyrene in toluene. Reproduced with permission of the American Chemical Society from ref. [29] where R9 is the radius of gyration of the polymer molecule and C is a parameter related to the molecular architecture and the thermodynamic environment. Incorporating the concentration dependence of D, TJQ2 = D0(I + kDc)(l + CRlQ2) Thus, as c->0 and Q->0, TJQ2 = D 0 and D 0 , kD and C can be obtained from a 'dynamic' Zimm plot (Figure 12.11). The slope of the line dependent on Q2 alone is CR2, whereas the slope of the line dependent on c only is DokD. Thus method has not been widely used; however, it has been applied to naturally occurring polymers to extract C and thus to enable something to be said about their structure. We noted earlier that each relaxation time will contribute to T and hence influence the shape of the correlation function. Consequently, all the information on polymer polydispersity is contained within the intensity correlation function because D is proportional to (molecular weight)"0. The extraction of the molecular weight distribution from the correlation function is an 'ill posed problem', as there are an infinite number of solutions to the Laplace inversion of the data that is required to obtain the distribution. Several attempts have been made at developing suitable computational methods to derive a distribution from a correlation function. Perhaps the most widely known and used is the constrained regularisation programme CONTIN [30,31]. In many cases the programme works well, but care has to be taken in choosing the right range of D to explore for a solution, and the original data must be of high quality, as 'noisy' data can lead to artefacts in the analysis. A comparison of CONTIN with maximum entropy methods has recently been published [32].
Relative Contribution
Diffusion Coefficient (cm 2 s'x)
x1
°
Figure 12.12 Distribution in diffusion coefficients for an aromatic terpolyester in a mixed solvent of trifluoroacetic acid and dichloromethane obtained by CONTIN analysis of quasi-elastic light scattering data Obtaining molecular weight distributions by this means has two benefits. Firstly, with a high power laser light source on the correlator and with fast data links to a work station, a full molecular weight distribution can be obtained in «2min. The second benefit is when only ferocious solvents are available, ones which would destroy size-exclusion chromatography (SEC) column packings; quasi-elastic light scattering then becomes a highly suitable method to obtain a molecular weight distribution. An example of this is the aromatic terpolyester prepared from hydroxybenzoic acid, isophthalic acid and hydroquinone [33], which is soluble in a mixture of trifluoroacetic acid and methylene chloride. The low refractive index of the solvents and the high refractive index of the polymer make the solutions extremely strong scatterers of light and ideal for CONTIN analysis, even with only a modest laser. An example of the distribution in diffusion coefficients (and hence molecular weight) is shown in Figure 12.12.
12.3.2 GELS A cross-linked polymer swollen by a solvent constitutes a gel, and if swollen sufficiently the concentration of polymer in the gel is that of a semi-dilute
solution, i.e. it is between c* and c** as defined by de Gennes. The gel has continual local fluctuations in the degree of swelling (equivalent to polymer concentration) which lead to variations in the local osmotic pressure. The analysis of the intensity correlation function obtained from the scattering of light by these fluctuations produces a co-operative diffusion coefficient. The first QELS experiments on gels and the theoretical analysis of the data were reported over 20 years ago by Tanaka et al. [34]. They showed that at a delay time of zero (i.e. extrapolating the correlation functiion to t = 0), the scattered light intensity above the background was equal to the osmotic moldulus Mos (= Kos + 4Gos/3 where Kos is the bulk osmotic modulus and Gos is the shear osmotic modulus), also known as the longitudinal modulus. The co-operative diffusion coefficient is given by DC = (KOS + 4GOS/3)(1 -
and Dc o n
Intensity (a.u.)
T = 308K Solvent C6H12
Baseline
Time (jus) Figure 12.13 Intensity autocorrelation function obtained for a randomly cross-linked network of polystyrene swollen in cyclohexane at 308 K was increased. Increased non-exponential behaviour has been identified with the overlapping of molecules and appears to be possible only when there are many loose dangling chains.
12.3.3 SEMI-DILUTE S O L U T I O N S A N D TRAPPED CHAINS The broad outlines of reptation theory are well known, and the detailed theory is available elsewhere [43,44]. Essentially, a polymer molecule in a melt is confined to a tube which is defined by the surrounding molecules, and can only move along the tube axis. The time dependence of the various dynamic modes of the molecule in the tube has been discussed by Doi and Edwards [45]. Additionally, de Gennes [46] has set out equations which relate the translational diffusion coefficient of a probe polymer to its molecular weight (Mp), the entanglement molecular weight of the matrix (MJ and the molecular weight between cross-links (AfJ. Three regimes are predicted: 1. Free draining (A/p < Afc, Afp > AfJ, D = D0M; K
DJm2S'1)
Figure 12.14 Co-operative diffusion coefficient as a function of volume fraction of polymer in cyclohexane swollen polystyrene networks; (o) 308 K, (o) 318 K, (•) 333 K 2. Simple reptation (M p > Me, M c > Mc), D = D0M tM; 2. 3. 'Strangulation' regime (Me > Mc, M p > Mc), Dt = D0M0M;
2
.
Attempts have been made at observing these regimes using semi-dilute solutions of a matrix polymer with a chemically identical probe of a different molecular weight incorporated in the solution. The conclusion of these experiments was that the reptation theory was inappropriate for such semi-dilute solutions [47,48]. A possible explanation for the failure of reptation theory may be in the recent analysis of Wang [49-51]. He shows that the quasi-elastic light scattering from a semi-dilute solution has contributions from both concentration fluctuations and density (pressure) fluctuations, and consequently the long time viscoelastic relaxation spectrum, usually observed by dynamic mechanical means, will also contribute to the autocorrelation function. The extent to which both contributions are seen depends on the frequency distribution of the stress relaxation modulus and a coupling parameter j8 (proportional to the partial
log[D t <M x >/M x ]
log M Figure 12.15 Diffusion coefficient of polystyrene tracer in polyvinyl methyl ether gels as a function of tracer molecular weight. Diffusion coefficients normalised by ratio of molecular weight between crosslinks of gels. Reprinted with permission from [52]. Copyright 1992 American Chemical Society
specific volume of the polymer minus the partial specific volume of the solvent). Very recently, QELS investigation of reptation predictions has been made using randomly cross-linked networks containing chemically distinct trapped chains. Rotstein and Lodge [52] prepared polyvinyl methyl ether gels containing trapped polystyrene chains, and obtain tracer diffusion coefficients for the toluene-swollen gels. Values of M c were calculated from swelling data, and 4 x 103 ^ Mc ^ 14 x 103. Figure 12.15 shows the diffusion coefficient data normalised by the ratio of the M c values for the three networks involved. There appears to be little or no influence of Mc even when M p » Me; furthermore, the probe molecular weight dependence of D (DocM~2S) is much stronger than predicted by reptation theory. Pajevik et al. [53] prepared randomly cross-linked polymethyl meth'acrylate gels containing polystyrene probe molecules. Their results are shown in Figure 12.16. When M p < Mc («80000) then D scales as Mp ° 6; above this molecular weight the influence of M p is marked and D scales as M~ l'*±°-29 i.e. almost exactly in agreement with reptation theories, CONTIN or an equivalent program was used in both investigations, and the isorefractivity of toluene with polyvinyl methyl ether and polymethyl methacrylate aids the
D*/D0
Mp
Figure 12.16 Ratio of polystyrene tracer diffusion coefficient (D1) in toluene swollen PMMA gel to diffusion coefficient of polystyrene in dilute toluene solution (•); (A) values for PS tracer in PMMA solutions. Reproduced with permission from the American Chemical Society from Ref. [53] process of extracting the probe diffusion coefficient. However, about 14 years [54] ago it was noted that, when polystyrene was dissolved in a semi-dilute benzene solution of polymethyl methacrylate, the value of D decreased as the polymethyl methacrylate concentration increased, i.e. rather similar to the molecular weight dependence seen by Pajevik et al, and this may be due to polymer-polymer interactions. To overcome these possible complications, polystyrene networks with trapped polystyrene molecules have been prepared [55] and are currently being investigated.
12.3.4 SURFACE QUASI-ELASTIC LIGHT SCATTERING (SQELS) A liquid surface is continually roughened by thermal excitations, which give rise to the hydrodynamic modes known as capillary waves. The r.m.s. amplitudes of the waves are small ( « 2A) but they are efficient light scatterers. The displacement of the liquid surface from its equilibrium position by a wave propagating in the x direction is: C(x,r) = C 0 exp(/ex-ho)0 where Q is the surface wavenumber or the scattering vector parallel to the liquid surface. The wave frequency o is a complex quantity given by a>0 + iT, where co0 is the capillary wave frequency and F is the decay rate of the waves. A dispersion equation relates co and Q, and for pure liquids the controlling factors (for fixed Q)
are the kinematic viscosity and the surface tension [56]. For most instruments the accessible range of Q is 100-2000Cm"1 and hence the wavelengths probed are « 600-30 /mi. If a polymer film is spread on the surface of the liquid, additional hydrodynamic modes modify the dispersion equation. Only the transverse modes (capillary waves) scatter light, but there is coupling with the longitudinal or dilational modes, and hence in principle some information is obtainable on both modes from the power spectrum of the scattered light. The parameters obtainable are the surface tension y and the dilational modulus e; both of these are viscoelastic properties, as energy dissipation takes place in the relaxation processes, and thus y = y0 + icoy' e = 6 0 + icoe'
where y0 and £0 are the static surface tension and dilational modulus I — I, \ A aA J y' is the transverse shear viscosity and e' is the in-plane dilational viscosity. Although direct measurement of the frequency broadening of the scattered light by the capillary waves has been used, the frequency shifts are rather small, and a more direct means of observing the frequency of the capillary waves is to use heterodyne quasi-elastic light scattering [57,58]. The experimental arrangement to collect such data is shown in Figure 12.17; the diffracted beams produced
Laser
rough
PM Tube
Figure 12.17 Schematic diagram of surface quasi-elastic light scattering apparatus. Ll, L2 = lenses, T = transmission grating, F = neutral density filter, Ml, M2, M3, M4 = mirrors
Normalised correlation function
Time (us)
Figure 12.18 Heterodyne correlation function for syndiotactic polymethyl methacrylate spread on water at a surface concentration of 1.7mgm~2 by the transmission grating act as the reference beam of zero frequency shift, and this beats with the scattered light at the photocathode to produce the typical correlation function shown in Figure 12.18. From these data the capillary wave frequency co and the decay constant F can be obtained. By assuming that y and e! are zero, y0 and e0 can be obtained from these values by solving the dispersion equation. Extracting the viscous moduli requires a non-linear least squares fit of the Fourier transform of the power spectrum equation to the data. A computational method for this process has been developed by Earnshaw et al. [59] and exhaustively justified [60]. Wider aspects of light scattering from liquid surfaces are discussed in the book edited by Langevin [61]. To date much of the work published on SQELS from spread polymers has emanated from Yu and colleagues [62-65], but assumed that the viscous moduli are zero. We have reported [66] a limited study of spread polymethyl methacrylates and polyethylene oxide. Figure 12.19 shows the variation in surface tension, shear viscosity and dilational modulus obtained from SQELS data as a function of surface concentration. The viscoelastic moduli both show maximum values at finite values of the surface concentration. As the capillary waves generate oscillatory stress and strain, these are related via the complex dynamic modulus of the surface a* =y*[G'(co) +iG"(co)]
Surface tension (mN nrr1) Shear viscosity (mN s rrr1)
Surface concentration (mg nrr2)
Surface concentration (mg m*2) Figure 12.19
(Continued)
Dilational modulus (mN nrr1)
Surface concentration (mg nrr2)
Figure 12.19 Derived parameters from surface quasi-elastic light scattering as a function of concentration of polymethyl methacrylate spread on water: (a) surface tension; (b) surface shear viscosity; (c) dilational modulus
where <x* is stress, y* is strain, G'(co) is the storage modulus (surface tension) and G"(a>) is the loss modulus (a>yf). Using volume fraction composition data obtained from neutron reflectometry on the spread polymer films, it is evident that the surface film loss modulus is linearly dependent on the volume fraction of polymer in the film. If we presume that the relaxation process in the surface film is described by a Maxwell model, then G'(co) = Ge + GCO2T2/(1 + CO2T2)
where Ge is the elastic modulus at co = O, i.e. the static surface tension. Further, if there is only one relaxation process in the spread film, then T = A7c/co2y' where An is the difference in the surface tensions measured by SQELS and from static (Wilhelmy) plate methods. The dependence of relaxation time on the volume fraction of the polymer shows an exponential increase, Figure 12.20. To obtain further insight into the relaxation mechanism requires the frequency dependence (i.e. different Q values) of the transverse shear viscosity to be known.
Relaxation time (s)
SYN PMMA SQELS DATA
VoI fraction of polymer
Figure 12.20 Relaxation time for spread polymethyl methacrylate as a function of volume fraction of polymer in the spread film
12.4 CONCLUSIONS An overview of some of the areas where light scattering has made contributions to polymer science has been given. The emphasis has been on dynamics, either by using light to follow a process (crystallisation or phase separation) or using dynamic light scattering per se. A broad range of polymer types and situations has been covered and the discussion has by no means been exhaustive. Evidently, despite its maturity as a laboratory technique, light scattering is still capable of providing much information on polymer systems. Furthermore, the development of newer applications such as surface quasi-elastic light scattering will enable investigations of surface gelation and surface ordering in polymer solutions, areas which have yet to be investigated.
12.5 REFERENCES [1] M.B. Huglin (Ed.), Light Scattering from Dilute Polymer Solutions, Academic Press, London, 1972. [2] P. Kratochvil, Classical Light Scattering from Polymer Solutions, Elsevier, Amsterdam, 1987.
[3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44]
H. Yamakawa, Modern Theory of Polymer Solutions, Harper, New York, 1971. GC. Berry, J. Polym. ScL, Polym. Symp., 1978,65,143. W.R. Krigbaum and G. Brelsford, Macromolecules, 1988, 21, 2502. Z. Tuzar, P. Kratochvil and D. Strakova, Eur. Polym. J., 1970,6,1113. BJ. Berne and R. Pecora, Dynamic Light Scattering, Wiley, New York, 1976. K.S. Schmitz, An Introduction to Dynamic Light Scattering by Macromolecules, Academic Press, San Diego, 1990. R.S. Stein and JJ. Keane, J. Polym. ScL, 1955,17, 21. R.S. Stein and M.B. Rhodes, J. Appl. Phys., 1960, 31,1873. R.S. Stein, P. Erhardt, JJ. van Aartsen, S. Clough and M. Rhodes, J. Polym. Sci. C, 1965,13,1. RJ. Samuels, J. Polym. Sci. C, 1965,13, 37. G.E. Wissler and B. Crist, J. Polym. Sci., Polym. Phys. Ed., 1985, 23, 2395. M. Ree, T. Kyu and R.S. Stein, J. Polym. ScL, Polym. Phys., 1987, 25,105. J.V.Champion,A.KilleyandG.H.Meeten,J.Polym.ScL,Polym.Phys.Ed., 1985,23, 1467. G.H. Meeten and P. Navard, J. Polym. ScL, Polym. Phys., 1989, 27, 2023. M. Desbordes, G.H. Meeten and P. Navard, J. Polym. ScL, Polym. Phys., 1989, 27, 2037. P.H. Richardson and R.W. Richards, unpublished work. WT. Culberson and M.R. Tant, J. Appl. Polym. ScL, 1993,47, 395. P.H. Richardson, ubpublished results. J.C. Schultz, Polymer Materials Science, Prentice-Hall, New Jersey, 1974. J.W. Cahn and J.E. Hilliard, J. Chem. Phys., 1958,28, 258. J.G. Connell, Ph.D. Thesis, University of Strathclyde, 1989. H.L. Snyder and P. Meakin, Macromolecules, 1983,16, 757. T. Hashimoto, M. Itakura and N. Shimidzu, J. Chem. Phys., 1986,85,6773. A. Cumming, P. Wiltzius, F.S. Bates and J.H. Rosedale, Phys. Rev. A, 1992, 45, 885. D.E. Koppel, J. Chem. Phys., 1972,57,4814. P.N. Pusey, D.E. Koppel, D.W. Schaefer, R.D. Camerini Otero and S.H. Koenig, Biochemistry, 1974,13, 952. W. Burchard, M. Schmidt and W.H. Stockmayer, Macromolecules, 1980,13,1265. S.W. Provencher, Comput. Phys. Commun., 1982,27, 213. S.W. Provencher, in E.O. Schulz-DuBois (Ed.), Photon Correlation Techniques in Fluid Mechanics, Springer, Berlin, 1983. S.W. Provencher, in S.E. Harding, D.B. Sattele and V.A. Bloomfield (Eds.), Laser Light Scattering in Biochemistry, Royal Society of Chemistry, Cambridge, 1992. A.D.W. McLenaghan, Ph.D. Thesis, University of Strathclyde, 1990. T. Tanaka, L.O. Hocker and G.B. Benedek, J. Chem. Phys., 1973,59, 5151. E. Gleissler and A.M. Hecht, J. Phys. (Paris) Lett., 1979,40, L173. A.M. Hecht and E. Geissler, J. Phys. (Paris), 1978, 39, 631. A.M. Hecht, E. Geissler and A. Chosson, Polymer, 1981,22, 877. E. Geissler and A.M. Hecht, J. Chem. Phys., 1982, 77,1548. EJ. Amis, P.A. Janney, J.D. Fery and H. Yu, Macromolecules, 1983,16,441. N.S. Davidson, Ph.D. Thesis, University of Strathclyde, 1984. N.S. Davidson, R.W. Richards and E. Geissler, Polymer, 1985, 26,1643. T.G. Scholte, J. Polym. ScL A2, 1970,8, 841. W.W. Merrill and M. Tirrell, in G.R. Freeman (Ed.), Kinetics of Nonhomogeneous Processes, Wiley, New York, 1987. T.P. Lodge, N.A. Rotstein and S. Prager, Adv. Chem. Phys., 1990,79,1.
[45] M. Doi and S.F. Edwards, The Theory of Polymer Dynamics, Oxford University Press, Oxford, 1986. [46] P.G. de Gennes, Macromolecules, 1986,19,1245. [47] W. Brown and P. Zhou, Macromolecules, 1989, 22, 3508. [48] T. Nicolai, W. Brown, S. Hvidt and K. Heller, Macromolecules, 1990,23, 5088. [49] CH. Wang, J. Chem. Phys., 1991,95, 3788. [50] CH. Wang, Macromolecules, 1992, 25, 1524. [51] CH. Wang and X.Q. Zhang, Macromolecules, 1993,26, 707. [52] N.A. Rotstein and T.P. Lodge, Macromolecules, 1992, 25,1316. [53] S. Pajevic, R. Bansil and C Konak, Macromolecules, 1993, 26, 305. [54] AJ. Hyde, J. Hadgraft and R.W. Richards, J. Chem. Soc. Faraday Trans II, 1979,75, 1495. [55] D.A. Davison, University of Durham, work in progress. [56] J.C Earnshaw and R.C McGivera, J. Phys. D, 1987,20, 82. [57] S. Hard and R.D. Neuman, J. Colloid Interface ScL, 1981,83, 315. [58] J.C. Earnshaw, in CA. Croxton (Ed.), Fluid Interfacial Phenomena, Wiley, New York, 1986. [59] J.C Earnshaw, R.C McGivern, A.C McLaughlin and P. J. Winch, Langmuir, 1990, 649. [60] J.C. Earnshaw and R.C. McGivern, J. Colloid Interface ScL, 1988,123, 36. [61] D. Langevin (Ed.), Light Scattering by Liquid Surfaces and Complementary Techniques, Dekker, Basel, 1992. [62] M. Kawaguchi, M. Sano, Y.-L. Chen, G. Zografi and H. Yu, Macromolecules, 1986, 19, 2606. [63] M. Kawaguchi, B.B. Sauer and H. Yu, Macromolecules, 1989, 22,1735. [64] B.B. Sauer, M. Kawaguchi and H. Yu, Macromolecules, 1987, 20, 2732. [65] K.-H. Yoo and H. Yu, Macromolecules, 1989, 22,1989. [66] J.A. Henderson, R.W. Richards, J. Penfold and R.K. Thomas, Macromolecules, 1993, 26,65.
13 NEUTRONSCATTERING FROM POLYMERS A. R. RENNIE Polymers and Colloids Group, Cavendish Laboratory, University of Cambridge, Madingley Road, Cambridge, CB3 OHE, UK
13.1 INTRODUCTION In the space of a few pages it would be impossible to provide a full description of the different investigations of polymers that can be made, or even have already been made, using neutron techniques. The intention is rather to provide an introduction to the methods. Those readers who may wish to exploit the special advantages of neutrons will be able to assess the feasibility of experiments and find some guide to the recent literature. The description of published work will necessarily be selective and will try to reflect some of the wide range of studies that are now in progress. There are several reviews of both the technique of neutron scattering and its application to the study of polymers. Although both provide little information about work on polymers, readers interested in a thorough description of the theory of neutron scattering should refer to the books by Lovesey [1] and Squires [2]. Reviews and books concerning neutron studies of polymers can be divided into those that are concerned with the technique [e.g. 3-5] and those that describe results, often in particular areas of the topic such as copolymers [6], networks [7], polymer motion [8-11], semi-crystalline polymers [12], polymer colloids [13] and biopolymers [14].
13.2 THE PRINCIPLES OF NEUTRON SCATTERING The scattering of neutrons can be treated formally as an inversion problem: the scattered intensity from a plane incident beam of wavelength A into a solid angle dQ, in the energy range d£, at a scattering angle 6 can be expressed as the Fourier transform in space and time of the pair correlation function of scattering length density. In this respect the theory of neutron scattering is identical to that of weak scattering of any other form of radiation. The advantages of neutron scattering Polymer Spectroscopy. Edited by Allan H. Fawcett © 19% John Wiley & Sons Ltd
arise from the magnitudes of the quantities mentioned above and their interrelationships. For example, the large mass of the neutron when compared with electrons or photons is important in providing good coupling between molecular motion and the energy of the scattered beam. In the next few paragraphs some of the formalism associated with the description of scattering will be presented. Further details of the principles of scattering and the properties of neutrons will be found in books on atomic physics such as that by Born [15]. Many people will not be concerned with such fundamentals of the theory, and the interpretation of many experimental results can be adequately performed using the simple relationships that derive from appropriate integrations of the wave equations. Some of these are described in the next section. It is first useful to recall that the neutrons can be considered as either particles or waves; the connection between the two descriptions is provided by the de Broglie relationships: E = hv p = hk/2n
(1) (2)
where E is the energy, p is the momentum (product of mass and velocity v), v is the frequency and k is the wave vector of the neutron. The magnitude of the wave vector is given by |k| = 2n/h where X is the wavelength. A schematic diagram of a general scattering experiment is shown in Figure 13.1. We can distinguish two general cases. First, the situation in which the energy and wavelength of the scattered neutron are equal to those of the incident beam. This is known as elastic scattering, and gives the simple result that the momentum transfer | Ql is (An/X) sin (0/2). More generally, there will be some energy transfer between the neutron and the sample. The experiment is then said to involve inelastic scattering or, if the energy transfer is small and corresponds only to a broadening of the incident wavelength distribution, quasi-elastic scattering. Advantages of neutrons over other types of radiation for studies of polymers arise both from the relationship of energy to wavelength and from the mechanism of scattering by nuclei. The calculation of scattering patterns is based on the summation of amplitudes or intensities scattered from all components. The intensity I of a wave described in the usual notation of complex variables is given by I = AA*
(3)
where A is the amplitude and the star represents a complex conjugate. The addition of wave amplitudes from different scattering centres must take account of the phase of each wave if the incident beam is a coherent wave front. Coherence in the context of neutrons will be discussed further below. The amplitude A scattered by a single nucleus at a position r from an incident plane wave of amplitude A0 is given by:
A = Aobc-^/\r\
(4)
Sample
Detector
Sample
Figure 13.1 (a) Schematic diagram of a scattering experiment; (b) the wave vectors that describe the scattering process. Q is the difference between the incident and scattered wave vectors
This is known as the Born approximation for scattering from weak potentials. It is seen that the scattered wave is spherically symmetric and decreases in intensity as l/|r| 2 , which is the usual inverse square law. For a distribution of scattering lengths described by a density p(r) the resulting amplitude is
A = A0^tp(r)e-^/\rndr3
(5)
and thus the scattered intensity / is / = AA* = Al f [p(r)e- fQr p(r')e^7|r| 2 ]d(r - r')3
(6)
The quantity p{r)p{r') is the spatial correlation function of the scattering length density, and the integral with e" lQ(r " r) is equivalent to a Fourier transform. It would be out of place to develop this formalism at great length. The results for elastic scattering are well described by Hukins [16].
It is possible to include the angular frequency co or energy E of the neutrons and the time variation of the scattering length density in Equation (4) to device related results for inelastic or quasi-elastic scattering. A more formal treatment would first treat this general case and then simplify the results for elastic scattering. These will not be derived but it is sufficient to quote the result:
dL/dodQ = I/Al
= I Mr,t)e-i(Qr+cor)p(^tV(Qr'+
- r'fdt
(7)
Details of the theory of inelastic scattering can be found in the books of Lovesey [1] and Squires [2]. The intensity or the differential scattering cross-section dZ/dcodfi can be calculated for various models of structure and dynamic behaviour using Equation (7). The models of particular importance to the study of polymers are similar to those relevant to quasi-elastic light scattering [17,18]. In practice, the calculation of such integrals is often simplified by recognising that they are Fourier transforms and that several theorems are available to describe their properties, and are included in textbooks on mathematical methods [20,21]. Very few polymers are completely crystalline, and so the present description differs from that of most scattering experiments, which are concerned with diffraction or inelastic scattering from a regular lattice. The problems of predicting scattering patterns can generally be reduced to integrals, results for many of which can be derived readily or calculated numerically. For the case of structural studies by elastic scattering, many of these results are available in the literature [22-24], even for amorphous or random structures. Most were originally derived for X-ray or light scattering. An important feature of neutron scattering is the neutron scattering length, designated b9 which describes the probability of scattering. Unlike photons or electrons the neutron is scattered by the atomic nucleus. Some values for elements commonly found in synthetic polymers are listed in Table 13.1. The scattering lengths are not correlated with atomic number and can change between isotopes of a single element. This allows isotopic labelling of individual molecules or parts of molecules without perturbing the chemistry, and has proved of particular value in the study of polymers, which often consist largely of atoms such as carbon, hydrogen, oxygen and nitrogen with low electron densities and thus poor contrast for other radiation. Of particular significance is the ability to label individual molecules among other chemically identical polymers to determine their size or investigate their mobility in the bulk. Most readers will be familiar with the idea that structural studies of materials carried out by light scattering can be performed only within the limits of coherence of the source. This is usually explained in terms of the finite time over which a wave is emitted from a source. Similar considerations apply to neutrons;
Table 13.1 Neutron scattering cross-sections [19] Element Hydrogen 1 H Deuterium 2 H Carbon Nitrogen Oxygen
Coherent scattering lengh/fm -3.74 6.67 6.65 9.21 5.9
Incoherent scattering cross-section/10"28 m 2 79^9 2.04 0.0014 0.49 0.015
there is, however, a further constraint on the condition for coherent addition of amplitudes. A neutron possesses the property known as spin, and must retain the same spin state for coherent scattering. The nuclei of atoms can also have spin, and will in general have two separate probabilities of scattering associated with the preservation and loss of spin coherence. It is usual to quote the coherent scattering length (often written as b) in units of length and the incoherent scattering as a cross-section
13.3 NEUTRON EXPERIMENTS There are many classes of experiment that can be made with neutrons; they cover a wide range of angles, wavelengths and energy transfer and it is not possible to give a full description of them all here. Some experiments are frequently used with polymers, and these will be outlined briefly in three categories after a few general remarks on neutron instrumentation. The description of scattering theory given above has indicated that the important variable is the momentum transfer Q rather than the wavelength or angle of scattering. It is possible to measure over the necessary range of Q values in two distinct ways. The first is to use a fixed wavelength and scan the angle. The second is to use a fixed angle and a range of wavelengths. This second technique can be exploited with advantage at pulsed neutron sources: the source provides a pulse consisting of many wavelengths which can be analysed using measurements of the time-of-flight of neutrons from the source to the detector. This
Table 13.2 Benefits of neutron scattering Advantages Disadvantages Contrast with isotopes Low flux Contrast variation Large, expensive source Measurements in bulk Easy control of sample environment Measurements of molecular motion Can be surface specific approach can exploit a wide range of wavelengths available in a pulse, as it avoids the need to provide a monochromatic beam. On continuous sources it is usual to use a monochromator, either crystals used for Bragg diffraction or mechanical velocity selection, and measurement is made at a range of angles. These two techniques are sometimes associated with spallation sources, which are often pulsed, and reactors, which are usually run as continuous sources. Exceptions to both these rules occur, in that pulsed reactors have been built and, even more simply, the continuous beam from a reactor can be pulsed using a rotating chopper. Continuous spallation sources can also be built. The details of design of both neutron sources and the diffractometers and spectrometers associated with them are to be found in the specialist literature [25-27]. A property common to all neutron sources is that of low flux or brightness. Even the best design of nuclear reactors cannot provide a flux of thermal neutrons in the moderator close to the core larger than a few times 10 15 n cm ~2 s ~ *. After selection of wavelength (monochromation) and collimation to define the incident beam, the flux available at the sample on a diffractometer or spectrometer is unlikely to be more than 108n Cm -2 S" 1 and is often much less. This is lower than the flux of photons available from typical laboratory sources of light or X-rays. A consequence of this low flux is the need to use large samples (crosssectional areas of % 1 cm2) and detectors of high efficiency covering a large solid angle around the sample. The combination of large samples and the need to obtain reasonable angular resolution on the detectors will usually lead to physically large instruments. An extreme example is the overall length of 80 m for one small angle scattering instrument known as Dl 1, which has been built at the Institut Laue-Langevin in Grenoble, France [28]. Characteristics of neutron scattering for comparison with other radiation are shown in Table 13.2.
13.3.1 STUDIES O F POLYMER DIMENSIONS: SMALL ANGLE SCATTERING The most frequently encountered use of neutrons in the study of polymers is the measurement of small angle scattering (SANS). Use of neutrons with wavelengths in the range 0.5-2 nm and scattering angles between 0.1° and 10° readily allows
investigation of structures in the size range l-200nm. This class of study can be used to give information about the size and conformation of individual polymer molecules. These may be either in solution or in the bulk. The technique is also used for the study of phase separation in polymer blends or copolymers and in the study of structures of composite materials. The theory of small angle scattering has several simplifications over the more general ideas outlined above. It is concerned only with elastic scattering, although in practice the total intensity measured includes both elastic and inelastic scattering [29, 30], which can give rise to some additional complications to measurements made with time-of-flight instruments. In the most straightforward case, the scattering from isolated objects is measured to determine their size and shape. The integral in Equation (6) is thus performed over a single object such as a polymer molecule in solution. There is no correlation between the positions of nuclei in different objects, and so the total intensity from a solution is found by adding the intensity from each object. Several simple results can be derived for the scattering at small Q. It is useful to write down the result of integrating Equation (6) in the following simplified manner /(Q) = CP(Q) (8) where P(Q) is a function describing the shape of the scattering normalised to unity at Q = 0. The constant C depends on the contrast (square of scattering length density difference), the number density of particles and the incident flux. A general result due to Guinier [23] states that for any object with spherical symmetry, in the limit of small Q9 P(Q) will be approximated by />(G) = exp(-e 2 R g 2 /3) (9) where Rg is the radius of gyration of the object. This is clearly of considerable value in determining the dimensions of polymer molecules, even in the absence of detailed models of the structure. It is trivial to show by way of series expansions that in this condition of QRg < 1 the result can also be written as 1/P(Q)=I-Q2
R2g/3
(10)
which is the result of Zimm [31] frequently encountered in light scattering. It is of course possible to extend the analysis to include virial interactions between molecules in the manner described in light scattering texts [32, 33]. In many circumstances it is possible to extend the range of measurements with neutrons to cover more than this low Q limit, and then more detailed models of the structure must be evaluated. Debye [22] has derived a result for the scattering from a Gaussian distribution of polymer segments appropriate to a random polymer coil which is of the form: P(Q) = (2/Q2Rl)lexp(-Q2R2g)-(l-Q2R2K (11) and can be used to fit data over a much wider range of momentum transfer until the Gaussian approximation for molecular structure fails at short distances.
The constant C describing the absolute intensity is of importance as it permits determination of the molecular weight of polymers. By rearrangement of the constants in Equation (S), it can be expressed as: I/C = ( N A / c M w ) p > p - p s ) 2
(12)
where NA is Avogadro's number, c is the concentration, Mw is the molecular mass, pm is the mass density of the polymer, and p p and p s are the scattering length densities of the polymer and solvent respectively [14, 34]. Measurements of the absolute intensity of scattering at low Q, which can be extrapolated to Q = 0 or fitted using one of the equations above, can thus give information about the weight average molecular mass. The real value of small angle neutron scattering lies in the realisation that the simple theory above, which is essentially identical to that of light scattering, can be extended in two ways. The possibility of isotopic substitution and contrast between chemically identical molecules can lead to measurements of molecular dimensions in the bulk rather than in solution. Some of the early work with SANS or polymers was concerned with the verification of the idea that screening of molecular interactions in the melt gave rise to molecular dimensions that were identical to those found in theta solvents [35-37]. This work has now been extended to a wide range of investigations of molecular conformation in bulk polymers, which can include amorphous glasses [38], melts [39], gels [40-42], elastomers [43-46] and semi-crystalline polymers [47-51]. A major boost to the application of SANS to polymers in the bulk came from the recognition that the screening of molecular interactions in bulk homopolymers could be used to extend the range of concentrations over which measurements can be made. This idea, which is known as the random phase approximation or RPA [52, 53], states that if there are no interactions, i.e. the second virial coefficient is zero, then measurements of molecular dimensions can be made at any concentration. In order to optimise count rates this may often be close to 50% blends of deuterated and protonated polymers. Several experiments have been performed to test this theory [54, 55], which is now widely applied. It should be remembered that there are many cases where the measurement of interactions is of importance. SANS has been widely used for the study of polymer blends. A simple extension of the theory gives the following expression for the scattering from a blend of two miscible polymers with a Flory-Huggins interaction parameter x'. 1/7(0 - 1 / P 1 ( Q ) + 1/P2(Q) -2X
(13)
where P 1 and P 2 are the Debye expressions for separate polymers as given by Equation (11). Other extensions of theory can be made to describe the scattering from copolymers [56,57], branched and star-like polymers [58] and also variety of geometrical shapes that may be appropriate to describe liquid crystalline and semi-crystalline structures [23, 24]. Local structure in polymers such as that
arising from chain rigidity has also been described [59]. It should also be mentioned that the scattering from polymers bound to the interface of colloidal particles has been the subject of several investigations with SANS. 13.3.2 POLYMERS AT S U R F A C E - R E F L E C T I O N Neutrons can be reflected from planar surfaces according to the usual laws of specular optics. The refractive index n for neutrons is given by: n=l-(p№t)
(14)
This equation indicates that for most materials the refractive index will be very close to, but generally slightly less than, unity. The condition in optics known as total internal reflection will thus be replaced by total external reflection [60]. This will occur at low angles, typically about one degree for thermal neutrons. At angles larger than the critical angle, the reflectivity is reduced, and it is this variation that provides information about the structure of the interface. The intensity of a neutron beam reflected from a surface can be calculated exactly using the optical matrix approach of approximating the interface to a series of thin layers and calculating the reflection at each boundary in the manner described by Born and Wolf [61], Heavens [62] or Abeles [63]. It is perhaps more instructive to consider an alternative procedure based on the kinematic theory of scattering [64] which gives the following approximate expression for the reflectivity R(Q) as a function of Q the momentum transfer normal to the interface: R(Q)= I6n2\ H(Q)2 \Q2 (15) where H(Q) is the Fourier transform of the scattering length density distribution normal to the interface p(z). This result is not valid close to critical reflection,but can now be extended [65] to provide analytical forms over the entire range of Q. The experimental arrangements for such measurements are shown schematically in Figure 13.2(a). This technique has emerged only recently but is rapidly growing. Reviews of the experimental technique [66] and the application to polymers [67] have already appeared. The application of neutrons to the study of polymer surfaces or interfaces can be divided into two categories. First, the study of polymers adsorbed or spread at liquid interfaces; secondly, the study of polymers in thin films. This second category provides interesting models for measurements on polymer compatibility and inter-diffusion. Reports of ordering of copolymers at surfaces [68], phase separation [69] and the width of polymer/polymer interfaces during diffusion [70] have been made. The results on adsorbed polymers using neutron reflection provide a useful complement to studies Qf adsorption by small angle scattering and other classical techniques. Measurements of the excess polymer at both solution/vapour [ 7 1 73] and solution/solid [74-76] interfaces have been described. Other studies
Detector
Neutrons
Sample
Solid
Solution Figure 13.2 (a) Neutron reflection experiment; the geometry is arranged so that only specular reflection (angle of incidence is equal to angle of reflection) is observed; (b) observations can be made either at the interface of a sample with vapour (i) or at a solution/solid interface (ii) provided that a material of adequate transparency to neutrons such as silicon or quartz is used to form the bulk of the solid have been made of the structure of polymer monolayers spread at the air/water interface on Langmuir troughs [ 7 7 - 7 9 ] .
13.3.3 POLYMER DYNAMICS—QUASI-ELASTIC SCATTERING The advantage of neutron scattering for studies of polymer dynamics is the direct information that is provided about the molecular correlation functions in time and space by the technique. Other spectroscopic probes such as NMR and dielectric or mechanical response can provide information about the time or frequency of relaxations, but do not directly provide information about the length scale on which the dynamic processes occur. Once again there is an analogy to dynamic light scattering, and most of the theory is similar. The derivation of the scattering laws describing the differential scattering cross-
precession
Figure 133 A schematic diagram of a neutron spin-echo spectrometer. The difference in velocities of the polarised neutron before and after the scattering process can be observed by measuring the precession in the regions of uniform magnetic field H. The difference in the precession is easily determined from the polarisation of the beam reaching the detector D if the fields before and after the sample S are symmetrical and the polarization is inverted with the flipper coil marked as U/2. The coils marked n/4 are used to provide magnetic fields that define the initial and final states of polarisation section for various models of polymeric motion, such as reptation of molecules through entanglements or the hydrodynamic screening described by Zimm, is complex. The main results can be found in the book by Doi and Edwards [80]. For our purposes it is sufficient to recognise that a diffusive process gives rise to a scattering law of the form: /(Q, Q) = DQ2/{Q2 + D2Q2)
(16)
where D is the diffusion coefficient and Q and Q refer to the momentum and the energy transfer in the scattering process. If measurements are made directly in the time domain, then the scattering law will ocrrespond to the Fourier transform of this Lorentzian, which is simply an exponential decay I{Q,t) = exp(-DQ2t) (17) Measurements of the diffusion of polymers in the melt and in solution have been made using neutrons. It is of particular interest that the distance scale over which diffusion is measured depends inversely on Q9 and so it is possible to probe both the regions of internal modes and overall intermolecular diffusion. The technique that has attracted most recent attention is that of neutron spin-echo (NSE) spectrometry. This method of measuring small energy transfers, which was proposed by Mezei [81, 82], cleverly avoids the need to monochromate the beam precisely and thus provides gains in the incident flux. The experiment (see Figure 13.3) measures the energy change of each neutron by observation of the precession of the spin in uniform magnetic fields. Experiments using the NSE technique have been used to verify the microscopic aspects of Rouse and Zimm models of the motion in polymer solutions [83]. They have also shown that the reptation process [84] of self-diffusion along a tube
can be well described by Rouse modes within the tube. The effects of entanglements (topological constraints) can be observed in the long distance (small Q) modes in melts and networks [85-88]. More recently, experiments have looked at the motion of copolymers [89] and of polymers close to the glass transition [90-92].
13.4 SOME EXAMPLES OF RECENT PROGRESS Here a few examples of recent work will be mentioned. The selection naturally reflects the author's interests, but is intended to illustrate those areas of neutron scattering that have particular prominence at present or are growing rapidly. Many of the early experiments with neutrons were concerned with the properties of homopolymers; indeed, physicists were often seeking the simplest systems that could be considered as uniform, flexible macromolecules to test fundamental models of polymer conformation and dynamics. Recent work has been characterised by an increasing complexity in the systems that are investigated to include copolymers, blends and composite materials. In some cases the neutron studies are a minor part of more widespread investigations of the properties of novel polymers, their synthesis or processing. The examples below are intended to provide some insight into the range of studies that can be made and the precision of data that can be obtained. 13.4.1 STUDIES O F COPOLYMERS The study of copolymers with scattering techniques is greatly aided by isotopic contrast variation. Studies of phase separated and homogeneous states have been made on several systems [6]. A description of the static scattering obtained from a diblock copolymer in the homogeneous melt has been given by Leibler [56]. It is characterised by a peak in the small angle scattering arising from the so-called 'correlation-hole'. A volume around a segment of a given type is depleted in that monomer by the constraints of the relative monomer density imposed by the molecular block structure. This is shown in Figure 13.4. Small angle scattering on such materials proves to be an excellent way of measuring the interaction, as the width of the peak seen in the scattering pattern is very sensitive to the interaction parameter xRecently the neutron spin-echo technique has been used in conjunction with isotopic labelling to test theories of copolymer dynamics. The theory has been reviewed in the context of light scattering [93]. Data for a diblock isoprenestyrene copolymer have been presented [89] that are in good agreement with the RPA theories. The findings demonstrate a general feature that is perhaps worthy of comment. The mobility of the more mobile isoprene segment can be characterised by a relaxation time r or characteristic frequency Q. The ratio Q/Q2 is not
Intensity/a.u.
Q/nrrr1
Figure 13.4 Scattering from a diblock copolymer in the homogeneous melt showing the variation with temperature. This data taken from the study of styrene-isoprene diblock polymers described in ref. [89] is measured with X-rays, although the principles are the same as for SANS measurements. If there is sufficient X-ray contrast (electron density difference) it is usually more economic to use X-rays. The width of the peak is a good measure of the interaction between the two components
Q? / nnr2 Figure 13.5 Motion of the polyisoprene block in blends of 50:50 polyisoprene/polystyrene diblock copolymer dissolved in polyisoprene at the different temperatures indicated. The data is displayed as the characteristic frequency O divided by Q2. The relaxation time is seen to tend to infinity at a finite Q vector that corresponds to the peak in the static structure seen in Figure 13.4. Full details of the interpretation of these data are to be found in ref. [89] constant, as might be expected for a normal diffusion process, but varies with Q. This is normal for the 'Rouse' or 'Zimm' relaxation modes of a polymer chain. The significant feature of the data for the copolymer shown in Figure 13.5 is that the ratio Q/Q2 tends to zero at a finite value of Q. This corresponds roughly to peak in the static structure, and demonstrates clearly that a static correlation between monomers of a given type at a given distance or Q-vector must be reflected in hindered diffusion in the same range.
IgR
Q/nnrr1
Figure 13.6 Reflectivity curves for deuterated polystyrene (390000 Mw) adsorbed at an amorphous quartz surface from a 0.1% w/w cyclohexane solution [94]. The scattering length density of the solvent was adjusted (HfD ratio) to match the quartz so that the only signal arises from the adsorbed polymer. The curves for 15 (•) and 350C (+) are shown; the adsorption between these limits was reversible. The continuous lines show approximate fits indicating an exponential polymer segment distribution with a characteristic length increasing from 95 to « 800 A
13.4.2 ADSORPTIONATSURFACES The use of neutrons to study interfaces is still relatively new, but several studies have already been made which demonstrate the scope of the technique. By way of example, some data for polystyrene adsorbing to amorphous silica [94] are shown in Figure 13.6. The curves show a large difference between 15 and 350C, which is associated with a rapid increase in adsorption as the temperature is decreased to 210C, which is the cloud point for the solution of deuterated polystyrene (molecular weight «390000) in a mixture of deuterated and protonated cyclohexane. The large increase in adsorption is substantially reversible in that, on heating, most of the polystyrene will be desorbed. This contradicts the frequently held view that physical adsorption of polymers is usually irreversible. However, it must be remembered that in this case the adsorbed layer really corresponds to a large thickness of concentrated solution. The layer is rather thicker than that of single polymer molecules, and many of the molecules may have no direct interaction with the silica surface. It must not be assumed that this type of
behaviour is typical of polymers in general, as rather few systems have been subject to detailed study, although reports in the literature do refer to polyethylene oxide adsorbing from aqueous solution [74] and to some copolymers [75]. A further point that emerges in such studies on polymers is that isotopic labelling can sometimes significantly alter the behaviour of a polymer solution or blend. Detailed studies of the system polystyrene/cyclohexane have been
Q2/nm- 2
Figure 13.7 (a) Schematic diagram of chain fragment diffusion in polymer glasses and melts; after the initial fragmentation, which can occur rapidly, the polymer segments diffuse apart and at long times will appear as separate smaller polymer molecules; (b) example data [101] for a deuterated polycar bonate/tetramethyl polycarbonate blend containing some tetraethylethane groups that can split readily under moderate heating (three or four per molecule). The data shows the diffusion at 1860C during which the polymer is observed
published [95], showing a variation of several degrees in the cloud point. Other work on deuterated polystyrene/protonated polystyrene has also shown interactions that can be significant in high molecular weight polymers [96,97].
13.4.3 KINETICS A N D POLYMER M O T I O N In recent years it has been possible to extend the use of small angle neutron scattering to study many processes that concern polymers in a variety of complex sample environments. These have included deformation and yield of elastomers and glassy polymers in conditions close to those occurring in processing and service. Other aspects of time dependent behaviour have included the study of polymerisation and observing the growth of polymer molecules during the scattering experiment. An example of the use of time dependent small angle scattering can be found in the chain fragment diffusion experiments described by Hellmann and co-workers [98-101]. Polycarbonate molecules with links that can be thermally degraded were included in a matrix of other polymers. As the chain fragments diffuse apart after the degradation, the change in apparent radius of gyration and molecular weight can be observed by SANS. This process is shown schematically in Figure 13.7(a), and some data showing the resulting data for a sample held at 186 0C are given in Figure 13.7(b). The times shown on the
Figure 13.8 Results of diffusion measurements over a range of temperatures for the system shown in Figure 13.7. The comparison with the standard WLF equation [102] (dashed line) and the glass transition temperature Tg is indicated
graph of normalised inverse intensity against Q2 indicate that the time scale of the diffusion process over a distance of about a molecular radius is several minutes. At a modern high flux reactor with good instrumentation it is possible to record data at approximately every minute. This has permitted measurements of diffusion coefficients in the range 10"1 4 - 1 0 ~ 1 8 cm2 s" 1 such as those shown in Figure 13.8.
13.5 FINAL REMARKS Neutrons provide a powerful investigative tool to study the structure and molecular motion in materials. Although available in only a few specialist laboratories, they have been widely exploited, and both sample preparation and data analysis are relatively straightforward. It is to be expected that neutron scattering will increasingly become a standard technique available to polymer scientists for characterisation of samples, and also for measurement of the structure of materials to correlate with physical properties. The possibilities of building realistic sample environments permit the study of polymers and their surfaces in conditions that are close to those in service or in polymerisation reactors.
13.6 REFERENCES [1] S. W. Lovesey, Theory of Neutron Scattering from Condensed Matter, VoIs 1 and 2, Oxford University Press, Oxford, 1984. [2] G.L. Squires, Introduction to the Theory of Thermal Neutron Scattering, Cambridge University Press, Cambridge, 1978. [3] A. Maconnachie and R.W. Richards, Polymer, 1978,9, 739. [4] J.S. Higgins and R.S. Stein, J. Appl. Crystallogr., 1978,11, 346. [5] G.D. Wignall, Neutron Scattering, Encyclopedia of Polymer Science and Engineering, Vol. 10, 2nd ed., John Wiley & Sons, 1987, pp. 112-184. [6] R.W. Richards, Adv. Polym. ScL, 1985, 71, 1. [7] F. Boue, Adv. Polym. ScL, 1987,82,47. [8] G. Allen, Makromol. Chem., Suppl., 1979, 3, 335. [9] J.B. Hayter, in S.H. Chen et al. (Eds.), Scattering Techniques Applied to Supramolecular and Non-equilibrium Systems, NA TO Adv. Study Inst. Ser., Ser. B, Vol. 73, pp. 615-638 (1988). [10] R.C. Oberthiir, Rev. Phys. Appl., 1984,19, 663. [11] A.R. Rennie, Springer Proceedings in Phys., 29, 72-84, Springer Verlag, Heidelberg. [12] D.M. Sadler in Comprehensive Polymers, Vol. 1, G. Allen and J.C. Bevington (Eds.) Pergamon, Oxford, 1989. [13] R.H. Ottewill, in F. Candau and R.H. Ottewill (Eds.), Scientific Methods for the Study of Polymer Colloids and their Applications, Kluwer Academic Publishers Dordreiht, 1990.
[14] B. Jacrot, Rep. Prog. Phys., 1976,39,911. [15] M. Born, Atomic Physics, Blackie, London, 1966. [16] D.W.L. Hukins, X-Ray Diffraction by Disordered and Ordered Systems, Pergamon Press, Oxford, 1981. [17] B. Chu, Laser Light Scattering, 2nd ed., Academic Press, San Diego, 1991. [18] W. Brown (Ed.), Dynamic Light Scattering, Oxford University Press, Oxford, 1993. [19] V.F. Sears, At. Energy Can. Ltd. [Rep.] AECL 8490,1984. [20] K. Riley, Methods of Mathematical Physics, Cambridge University Press, Cambridge, 1973. [21] H. Jeffreys and B. Swirles, Methods of Mathematical Physics, Cambridge University Press, Cambridge, 1972. [22] P. Debye, J. Phys. Colloid Chem., 1947,51,18. [23] A. Guinier and G. Fournet, Small Angle Scattering X-Rays, John Wiley & Sons, New York, 1955. [24] O. Glatter and O. Kratky (Eds.), Small Angle X-Ray Scattering, Academic Press, London, 1982. [25] G.E. Bacon, Neutron Diffraction, 3rd ed., Clarendon Press, Oxford, 1971. [26] J.M. Carpenter, G.H. Lander and CG. Windsor, Rev. Sci. Instrum., 1984,55,1019. [27] CG. Windsor, Pulsed Neutron Scattering, Taylor and Francis, London, 1981. [28] K. Ibcl, J. Appl. Crystallogr., 1976, 9, 296. [29] R.E. Ghosh and A.R. Rennie, Can we justify conventional SANS data analysis? Conf. Ser.—Inst. Phys., 1990, (107), 233. [30] A. Maconnachie, Polymer, 1984,25,1068. [31] B.H. Zimm, Journal of Chemical Physics, 1948,16,1093 & 1099. [32] M.B. Huglin (ed.) Light Scattering from Polymer Solutions, Academic Press, London, 1972. [33] F. W. Peaker in Techniques of Polymer Characterization, P. W. Allen (Ed.) Butterworths Scientific Publications, London (1959). [34] B. Jacrot and G. Zaccai, Biopolymers, 1981,20, 2413. [35] R.G. Kirste, W.A. Kruse and J. Schelten, Makromol. Chem., 1973,162, 299. [36] H. Benoit, D. Decker, J.S. Higgins, C. Picot, J.P. Cotton, B. Farnoux, G. Jannink and R. Ober, Nature, Phys. ScL, 1973,245,13. [37] R.G. Kirste, W.A. Kruse and K. Ibel, Polymer, 1975,16,120. [38] G.D. Wignall, D.G.H. Ballard and J. Schelten, Eur. Polym. J., 1974,10,861. [39] J. Schelten, G.D. Wignall, D.G.H. Ballard, G. Longman and W. Schmatz, Polymer, 1976,17, 751. [40] E. Geissler, R. Duplessix and A.-M. Hecht, Macromolecules, 1983,16, 712. [41] S. Mallam, F. Horkay, A.-M. Hecht, E. Geissler and A.R. Rennie, Macromolecules, 1991,24, 543. [42] M. Klein, A. Brulet and J.-M. Guenet, Macromolecules, 1990,23,4582. [43] S.B. Clough, A. Maconnachie and G. Allen, Macromolecules, 1980,13, 774. [44] M. Beltzung, J. Herz and C. Picot, Macromolecules, 1983,16, 580. [45] N.S. Davidson and R.W. Richards, Macromolecules, 1986,19, 2576. [46] T.A. Vilgis and F. Boue, J. Polym. Sci. B: Polym. Phys., 1988, 26, 2291. [47] E.W. Fischer, K. Hahn, J. Kugler, U. Struth, R. Born and M. Stamm, J. Polym. Sci., Polym. Phys. Ed., 1984, 22, 1491. [48] D.M. Sadler and A. Keller, Polymer, 1976,17, 37. [49] D.M. Sadler and A. Keller, Macromolecules, 1977,10,1128 [50] D.G.H. Ballard, P. Cheshire, G.W. Longman and J. Schelten, Polymer, 1978,19,379. [51] D.G.H. Ballard, J. Schelten, G.D. Wignall and G.W. Longman, Polymer, 1977,18, 1111.
[52] H. Benoit and M. Benmouna, Polymer, 1984, 25, 1059. [53] H. Benoit, M. Benmouna and Th. Vilgis, CK. Hebd. Seances Acad. ScL Ser. II, 1991, 313, 869. [54] G.D. Wignall, R.W. Hendricks, W.C. Koehler, J.S. Lin, M.P. Wai, E.L. Thomas and R.S. Stein, Polymer, 1981,22, 886. [55] S. Sakurai, H. Hasegawa, T. Hashimoto and CC. Han, Polymer, 1990,31,99. [56] L. Leibler, Macromolecules, 1980,13,1602. [57] H. Benoit and G. Hadziioannou, Macromolecules, 1988, 21,1449. [58] H. Benoit, J. PoIm. ScL, 1953,11, 507. [59] H. Benoit and P. Doty, J. Phys. Chem, 1953,57,958. [60] J. Lekner, Theory of Reflection, Martinus Nijhoff Publishers, Dordrecht, 1987. [61] M. Born and E. Wolf, Principles of Optics, 5th ed., Pergamon Press, Oxford, 1975. [62] O.S. Heavens, Optical Properties of Thin Solid Films, Butterworths, London, 1955. [63] F. Abeles, Ann. Phys., 1950,5, 596, 706. [64] T.L. Crowley, E.M. Lee, E.A. Simister, R.K. Thomas, J. Penfold and A.R. Rennie, Colloids Surf, 1990,52,85. [65] T.L. Crowley, Physica A, 1993,195, 354. [66] J. Penfold and R.K. Thomas, J. Phys. Condensed Matter, 1990, 2,1369. [67] T.P. Russell, Mater. ScL Rep., 1990,5,171. [68] A. Menelle, T.P. Russell, S.H. Anastasiadis, S.K. Satija and CF. Majkrzak, Phys. Rev. Lett., 1992,68, 67. [69] R.A.L. Jones, LJ. Norton, EJ. Kramer, RJ. Composto, R.S. Stein, T.P. Russell, A. Mansour, A. Karim, G.P. Felcher, M.H» Rafailovich, J. Sokolov, X. Zhao and S.A. Schwartz, Europhys. Lett., 1990,12,41. [70] M.L. Fernandez, J.S. Higgins, J. Penfold, R.C. Ward, CS. Shackleton and DJ. Walsh, Polymer, 1988,29,1923. [71] A.R. Rennie, RJ. Crawford, E.M. Lee, R.K. Thomas, T.L. Crowley, S. Roberts, M.S. Qureshi and R. W. Richards, Macromolecules, 1989, 22, 3466. [72] L.T. Lee, O. Guiselin, B. Farnoux and A. Lapp, Macromolecules, 1991,24, 2518. [73] O. Guiselin, L.T. Lee, B. Farnoux and A. Lapp, J. Chem. Phys., 1991,95,4632. [74] E.M. Lee, R.K. Thomas and A.R. Rennie, Europhys. Lett., 1990,13,135. [75] S.K. Satija, CF. Majkrzak, T.P. Russell, S.K. Sinha, E.B. Sirota and GJ. Hughes, Macromolecules, 1990, 23, 3860. [76] J.B. Field, C Toprakcioglu, R.C. Ball, H.B. Stanley, L. Dai, W. Barford, J. Penfold, G. Smith and W. Hamilton, Macromolecules, 1992,25,434. [77] L.T. Lee, E.K. Mann, D. Langevin and B. Farnoux, Langmuir, 1991, 7, 3076. [78] J.A. Henderson, R.W. Richards, J. Penfold, C Shackleton and R.K. Thomas, Polymer, 1991,32, 3284. [79] LR. Gentle, P.M. Saville, J.W. White and J. Penfold, Langmuir, 1993,9, 646. [80] M. Doi and S.F. Edwards, The Theory of Polymer Dynamics. Clarendon Press, Oxford, 1986. [81] P.A. Dagleish, J.B. Hayter and F. Mezei, in F. Mezei (Ed.), Neutron Spin Echo, Lect. Notes Phys., 1980,128, pp. 66-71. [82] L.K. Nicholson, Contemporary Physics, 1981, 22,451. [83] B. Ewen, B. Stiihn, K. Binder, D. Richter and J.B. Hayter, Polym. Commun., 1984, 25,133. [84] P.G. de Gennes, J. Chem. Phys., 1971, 55, 572. [85] J.S. Higgins, L.K. Nicholson and J.B. Hayter, Polymer, 1981, 22,163. [86] J.S. Higgins, K. Ma, L.K. Nicholson, J.B. Hayter, K. Dodgson and J.A. Semlyen, Polymer, 1983, 24, 793. [87] D. Richter, B. Ewen, B. Farago and T. Wagner, Phys. Rev. Lett., 1989,62, 2140.
[88] D. Richter, B. Farago, LJ. Fetters, J.S. Huang, B. Ewen and C. Lartigue, Phys. Rev. Lett., 1990,64,1389. [89] B. Stuhn and A.R. Rennie, Macromolecules, 1989, 22, 2460. [90] D. Richter, B. Frick and B. Farago, Phys. Rev. Lett., 1988,61, 2465. [91] D. Richter, R. Zorn, B. Frick and B. Farago, Ber. Bunsen-Ges., 1991,95,1111. [92] D. Richter, R. Zorn, B. Frick, B. Farago and LJ. Fetters, Phys. Rev. Lett., 1992,68, 71. [93] A.Z. Akcasu, in W. Brown (Ed.), Dynamic Light Scattering, Oxford University Press, Oxford, 1993. [94] J. McCarney. J.R. Lu, R.K. Thomas and A.R. Rennie, Colloids and Surfaces A Physicochemical and Engineering Aspects, 1994,86,185.
[95] Th. Engels, L. Belkoura and D. Woermann, Ber. Bunsen-Ges. Phys. Chem., 1988,92, 1544. [96] F.S. Bates and G.D. Wignall, Mactromolecules, 1986,19, 932. [97] F.S. Bates and G.D. Wignall, Phys. Rev. Lett., 1986,57,1429. [98] E.H. Hellmann, G.P. Hellmann and A.R. Rennie, Colloid Polym. Sci., 1991,269,343. [99] E.H. Hellmann, G.P. Hellmann and A.R. Rennie, Macromolecules, 1991,24, 3821. [100] E. Meurer, KJ. Kuhn, G.P. Hellmann and A.R. Rennie, J. Polym. Sci. B: Polym. Phys., 1990, 28, 893. [101] E. Hellmann, 'Kettenfragment diffusion in Blends am Glassubergang\ Doktorarbeit, Universitat Mainz, 1991. [102] J.D. Ferry, Viscoelastic Properties of Polymers, 3rd ed., John Wiley & Sons, New York, 1980.
14 OPTICAL ACTIVITY AND THE STRUCTURE OF MACROMOLECULES F. CIARDELLI Department of Chemistry and Industrial Chemistry, University of Pisa, Italy and CNR, Center for Stereordered and Optically Active Macromolecules, Pisa, Italy
O. PIERONI Department of Chemistry and Industrial Chemistry, University of Pisa, Italy and CNR, Institute of Biophysics, Pisa, Italy and
A. FISSI CNR, Institute of Biophysics, Pisa, Italy
14.1 INTRODUCTION 14.1.1 ORIGIN OF OPTICAL ACTIVITY IN MACROMOLECULES As in low molecular weight compounds, optical activity can be observed only in chiral macromolecules, that is, macromolecules for which all allowed conformations lack reflection symmetry elements. The identification of chiral macromolecules differs from that of low molecular weight molecules because of the substantially linear structure along the chain backbone. Accordingly, analysis of the symmetry properties has been carried out on the basis of three different models: (i) the infinite length chain; (ii) the finite length chain with equal end groups; and (iii) the finite length chain with different end groups. Point symmetry valid for molecules having definite and 'discrete' dimensions in all directions can be used only for the last two models, whereas linear symmetry must be used for the first model, which implies an infinite dimension [I]. In linear symmetry, in contrast to point symmetry, the new symmetry operation 'translation' and the new symmetry element 'translation axis' are introduced. The analysis for aflexiblemacromolecule, which can assume an extremely large number of conformations, is conveniently carried out on the most symmetric of these conformations, which is usually the 'planar zigzag'. The Polymer Spectroscopy. Edited by Allan H. Fawcctt © 19% John Wiley & Sons Ltd
analysis of the derived Fischer projection of an infinite chain indicates that this is chiral when a symmetry plane containing the chain, those perpendicular to the chain and that with translation containing the chain are all lacking [ I ] . For finite length chains, chirality is guaranteed by lack of symmetry in the plane containing the chain and the plane perpendicular to the chain at its central point [I]. Thus, in vinyl polymers, only atactic macromolecules can be chiral in the first model with infinite length chain. The atactic and the syndiotactic macromolecules with an even number of monomer residues are chiral in the finite chain model with identical end groups, whereas all isotactic, syndiotactic and atactic chains have a chiral structure for the last model with different end groups [1,2]. Even in vinyl polymers consisting of chiral macromolecules, extremely low or vanishing optical rotation can be predicted when the molecular weight is high, even if a complete separation of the enantiomeric pair were to be possible. Indeed, in vinyl polymers every stereogenic carbon atom is flanked by two CH 2 groups, and its chirality arises only from the different lengths of the two chain sections attached to it. Thus an appreciable contribution to chiroptical properties is conceivable only for asymmetric centers close to the chain ends, the concentration of which decreases with increasing molecular weight. The same holds for conformational optical activity referred to the presence of secondary structures, involving the macromolecule as whole or a substantial fraction of it, with a predominant handedness. This last is not attributable simply to stereogenic centers (asymmetric centers) with a single absolute configuration in each repeating unit. Indeed, the existence of purely conformational optical activity is not a unique macromolecular requisite, being Well known in low molecular weight atropisomerism. However, in polymers it assumes a very specific characteristic connected with the occurrence of cooperative effects which allow transmittance of molecular asymmetry along the chain to very long distances [3]. Most isotactic vinyl polymers assume helical conformations in the crystalline state [4], but owing to the substantial achirality of the macromolecules both screw senses are found in the lattice cells in equal amounts. This is even more true in a melt or in solution, where left-handed and right-handed helical sections alternate even within the same macromolecular chain. Certainly, an appreciable optical rotation would be observed in the crystalline state provided that crystallization occurred under a chiral field inducing a single screw sense helicity in all chains. Such an optical rotation would promptly be lost on melting or dissolution, as an immediate equilibration between the two opposite helical senses would occur. Isotactic macromolecules derived from achiral monomers have no preference for right- or left-handed screw senses, and the two are perfectly balanced at interand intramolecular level. However, distribution of left- and right-handed helical secondary structures affects markedly the free energy of the system, alternation of the two senses in the same chain being favored for entropic reasons [4,5]. If this last situation takes place, conformational optical activity cannot be obtained
because of intramolecular conformational compensation, which hinders any isolation of chains with a predominant handedness. The above considerations are described in fuller detail in previous papers [6,7] and indicate that macromolecules assuming a single chirality conformation can show chiroptical properties characteristic of the conformation itself. Moreover, if chromophores are present in the side chains specific chiroptical properties can arise from dipole-dipole electronic interactions among these chromophores disposed along the chirally arranged backbone. This situation is clearly shown in poly-a-amino acids, in which specific and typical chiroptical properties are associated with specific and typical conformations (a-helix, /?-structures, random coil) [8]. In order to make one screw sense largely predominant in a single macromolecule, the intramolecular equilibration must be hindered by building very rigid chains. In the limiting case the chains will form rods having helical structure with either left- or right-handed helicity. Even if hindering of equilibration can be considered as a kinetic effect, it cannot be excluded that thermodynamic contributions are involved, particularly when rigidity is due to bulk side chains and conformational reversals have a very high internal energy [5,9]. In other words, bulky side chains in a vinyl type structure, for instance, can favor the formation of longer helical chain sections, as the lower entropy is balanced by the gain in internal energy due to minimization of the number of conformational reversals. Indeed these last have, in the case of macromolecules with bulky and branched side chains, larger internal energy per structural unit than the same unit in the helical conformation [10]. Which mechanism is actually operating cannot be established from primary structure as a general rule, an increase in temperature in both cases favoring the equilibration of the two screw senses within each macromolecular chain. Chiroptical properties (molar rotatory power [ $ ] and molar ellipticity [0]) result from a weighted average of the contributions from different conformers, as shown in the following equations:
where N1 indicates the molar fraction and [O] 1 (or [G]1) indicates the molar rotatory power (or molar ellipticity) of the i-th conformer. In macromolecules the molar entities refer to one residue; thus the chiroptical properties are independent of molecular weight, at least when this is very high. It has been demonstrated that in isotactic polymers of optically active a-olefins the molar optical rotation per monomeric residue can be interpreted in terms of the prevalence of few conformations with very high optical rotation of the same sign, corresponding to those allowed to the structural unit inserted in an one screw sense helix [11].
Moreover, in coisotactic copolymers of optically active a-olefins with vinylaromatic monomers, it was shown that the aromatic groups in the side chains give rise to dichroic bands in the spectral region of the n->n* electronic transitions. In several cases exciton splitting was also observed, corresponding to the strong allowed n-m* electronic transition [12]. This circulardichroism (CD) couplet was confirmed to be connected to the dipole-dipole electronic interaction of transition moments of aromatic moieties disposed in a mutual chiral geometry with a predominant handedness, such as that of a one screw sense helix [13,14]. This clear demonstration that the extrinsic CD bands of side chains are related to main chain conformation indicates their usefulness as 'conformational probes'. A typical case comprises polypeptides with aromatic side chains masking the peptide absorption bands [15].
14.1.2 OBJECTIVE With reference to the concepts summarized in the previous section, optical activity measurements can be particularly effective in providing structural information on polypeptides with side chains absorbing at a wavelength clearly distinct from that of the peptide group. In some favorable cases, moreover, CD spectra can allow the detection of very specific structural features, including non-bonding interactions. On the other hand, the same data can be used for monitoring even subtle structural changes induced by external factors. Accordingly, the evaluation of chiroptical properties allows one to follow crucial conformational changes accompanying such biological phenomena as substrate binding, macromolecule-macromolecule interactions, and so on. In order to substantiate these last considerations, the present paper is devoted to the description of studies, using mainly CD spectra, concerning poly (Lglutamic acid)- and poly(L-lysine)-bearing photochromic side chains. Light irradiation of these polypeptides gives rise to reversible isomerization of the photoresponsive chromophores attached to the macromolecule's backbone, which itself can then undergo reversible conformational changes. These may be accompanied by reversible variations of the polymer's properties, such as viscosity, solubility and so on. The main objective of this paper is to show that the examination of the chiroptical properties during the above mentioned phenomena can allow the correlatation of changes in conformation and properties with the photoresponse of the chromophores. On the same basis, an interpretation at the molecular level can be put forward of the reversible variation of viscosity and the solubility. It is also hoped that these indications may be useful for developing photorecording devices which can be read using their chiroptical properties.
14.2 CHIROPTICALPROPERTIESOF PHOTOCHROMIC POLYPEPTIDES 14.2.1 POLYPEPTIDES PHOTORESPONSIVE TO UV LIGHT 14.2.1.1 Azobenzene-containing Polypeptides
Polypeptides sensitive to irradiation with near UV light can be prepared by introducing photochromic azobenzene units into the side chains of high molecular weight (Mv = 100000-250000) poly(a-amino acid)s, such as poly(L-glutamic acid) or poly(L-lysine). Macromolecules having the structures represented in Figure 14.1, and containing various percentages of azo groups, can be obtained under various reaction conditions [16,17]. The photoisomerization of azobenzene moieties (Figure 14.2) is the event responsible for the photochromic behavior of these macromolecules. At room temperature in the dark all azo groups are in the trans configuration, which is planar and then fully conjugated. Irradiation produces isomerization to the cis configuration which, by contrast, is not planar for steric reasons. At the photostationary state, the relative composition of the two isomers depends only on the incident light. The maximum photoconversion to the cis isomer (85 %) is achieved by irradiating at 350-370 nm, whereas the maximum yield of the back reaction from the cis to the trans isomer (80%) is achieved by irradiating at 450 nm. With a lamp having a power of 100 W, irradiation for 1 or 2 min is enough to achieve the photostationary state. By dark adaptation, the metastable cis chromophores decay again to the trans form. The thermal decay at room temperature in the dark is rather slow for azo-modified poly(L-glutamic acid) and takes more than 20Oh to restore the all-trans isomeric composition; for azo-modified poly(L-lysine) the decay in the
Figure 14.1 Chemical structures of poly(L-glutamic acid) and poly(L-lysine) containing azobenzene units in the side chains
Figure 14.2 Photochromic behavior of azobenzene-containing poly(L-glutamic acid). Reproduced by permission of Elsevier Science S.A. from J. Photochem. Photobiol. B: Bioi 1992,12,125-140
dark takes place so slowly that it cannot be observed under normal experimental conditions. The photochromic cycles are completely reversible and can be repeated at will, without any apparent fatigue. As a consequence of the different electronic situations, the two isomers have markedly different absorption, and the photo-isomerization is accompanied by strong variations in the spectra (Figure 14.2). In particular, the trans-to-cis isomerization is revealed by a strong decrease of the intense band at « 350 nm associated with a n-n* transition and a contemporaneous increase of the band at 450 nm associated with the n-n* transition of the azo-chromophore. 14.2.1.2 Light-induced Conformational Changes
Poly(L-glutamate)s having azobenzene units in the side chains, in organic solvents such as trimethyl phosphate (TMP), show the typical CD curve of the a-helix structure, with two minima at 208 and 222 nm. Above 250 nm, the dark-adapted samples exhibit also a couplet of bands centered at 350 nm, corresponding to the n-n* transition of the azo chromophore in trans configuration. The trans-to-cis photoisomerization completely cancels the side chain CD bands in the region of 35Onm, but does not modify at all the CD spectra in the peptide region. This indicates that, in these solvents, light causes the isomerization of the azo side chains, but the isomerization does not induce any variation of the polypeptide main chain.
Figure 143 CD spectra of poly(L-glutamic acid) bearing 36 mol% azobenzene units, before ( ) and after ( ) irradiation, in aqueous solution at various pH values: A, pH 4.8; B, pH 6.5; C,pH 8.0 The secondary structure in water depends on the molar content of azobenzene units and also on the degree of ionization of the unmodified COOH side chains. Below pH 5, a sample of poly(Glu) bearing 35 mol% of azobenzene units assumes a /^-structure. Irradiation does not induce any variation of the polypeptide conformation. At pH values above 7, the polypeptide adopts a random coil conformation which is again not affected by the photoisomerization of the azo side chains. However, at pH values of 5-7, irradiation produces a remarkable decrease of the ordered structure (Figure 14.3). In this range of pH the trans-to-cis isomerization produces a higher degree of ionization of the unmodified COOH side chains, thus amplifying the first light effect and causing unfolding of the polypeptide. Cationic surfactants are known to affect the conformation of poly(L-glutamic acid). This suggested to us that it might be possible to combine the isomerization of the photochromic side chains with the surfactant effect to obtain an amplification of the photoresponse. The expectation was realized by irradiating azomodified poly(L-glutamic acid) in the presence of dodecylammonium chloride (DAC) at the critical micelle concentration (c.m.c) [18]. Figure 14.4 shows the CD spectra of a 20% azo-modified poly(Glu) both in the absence and in the presence of DAC. In the absence of detergent at pH 7, the polymer is completely in random coil conformation and not affected at all by irradiation. In the presence of detergent at the c.m.c, irradiation at 35Onm (trans-to-cis isomerization) induces
Figure 14.4 CD spectra of poly(L-glutamic acid) bearing 20 mol% azobenzene units, at pH 7.6, before ( ) and after ( ) irradiation: A, in the absence of dodecylammonium chloride (DAC); B, in the presence of DAC, below the c.m.c; C, in the presence of DAC, at the c.m.c. Reproduced by permission of Elsevier Science S.A. from J. Photochem. Photobioi B: BioL, 1992,12,125-140 an evident coil-to-helix transition. The variation is completely reversible when the sample is dark-adapted or irradiated at 450 nm (cis-to-trans isomerization). Thus, in the presence of DAC micelles, the polypeptide conformation can be photomodulated by exposure alternately to light or dark, or by irradiating at two different appropriate wavelengths. The key factor responsible for the photoinduced variations of conformation is the affinity of the azo-polymer for the micelles. Such an affinity, in fact, is likely to be different when the azo side chains are in trans or in cis configuration. When azo-units are in the planar, apolar, trans form, they dissolve within the hydrophobic core of the micelles, forcing the polypeptide chains to assume a coil conformation. Isomerization of the azo units to the skewed, polar, cis form inhibits hydrophobic interactions and causes the azo-units to leave the micelles, thus allowing the polypeptide chains to adopt the a-helix structure (which is favored in the absence of micelles). In other words, the primary photochemical event is the trans ^ cis isomerization of the azobenzene units, but the driving force of the photoresponse should be the different location of the macromolecules relative to the micelles. Dark-adapted (all trans azo groups) poly(L-lysine) bearing 43 mol% of azobenzene groups, in a medium of hexafluoroisopropanol/water/sodium dodecyl sulfate, shows a CD spectrum which can be attributed to the presence of a /?-form. Irradiation at 340 nm causes the disruption of the /!-structure and promotes the formation of an a-helix (helix content % 50%), as revealed by the appearance of the typical CD pattern. Upon irradiation at 450 nm, the spectrum reverses again
to the original one. The photoinduced /J;=± helix conformational change is completely reversible and the two conformations can be obtained by irradiating alternately at the two different wavelengths. This photoinduced fi^± helix change can readily be explained on the basis of the different geometry and hydrophobicity of the trans and cis azobenzene units. The /J-form is stabilized by hydrophobic interactions among the side chains and is favored when the azobenzene units have the planar geometry and the high hydrophobicity of the trans configuration. The interactions are inhibited when the azo units are isomerized to the skewed cis configuration, and thus the /^-structure is destabilized and destroyed. The polypeptide chains then adopt the a-helix form in the helix-supporting solvent hexafluoroisopropanol. The photochromic behaviour of azobenzene-containing poly(L-lysine) has also been reported in the monolayer state [19]. When the polypeptide monolayer is kept at constant area, alternate irradiation with visible and ultraviolet light produces reversible changes (% 25%) of the surface pressure of the monolayer. 14.2.13 Photostimulated 'Aggregation-Dissaggregation' Effects
CD data provided evidence that azo-modified poly(Glu) containing azobenzene units can undergo reversible aggregation-disaggregation processes upon exposure to light or dark conditions [20]. Samples stored in the dark or irradiated at 450 nm (azo groups in the trans configuration) show variations of their CD spectra on aging in a TMP/water solution (Figure 14.5). The time dependence is characterized by the gradual appearance of an intense side chain CD couplet together with a progressive distortion of the a-helix pattern, typical of the effects produced by aggregates of polypeptide chains [21,22]. Irradiation at 361 nm (tran to cis isomerization) produces the full restoration of the initial CD spectra, indicating dissociation of the aggregates. The spectra revert again to the distorted ones on irradiating at 450 nm or by dark adaptation, thus confirming the reversibility of the light-induced effect. Investigation of azo-modified poly(Glu) containing 85 mol% azobenzene units in the side chains has provided confirmation of the occurrence of aggregationdisaggregation processes induced by light, together with the possibility of photoregulating polymer solubility [23]. This polypeptide, when stored in the dark, assumes the a-helix structure in hexafluoroisopropanol (HFP). Addition of a small amount of water (15 vol%) to the HFP solution causes the formation of aggregates, followed by precipitation of the polymer as a yellow material. The precipitation is total and quantitative, as can be seen by recording the absorption spectrum of the filtered colorless liquid. Complete dissolution of the polymer was obtained by irradiation of the suspension for a few seconds at 350 nm; irradiation at 450 nm or dark adaptation of the solution caused the polymer to precipitate. In a HFP/water = 85/15 solvent mixture, therefore, the 'precipitation-dissolution' cycles can be reversibly
Figure 14.5 Poly(L-glutamic acid) bearing 20mol% azobenzene units. CD spectra in trimethyl phosphate/water = 50/50, recorded at various aging times: (1) freshly prepared solution; (2) aged 1 day; (3) aged 2 days; (4) aged 3 days. ( ) Dark-adapted samples; ( ) irradiated at 360 nm, at any aging time. Reprinted with permission from [23]. Copyright 1989 American Chemical Society
repeated by irradiation and dark adaptation, or by irradiating at two different wavelengths. The dependence of the polymer solubility on the cis/trans composition of the azobenzene side chains was investigated by performing irradiation experiments at various wavelengths of the incident light. The considerable amount of photodissolved polymer allowed its determination by evaporating the solutions obtained upon irradiation and weighing the dry residue. The solubility of the polymer, as a function of the cis/trans ratio of azobenzene side chains, is described by a sharp sigmoidal curve. The polymer is fully insoluble when more than 60% of the azo groups is in a trans configuration. By contrast, the maximum amount of
photosolubilization is achieved when 60% of azo groups are in the cis configuration; the solubility then remains unaffected at higher values of cis content [23]. The described photoresponse effects can be well interpreted on the basis of association among macromolecules through hydrophobic interactions and stacking of azobenzene side chains. The planar, apolar, trans configuration gives aggregation and precipitation; when the azo moieties are photoisomerized to the skewed, polar, cis configuration, interactions and stacking between azo-groups are inhibited, so that disaggregation of the macromolecules takes place and polymer dissolution occurs. 14.2.2 PHOTOMODULATION OF POLYPEPTIDE CONFORMATION BY SUNLIGHT 14.2.2.1 Spiropyran-containing Polypeptides Azo-modified polypeptides could be considered as models for photoregulated processes occurring in nature, but the generation of cis and trans photoisomers, and hence photoregulation, requires artificial sources of ultraviolet light. Ideally, one would like to have a model system responding to the presence or absence of sunlight, such as polypeptides bearing spiropyran groups attached to poly(Lglutamic acid) [24] or poly(L-lysine) [25] (Figure 14.6). Spiropyran-modified poly(L-glutamate)s in hexafluoroisopropanol (HFP) exhibit 'reverse photochromism', that is, a photochromic behavior which is
Spiro
Spiro
Spiro
group
Figure 14.6 Chemical structure of poly(L-glutamic acid) and poly(L-lysine) bearing spirobenzopyran units in the side chains
dark light
Figure 14.7 Structure and reverse photochromic reactions in HFP of poly(L-glutamic acid) containing spiropyran units in the side chains opposite to that usually observed in most common organic solvents. Thus, HFP solutions kept in the dark at room temperature show a yellow-orange color which is completely bleached upon exposure to visible light and is reversibly restored in the dark. NMR data confirm that the photochromism in HFP involves the well-known interconversion between the colorless closed spiro structure / and the colored ring-opened merocyanine structure II (Figure 14.7). Accordingly, in the 13C NMR spectra of the colorless solution the resonances of the geminal methyl groups appear as two separate peakes, 27.0 and 21.0 ppm, as a consequence of the presence of the chiral spiro carbon atom. In the colored solution, by contrast, the two methyl group resonances merge to a single signal at 28.7 ppm, analogously to that observed for the proton resonances. The spectra of the colored species kept in the dark do not show nuclear resonances associated with the spiro form, indicating that in HFP the spiropyran ;=± merocyanine equilibrium is fully shifted to the right. The very polar solvent HFP is probably responsible for the reverse photochromism by stabilizing the charged merocyanine species II more than the apolar spiropyran species I. A protonated open structure III might also be formed between the zwitterionic species II and HFP, with the solvent functioning as an acid, as will be described in the following section. The dark-adapted sample shows a spectrum which displays two absorption maxima, at 500 and 370 nm, of about the same intensity (Figure 14.8). Irradiation with visible light (500-550 nm) or exposure to sunlight for a few seconds completely dispels the absorption in the visible region and gives rise to the spectrum of the colorless spiro form, characterized by absorption maxima at 355 and 272 nm. In the dark, the original spectrum is progressively restored. In the colorless indolinospiropyran species, the two halves of the molecule are topologically independent, so the absorption spectrum consists mainly of
light
dark
A
Figure 14.8 Variation of the absorption spectra as a function of irradiation and darkadaptation time for poly(L-glutamic acid) bearing 85mol% spiropyran units in HFP (c = 5.01 x 10"2 g/1; / = 1 cm); 1, dark-adapted solution; 2, irradiated solution
localized transitions belonging to a particular half of the molecule, rather than delocalized transitions belonging to the molecule as a whole [26]. The electronic transition at longer wavelength, which in HFP occurs at 355 nm, has been assigned to the benzopyran, and the second transition, which in HFP is seen at 272 nm, has been assigned to the indoline portion of the molecule [26]. In the colored species, the absorption band at 500 nm can be assigned to a n-n* electronic transition of the extended and conjugated merocyanine chromophore, and the 370 nm band can be attributed to a charge-transfer transition from the oxygen atom of the benzopyran ring to the electron-accepting nitro substituent [27,28].
light dark
Figure 14.9 Reverse photochromic reactions of spiropyran salts (see Fig. 14.7) Considering the acidity of HFP, the band at longer wavelengths should be assigned to the zwitterionic ring-opened form II (Figure 14.9), whereas the band at shorter wavelengths might be assigned to the presence of the ring-opened species III formed between the zwitterionic species and HFP, with this last acting as an acid (pKa = 9.30) [29] (shown in Figure 14.9). In the polymers, protonation of the open form by unmodified COOH side chains may also occur [27], even though this effect cannot play a relevant role in the 85 mol% modified polymer shown in the figure. The presence of a well-defined isosbestic point should indicate only two interconverting species. However, the salt of the spiropyran with trifluoroacetic acid exibits exactly the same isosbestic point (see the following section), so that one cannot exclude the presence of both the zwitterionic and the protonated merocyanine forms. 14.2.2.2 Photomodulation of Conformation
The CD spectra of the dark-adapted samples of poly(Gluj bearing 85 mol% photochromic units are those of random coil polypeptides. CD bands of small intensity are also present in the near UV-visible region, in correspondence with the merocyanine electronic transitions. The solutions bleached after exposure to visible light display the typical pattern of the a-helix, with the two minima at 222 and 208 nm, thus indicating that the isomerization of the side chains produces spiralization of the main chain. The back reaction in the dark is accompanied by a progressive decrease of the helix content and restoration of the original disordered conformation (Figure 14.10). The photoinduced helix content can be only approximately estimated on the basis of the CD spectra. In fact, several polypeptides, all having a-helical
light
dark
Figure 14.10 Effect of irradiation and dark adaptation on CD spectra of poly(L-glutamic acid) bearing 85 mol% spiropyran units in HFP at 250C: 1, kept in the dark; 2, exposed to sunlight; dotted lines are CD spectra recorded during decay in the dark over 8h. Below 250 nm, CD data are expressed in terms of molar ellipticity based on the mean residue molecular weight; above 250 nm, the molar ellipticity is referred to one spiropyranglutamyl residue conformation, were reported [30] to show significant variations of the maximum ellipticities when CD spectra were measured in HFP. On the basis of the literature values [30] of [ G ] 2 2 2 ( - 3 0 0 0 0 — 40000) for 100% a-helix in HFP, the photoinduced helix content can be evaluated as 90-70%. The photochromic reaction involves the reversible conversion of the zwitterionic merocyanine (sample kept in the dark) to the uncharged spiro form (sample exposed to light); the isomerization is thus accompanied by large variations of the electrostatic interactions among the side chains of the polypeptides. Intrachain interactions should produce loops in the macromolecules, whereas intermolecular stacking should produce aggregation phenomena. As
a result, the macromolecules are forced to adopt a disordered structure. When the sample is exposed to light and merocyanines are converted to the neutral spiro form, electrostatic interactions are removed and the polypeptide can adopt the a-helix structure. When spiropyrans are treated with acids they are converted into 'spiropyran salts', which exhibit photochromic behavior differing from that of the parent spiropyran compounds. The gross mechanism proposed is illustrated in Figure 14.9. In the dark at room temperature, the compounds give colored solutions due to the presence of the O-protonated merocyanine species III. The open form is converted by irradiation with visible light to the iV-protonated spiro form IV. As spiropyrans are fairly strong bases in the open form but very weak bases in the closed spiro form, the charged species IV can lose a proton and the neutral species I can be actually formed. Comparison of Figure 14.9 with Figure 14.7 shows that different photoisomers are involved in acidic or non-acidic solution. Therefore we may expect spiropyran-containing polypeptides to be affected by light in a different way depending on whether they are irradiated in the absence or in the presence of acid. Poly(spiropyran-L-glutamate) in HFP solution in the presence of TFA does not give light-induced conformational changes. Actually, the solutions show the typical CD spectra of random coil polypeptides both when they are kept in the dark and when they are exposed to light. The addition of methanol as a cosolvent induces the coil -> helix conformational transition, as for other polypeptides having salified side chains [31]. The most remarkable aspect of this system is that two distinct curves are observed for the dark-adapted sample and for the irradiated one (Figure 14.11). In particular, for the polymer containing 85mol% photochromic units, the concentration of methanol needed to induce the conformational transition is « 10-40% for the sample kept in the dark and « 5-10% for the sample exposed to light. Therefore, at any solvent composition in the range between the two curves, exposure to light produces reversible variations of the helix content (Figure 14.12). The photochromic reactions schematized in Figure 14.9 and the above discussed absorption spectra allow us to explain the conformational behavior. In HFP, in the presence of TFA, the photochromic side chains are protonated by the acid either when the sample is kept in the dark (photochromic units present as open species III) or when the sample is exposed to light (photochromic units present as closed species IV). In both cases the polypeptide is essentially a polycation, so the repulsive forces among the side chains make the macromolecules adopt an extended coil conformation, and no photoresponse is observed. In the presence of methanol (> 10%) the equilibrium between the two colorless species IV and I (Figure 14.9) is shifted toward the neutral spiro structure I. In these conditions the 'folding-unfolding' of the macromolecules is photocontrolled by the isomerization of the photochromic units. In the dark, they are present as charged species, so the macromolecules adopt a disordered conformation.
MeOH,%
dark
light
Figure 14.11 Variation of ellipticity at 222 nm as a function of methanol concentration (v/v) for poly(Glu) bearing 85mol% spiropyran units in HFP/MeOH/TFA solvent system, at 250C: ( ) dark-adapted sample; ( ), irradiated sample
Figure 14.12 Effect of irradiation on CD spectra for poly(Glu) containing 85 mol% spiropyran units at 250C in various HFP/MeOH/TFA solvent mixtures (c = 2.59 x 10"2 g/1; TFA = 1 x 10"3 ml in 2 ml of mixed solvent); MeOH % (v/v): (a) 0-5%; (b) 10%; (c) 20%; (d) 40%. ( ), dark-adapted; ( ) irradiated samples
a-helix v a r i a t i o n , %
merocyanine
form , %
Figure 14.13 PoIy(GIu) bearing 85 mol% spiropyran units. a-Helix relative variation as a function of spiropyran/merocyanine isomeric composition of the side chains, in pure HFP ( ) and HFP/MeOH/TFA = 90/10/5 x 10 " 2 ( ). The a-helix variation in % was estimated as {[@]V[©]°} x 100, where [0]° and [ 0 ] ' are the ellipticity values measured at 222 nm at the beginning and at the time t during decay, at 250C Exposure to light and the consequent photoconversion of the side chains to the apolar spiro form make the macromolecules adopt the a-helix conformation. In order to investigate the dependence of the secondary structure on the isomeric composition of the photochromic side chains, the rate of the helix-tocoil conversion and the rate of appearance of absorbance at the longer wavelength in the dark were measured simultaneously. The helix content was then plotted as a function of the photochromic units present in the merocyanine form (Figure 14.13). In pure H F P (Figure 14.13, dotted line), the helical structure starts to break up rapidly as soon as the merocyanine species begin to be formed, and the helix -* coil conformational change takes place almost entirely following conversion of % 30% of spiropyran to merocyanine groups. A rather different behavior is observed in HFP/MeOH/TFA (Figure 14.13, full line): the helix content decreases slowly with increasing merocyanine percentage, but the helical structure does not collapse until « 50% of the spiro groups are isomerized to the merocyanine form. The different dependence of the helix structure on the percentage of photochromic groups present in the merocyanine form is a confirmation that denaturation of the macromolecules in the dark occurs through different mechanisms in non-acidic and acidic media. In the former case denaturation should be caused by
stacking and aggregation between zwitterionic merocyanine species II. In the latter case, denaturation should be caused by repulsive forces among the cationic side chains III. In both cases, exposure to light removes the electrostatic interactions between side chains, allowing the formation by the polypeptide chains. Also poly(Lys)-containing spirobenzopyran side chains, as well as low molecular weight model compounds, exhibit intense 'negative' photochromism in HFP [25]. In the dark the solutions are orange, with two absorption maxima at «470nm (£mol = 31700) and 370 nm (emol = 32 000). Exposure to sunlight is accompanied by prompt bleaching, with a shift of the absorption maxima to 353 nm (emol = 11200) and 270-272 nm (emol = 16 200). The original spectrum is reversibly restored when the illumination is stopped. Decay in the dark at 25 0 C follows first order kinetics for the model compounds, with a rate constant of 5.7 x 10" 3 min" 1 and a half-life of «122min. For the polymer, the kinetics deviate slightly from monoexponential and biexponential decay; the time necessary to restore half of the original absorbance is « 80 min. The analogy of these reversible processes with those observed in spiropyranmodified poly(Glu) suggests the occurrence of similar photochromic reactions. Accordingly, HFP stabilizes the colored ionic merocyanine structure, while irradiation gives the colorless spiro structure. In pure HFP, the CD spectra are consistent with those of random coil polypeptide chains, and the photoisomerization reaction does not affect the polymer conformation at all. Addition of triethylamine to the HFP solution induces the coil-•helix transition, but the amount of base necessary to induce the transition is different for the dark-adapted sample (15% v/v) than for the irradiated one (30% v/v). Thus, at any composition in the range 3-15% v/v of NEt 3 , exposure to sunlight produces reversible variations of the helix content. Combination of the effects due to the photochromic behavior with appropriate amounts of NEt 3 allows modulation of the extent of the photoresponse. It appears that in pure HFP the conformation for poly(Lys)-containing spiropyran is determined by the unmodified Lys side chains protonated by the acid solvent; as a consequence, the polypeptide assumes a coil conformation which is not affected by the isomerization of the photochromic groups. Addition of a moderate amount (3-15%) OfNEt3 removes protons from Lys side chains, whose basicity depends on isomeric composition of the photochromic moieties. In the range between the transition curves of the dark-adapted and the irradiated sample, the chain folding ;=± unfolding is then controlled by the isomerization of the photochromic side chains: when these are in the charged merocyanine form, the polypeptide chains are in the random coil arrangement, but photoconversion to the apolar spiropyran form causes the macromolecules to assume a helical conformation. At NEt 3 contents above 15%, the high concentration of a NEt 3 • HFP saline complex can probably exert a shielding effect on the charged
side chains, allowing the polypeptide to stay in the helical conformation at any photoisomeric composition. The system described provides a well defined example of the combined action of light and environment on the secondary structure of polypeptides. It can thus be considered as a macromolecular model resembling the behavior of naturally occurring photoreceptors [32]. 14.2.2.3 Photoinduced Variations of Viscosity
a-helix v a r i a t i o n , %
The colored solutions of poly(L-glutamic acid) and poly(L-lysine) containing spiropyran, when kept in the dark, are characterized by very high values of viscosity, typical of those displayed by polyelectrolytes. The viscosity decreases dramatically upon exposure to sunlight and returns to the original value along with the reappearance of the absorption in the visible region. In order to correlate viscosity changes with conformational changes, samples of photochromic polypeptides were exposed to light, then the viscosity and the CD spectra were measured over time in the dark. Viscosity progressively increases with the gradual decrease of the helix content for both spiropyrancontaining poly(L-glutamic acid) (Figure 14.14) and poly(L-lysine) (Figure 14.15). The high viscosity of the solutions in the dark is essentially due to the side chains, which are charged when macromolecules are in disordered conformation. In these conditions the polypeptides are able to coordinate many solvent molecules to give highly solvated an extended macromolecules with a large hydrodynamic
t i m e , mi n Figure 14.14 PoIy(GIu) bearing 85mol% spiropyran units. a-Helix content ( ) and viscosity ( ) variation during decay in the dark at 25 0C. HFP solutions were irradiated, then dark adapted and monitored over time. The percentage of a-helix variation is estimated as indicated in Figure 14.13
a-helix variation,%
t i m e , min Figure 14.15 Helix content ( ) and viscosity ( ) variation during decay in the dark for poly(Lys) bearing 46 mol% spiropyran units, in HFP/NEt3 = 94/6
volume, thus exhibiting high values of viscosity. Aggregation phenomena, through interactions between merocyanine side chains, can also contribute to viscosity increases. From the figures it appears that viscosity keeps on increasing even when the a-helix is completely destroyed. In fact, the helix is fully destroyed by conversions of the spiro to the merocyanine form of « 60%, (Figure 14.13), but the macromolecules go on expanding until conversion to the merocyanine form reaches 100%.
14.3 REFERENCES [1] (a) M. Farina, Chim. Ind. (Milan), 1986, 68, 62; (b) M. Farina, Top. Stereochem., 1987,17,1. [2] P. Pino, Adv. Polym. ScL, 1965,4, 393. [3] F. Ciardelli, M. Aglietto and G. Ruggeri, in M. Fontanille and A. Guyot (Eds.), Recent Advances in Mechanistic and Synthetic Aspects of Polymerization, Reidel, Dordrecht, 1987, p. 409. [4] G. Natta, Makromol. Chem., 1960,35,94. [5] P. Pino, F. Ciardelli and G.P. Lorenzi, J. Polym. ScI, Part C, 1963, 4,21. [6] P. Pino, F. Ciardelli and M. Zandomeneghi, Annu. Rev. Phys. Chem., 1970, 21, 561. [7] F. Ciardelli and P. Salvadori, Pure Appl. Chem., 1985,57,931. [8] E.R. Blout, in F. Ciardelli and P. Salvadori (Eds.), Fundamental Aspects and Recent Developments in ORD and CD, Heyden, London, 1973, Chs. 4 and 5.
[9] P.L. Luisi and F. Ciardelli, in A.D. Jenkins and A. Ledwith (Eds.), Reactivity, Mechanism and Structure in Polymer Chemistry, John Wiley & Sons, New York, 1974, p. 471. [10] P.L. Luisi and P. Pino, J. Phys. Chem., 1968,72,2400. [11] P. Pino, F. Ciardelli, G.P. Lorenzi and G. Montagnoli, Makromol. Chem., 1963,61, 207. [12] F. Ciardelli, P. Salvadori, C. Carlini and E. Chiellini, J. Am. Chem. Soc, 1972, 94, 6536. [13] W. Hug, F. Ciardelli and I. Tinoco, Jr, J. Am. Chem. Soc, 1974,96, 3407. [14] F. Ciardelli, C. Righini, M. Zandomeneghi and W. Hug, J. Phys. Chem., 1977, 81, 1948. [15] F. Ciardelli and O. Pieroni, Chimia, 1980, 34, 301. [16] F. Ciardelli, O. Pieroni, A. Fissi and J.L. Houben, Biopolymers, 1984,23, 1423. [17] A. Fissi, O. Pieroni and F. Ciardelli, Biopolymers, 1987, 26,1993. [18] O. Pieroni, D. Fabbri, A. Fissi and F. Ciardelli, Makromol. Chem., Rapid. Commun., 1988,9,637. [19] B.R. Malcolm and O. Pieroni, Biopolymers, 1990, 29,1121. [20] O. Pieroni, A. Fissi, J.L. Houben and F. Ciardelli, J. Am. Chem. Soc, 1985,107,2990. [21] M.M. Long and D.W. Ury, in E. Grell (Ed.), Membrane Spectroscopy, SpringerVerlag, Berlin, 1981, pp. 143-171. [22] P. Bayley, in S.B. Brown (Ed.), An Introduction to Spectroscopy for Biochemists, Academic Press, London, 1980, pp. 148-234. [23] A. Fissi and O. Pieroni, Macromolecules, 1989,22,1115. [24] F. Ciardelli, D. Fabbri, O. Pieroni and A. Fissi, J. Am. Chem. Soc, 1989, 111, 3470. [25] O. Pieroni, A. Fissi, A. Viegi, D. Fabbri and F. Ciardelli, J. Am. Chem. Soc, 1992,114, 2734. [26] N.W. Tyer, Jr, and R.S. Becker, J. Am. Chem. Soc, 1970,92,1289. [27] T.M. Cooper, K.A. Obermeyer, L.V. Natarajan and R.L. Crane, Photochem. Photobiol., 1992,55,1. [28] A.S. Kholmanskii and K.M. Dyumaev, Russ. Chem. Rev. (Engl. TransL), 1987, 56, 136. [29] WJ. Middleton and R.V. Lindsey, Jr., J. Am. Chem. Soc, 1964,86,4948. [30] (a) J.R. Parrish, Jr., and E.R. Blout, Biopolymers, 1971,10,1491; (b) R.W. Woody, J. Polym. ScL, Macromol. Rev., 1977,12,181. [31] (a) M. Satoh, Y. Fujii, F. Kato and J. Komiyama, Biopolymers, 1991,31, l;(b) R.F. Epand and H. Scheraga, Biopolymers, 1968, 6,1383. [32] B.F. Erlanger, Annu. Rev. Biochem., 1976,45,267.
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 luminescence, 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 photoemission 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 relaxation of chains and the viscosity of the microenvironment of the probe molecule. The study of luminescence intensity of probe molecules as a function of temperature has been used as a method of studying transition temperatures and subgroup motion in polymers. (c) Luminescent species in polymer photooxidation: the problems associated with establishing a mechanism for the photooxidation and weathering of synPolymer Spectroscopy. Edited by Allan H. Fawcett © 19% John Wiley & Sons Ltd
S-S absorption
lntersystem crossing
Fluorescence
Vibrational relaxation
Absorption
lntersystem crossing
lntersystem crossing Vibrational relaxation
Phosphorescence
Vibrational relaxation
T-T absorption
Internal conversion
Internal conversion
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 spectroscopy, luminescence methods have been employed. (d) Identification of polymers: luminescence spectroscopy can provide a convenient 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 structure 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 mechanical 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 identical 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 distinct 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 transition 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 relaxations, 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 UV-ABSORBER
«'•
RADICAL SCAVENGER
*•
RO #
ROO,
MOLECULAR DISSOCIATION
QUENCHER
ROOH
METAL 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 molecules 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, excimer, N 2 pulse duration
Typical picosecond laser pulse duration
Shortest laser pulse 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 separator Cavity dumper
KTP rfout sync Cavity dumper Mode-locker driver out driver sync out
KDP
Sample
Fast photodipde
rfout
Timing filter amplifier
Harmonic separator
Constant fraction timing discriminator CFTD
Filter
MicroChannel plate
CFTD
PC/AT computer
TAC/SCA
Multichannel analyser
X100 Amplifier
Time-to-amplitude converter/single channel analyser
Figure 15.8 Time-correlated single-photon counting spectrometer based on CW modelocked Nd: Y AG laser
subject to environmental influence, such as diffusion, energy transfer and migration, 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 counting [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, timeresolved 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*] = kD^ll.^*\cxp(-
A l t )exp(- A2t)]
(D (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
(4)
2
1/2
(5)
A1 = 1/2(Z + kD + kM - {(kD + fcM - X) + 4fcMD/cDM[M]} ) A2 = 1/2(Z + kD 4- /cM + {(kD + fcM - X) +4fcMD/cDM[M]} )
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 k(t) = 4TIDABR'N(
1+
R
)
(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 approximately 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 approach 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 necessarily 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 fluorescence 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 concentration 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 =
-) 2 exp(— -q ['2WU0(2W T)+ I1VWT)] exp(-2WT) \
T
<1T)
Jo
/
(12) 1
2
*E(0 =
I1(HVT)IdT)
-qFEkE(kM-kE)(l-q)2[texp(-u(kE-T-1)Jo V
— -q[t "2W T Jo
x exp(-2WT)lI0(2WT) +I^WTftdTjdu
(13)
We have tested some of the above models using data from careful timeresolved 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.
[
AQ
-(kM + krJt—^-J
/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
Temp (K)
*2(FF)
290 270 250 230 210
U5 1.11 1.09 0.99 1.26
2 112 x (t )
1.11 1.05 1.09 1.06 1.30
Tw (FF)/
Tw (r 1/2 )/
kM +1/2kTOt 4(1-2/7T)(r )/ q2W/
10 7 S" 1
10 7 S- 1
(xlO" 4 )
10 7 S" 1
O20 0.12 0.049 0.041 0.053
019 0.12 0.047 0.038 0.046
115 116 114 1.95 1.71
O30 0.18 0.07 0.06 0.08
/cM + /crot (FF)/ 10 7 S" 1 236 130 118 1.99 1.76
Table 15.2 Fitting parameters for actual and synthesized data of Tao and Frank; 2-vinylnaphthalene homopolymer Temperature/K 293 273 253 233 213 193 173 153 133 113
* 2 (FF)
* 2 (' 1/2 )-X 2 (FF)
4(1-2/Tr)^2WyIO7S-1
ikM +Jk n ^lO 7 S" 1
L29 1.18 1.10 1.10 1.08 1.05 1.06 1.05 1.02 1.03
017 0.11 0.05 0.03 0.02 0.01 <0.01 <0.01 <0.01 <0.01
15 1.9 1.0 0.62 0.44 0.24 0.09 0.08 0.02 0.02
11 3.1 2.8 2.4 1.9 1.6 1.5 1.5 1.4 1.4
fcM + kTOV until the condition that 4(1 — 2/^)9 W < WkM + fcrot is no longer satisfied 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 experimental 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 approximates 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 polarization of excitation: Ht) = [Z1W - IAt)WiIt)
+ 2Z1(O] = D(t)/S(t)
Time-resolved fluorescence anisotropy measurements [47] can provide detailed information on the reorientation dynamics of molecules in solution. Until recently, however, this information has been limited to single rotational correlation times, which are only strictly appropriate for the diffusion of spherically symmetric molecules. Improvements in instrumentation and data analysis techniques during the last decade have led to increasingly accurate measurements of fluorescence lifetimes, with parallel improvements in determinations of fluorescence 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., wobbling within a cone), and the segmental motion of synthetic macromolecules [8]. Analyses of these models point to experimental situations in which the anisotropy 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
PMA/ACE
PMA/VN
PMMA/ACE
PMMA/VN
T/K
Tf/ns
To
TR
298 ±2 260±2 245±2 230±2 289±2 275±2 260±2 245±2 230±2 298±2 275±2 260±2 245±2 23O±2 298±2 275±2 260±2 245±2 23O±2
17.4 ±0.2 17.4±0.3 17.4±0.3 17.5±0.3 15.1 ±0.1 14.9±0.1 14.8±0.1 14.9±0.1 14.9±0.1 15.5±O.l 15.7±O.l 15.4±0.2 15.5±0.2 15.6±0.1 15.9±0.02 15.6±0.2 15.5±O.l 15.4±0.1 15.4±0.1
0.10 ±0.01 0.10±0.01 0.11 ±0.01 0.12±0.02 0.13±0.01 0.13±0.01 0.14±0.01 0.14±0.01 0.15±0.01 0.13±0.01 0.13±0.01 0.13 ±0.01 0.13±0.01 0.11 ±0.02 0.15±0.01 0.16±0.01 0.14±0.01 0.15±0.01 0.16±0.01
0.8 ±0.3 1.3±0.2 1.8±0.3 2.5±0.3 0.5±0.1 0.8±0.1 1.0±0.2 1.3±0.3 1.7±0.3 1.3±0.1 2.2±0.2 3.2 ±0.5 4.5±0.7 5.6±0.7 1.3±0.2 2.2±0.5 2.7±0.3 3.6±0.5 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. Irradiation at 300 nm produces excitation of both absorption bands, and so naphthalene, 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 chromophore, would appear to be capable of motion independent of the polymer backbone 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 Molecular Motion in Synthetic Polymers, Chapman Hall, London, 1985.
[2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27]
S.W. Beavan, J.S. Hargreaves and D. Phillips, Adv. Photochem. 1979,11, 207. Y. Nishijima, J. Polym. ScL Polym. Symp., 1970, 31, 353. A.C. Somersall, E. Dan and J.E. Guillet, Macromolecules, 1974,7, 233. D. Phillips, Br. Polym. J., 1987,19,135. 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. 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. 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. D. Phillips, in CE. Hoyle and J.M. Torkelsen (Eds.), Photophysics of Polymers, ACS Symp. Ser., 1987, (358), 308. J.B. Birks, Photophysics of Aromatic Molecules, Wiley-Interscience, London, 1970, pp. 322-335. J.B. Birks, DJ. Dyson and T.A. King, Proc. R. Soc. London, Ser. A, 1964, 277, 270. AJ. Roberts, D.V. O'Connor and D. Phillips, Ann. N.Y. Acad. ScL, 1981,366,109. D. Phillips, AJ. Roberts and I. Soutar, Polymer, 1981, 22, 293. D. Phillips, AJ. Roberts and I. Soutar, J. Polym. ScL, Polym. Phys. Ed., 1982,20,411. D. Phillips, AJ. Roberts and I. Soutar, Polymer, 1981, 22,427. D. Phillips, AJ. Roberts and I. Soutar, J. Polym. ScL, Polym. Phys. Ed., 1980, 18, 2401. D.A. Holden, P.Y.K. Wang and J.E. Guillet, Macromolecules, 1981,14,405. F.C. DeSchryver, K. Demayer, M. Van der Anweraer and E. Quanten, Ann. N.Y. Acad. ScL, 1981, 109. AJ. Roberts, D. Phillips, F. Aboul-Rasoul and A. Ledwith, J. Chem. Soc, Faraday Trans. 7,1981,77,2725. K. Sienicki and G. Durocher, Macromolecules, 1991, 24,1102. C. David, M. Piens and G. Geuskens, Eur. Polym. J., 1972,8,1019. C David, M. Piens and G. Geuskens, Eur. Polym. J., 1972,8, 1291. CE. Hoyle, T.L. Nemzek, A. Mar and J.E. Guillet, Macromolecules, 1978,11, 429. G.E. Johnson, J. Chem. Phys., 1975,62,4697. J.C. Andre, F. Baros and M.A. Winnick, J. Phys. Chem., 1990, 94, 2942. T.C. Nemzek and W.R. Ware, J. Chem. Phys., 1975, 62,477. M.V. Smoluchowski, Z. Phys. Chem., 1917,92, 129.
[28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49]
F.C. Collins and GE. Kimball, J. Colloid ScL, 1949,4,425. D.L. Huber, Phys. Rev. B, 1979, 20, 2307. B. Movaghar, G.W. Sauer and D. Wurtz, J. Slat. Phys., 1982, 27, 473. B. Movaghar, J. Phys. C (Solid State), 1980,13,4915. E.W. Montroll, J. Math. Phys., 1969,10, 753. J. Klafter and R. Silbey, J. Chem. Phys., 1981,74, 3510. CR. Gochanour, H.C. Andersen and M.D. Fayer, J. Chem. Phys., 1970,70,4254. R.F. Loring, H.C. Andersen and M.D. Fayer, J. Chem. Phys., 1982,76, 2015. G.H. Fredrickson and H.C. Andersen, Macromolecules, 1984,17, 54. M. Griinewald, B. Pohlmann, B. Movaghar and D. Wurtz, Philos. Mag. B, 1984,49, 341. R. Richert, B. Ries and H. Bassler, Philos. Mag. B, 1984, 49, L25. A.D. Stein, K.A. Petersen and M.D. Fayer, J. Chem. Phys., 1990,92, 5622. W. Weixelbaumer, J. Burbaumer and H.F. Kaufmann, J. Chem. Phys., 1985,83,1980. K. Sienicki and M.A. Winnick, J. Chem. Phys., 1987,87, 2766. M.N. Berberan-Santos and J.M.G. Martinho, J. Chem. Phys., 1991,95,1817. J. Baumann and M.D. Fayer, J. Chem. Phys., 1986,85,4087. G.H. Frederickson and CW. Frank, Macromolecules, 1983,16, 572. W.C Tao and CW. Frank, J. Phys. Chem., 1989,93, 776. R. Gelles and CW. Frank, Macromolecules, 1982,15, 747. D. V. O'Connor and D. Phillips, Time-Correlated Single Photon Counting, Academic Press, London, 1984. R.L. Christensen, R.C Drake and D. Phillips, J. Phys. Chem., 1986,90, 5960. R.C Drake, Ph.D. Thesis, University of London, 1986.
Index Index terms
Links
A AB quartet
18
acrylonitrile-furan copolymers
27
adsorbtion
242
afine deformation
183
aggregation
355
alignment (Homeotropic, planar)
282
alkylcyanobiphenyl
282
alpha (shift) effect
339
9
56
alpha (greek) process
277
alpha helix
355
amorphous polymers
276
332
anisotropies
173
235
APT (attached proton test)
11
autocorrelation function
278
azobenzene groups
351
B Bernoullian statistics beta (shift) effect
21 9
beta (greek) process
277
biaxial orientation
176
blends
245
57
340
This page has been reformatted by Knovel to provide easier navigation.
391
392
Index terms
Links
C catalysts
33
C– –C(triple) stretch
214
CD Circular dichroism
350
chain transfer
32
charge-transfer complex participation
22
chiral macromolecules chemical shift chemical shift image cis double bonds
347 2 157
159
36
Cobalt-60 gamma irradiation
263
Cole-Cole plot
290
composites
159
conformation
55
conformational data
97
conformational transitions
352
correlation times
235
COSY
221
221
84
coupling constant cross polarization Robin
7
111
135
crosslinked systems
12
Cryogenic trapping
270
crystalline polymers
280
261
D deformation
203
degradation
253
263
11
73
DEPT Assignment technique
This page has been reformatted by Knovel to provide easier navigation.
393
Index terms
Links
desorption
165
diacetylene monomers
204
dichroic difference
187
dichroic ratio
180
dielectric constant/permittivity
258
dielectric relaxation modes
279
Dielectric Relaxation Spectroscopy (DRS)
275
dienes as monomers
287
25
diffusion
162
167
diglycidyl ether of bisphenol-A (DGEBA)
121
288
domain size
125
146
double bond dyada, triads dynamic dichroic
38 192
E elastomers
197
electroactive
282
electron spin resonance (esr)
231
electrooptical
282
enantiomer copolymerization
41
ethylene glycol dimethacrylate
254
ethylene-vinyl acetate copolymers excimer
253
88 374
F Fourier transform (NMR) Fourier transform vibration spectroscopy fractal
8 173
257
15
free radical
231
255
This page has been reformatted by Knovel to provide easier navigation.
315
394
Index terms fumarate-vinyl acetate copolymer furan-acrylonitrile copolymers
Links 245 27
G g value (esr) gamma gauche effect
235
255
57
64
97
gamma rays
263
gamma relaxation
280
gels
259
311
332
glass transition temperature
142
165
240
Goldman-Shen pulse sequence
126
H helix content
354
367
I imaging (NMR) INEPT assignment technique
151 72
inhomogeneities
151
interfacial
224
interferometer
177
IR
173
isotope enrichment (D)
257
30
K Karplus
111
Kerr Constant
280
Kevlar
204
Kohlrausch-Williams-Watts
277
207
This page has been reformatted by Knovel to provide easier navigation.
395
Index terms
Links
L lamellar morphology Lewis acid effect
125 29
light scattering
297
linear response theory
278
liquid crystalline polymers
145
282
loss (dielectric)
275
283
luminescence
369
M Magic-angle Spinning (MAS) NMR Markov statistics
92 144 359
meso dyads
18
metallacarbenes
35
methacrylate radical
260
methylmethacrylate
254
41
8
18
55
97
molecular anisotropy
173
molecular size
331
morphology
136
Motion (of polymer)
231
Motionally-narrowed
239
multidimensional NMR
135
N natural rubber
138
21
merocyanine
microstructure
119
161
198
This page has been reformatted by Knovel to provide easier navigation.
32
396
Index terms
Links
near lR
257
neutron scattering
325
nitroxide
232
NMR
7
117
135
160
281
27
103
151 NOESY NMR assignment
87
Norris-Tromsdorf region
257
nylon
128
O optical activity
347
orientation
173
P permittivity(dielectric)
275
phase separating mixtures
305
photophysics
369
pixel
165
poly(acrylonitrile)
19
poly(alpha amino acids)
351
poly(alpha methyl styrene)
265
poly(bisphenylurethane-2,4-hexadiyne-1,6-diol)
205
poly(butadiene)
27
poly(but-l-ene sulphone)
16
poly(isobutylene)
199
polycarbonate
167
poly(diacetylene) single crystal fibre
205
poly(dimethyl siloxane)
161
poly(2,6-dimethyl-l,4-phenylene oxide)
196
This page has been reformatted by Knovel to provide easier navigation.
21
397
Index terms polyester(branched) poly(ethyl cyanoacrylate) poly(2-ethylhexyl methacrylate) polyethylene
Links 9
12
15 248 9
181
184
210
212
281
polyethylene terephthalate)
145 277
182 281
189
poly(L-glutamic acid)
352
poly(2,4-hexadiyne-1,6-bis (p-toluenesulphonate))
205
poly(isoprene)
246
poly(L-lysine)
351
polymer dynamics
136
275
338
polymerization
253
poly(methacrylonitrile)
107
poly(methyl acrylate)
277
poly(methylmethacrylate)
7 165 259
10 246 260
107 254 268
poly(n-butyl isocyanate)
235
poly(norbornene)
35
43
poly(oxymethylene)
140
143
281
polypropylene
10 103
58 122
66
68
277
190 267
193 315
10
92
poly(propylene oxide) polystyrene polystyrene(syndiotactic) poly(styrene block butadiene block styrene) polysulphide
196 15
This page has been reformatted by Knovel to provide easier navigation.
246 373
398
Index terms
Links
poly(vinyl acetate)
243
277
poly(vinyl alochol)
104
371
poly(vinyl chloride)
105
130
poly(vinyl fluoride)
83
281
poly(vinylidene fluoride)
81
142
246
297
309
334
radical
231
255
Raman Spectroscopy
173
203
regio selectivity
18
38
40
resolution enhancement
12
Ring Opening Metathesis Polymerization (ROMP)
30
RIS (rotational isomeric state)
62
98
109
183
281 poly(vinyl methyl ether)
141
powder pattern spiess
137
prochiral face
35
Q quasi elastic scattering
R
Rotor-synchronized MAS NMR
144
Rouse modes
338
S SALS (small angle light scattering)
298
SANS (small angle neutron scattering)
330
semicrystalline polymers
298
semidilute
313
silica absorbent
242
This page has been reformatted by Knovel to provide easier navigation.
399
Index terms
Links
silicone polymer
282
simulation (esr spectrum)
266
solid polymer NMR
117
solution NMR
135
7
specular reflectance
177
spin density
157
spin label
231
spin probe
231
spin relaxation (proton)
125
spirobenzopyran
357
steroselectivity
35
stereospecific polymerization
32
strain
185
203
stress
185
203
styrene-MMA copolymers
84
substituent effects
56
surfaces
316
333
T1
119
125
T1rho(ρ)
125
T2
125
T
tacticity
15
Tactic sequence
98
tensile stress
222
thermoset (epoxy amine)
288
time-resolved measurements
184
two-dimensional IR
192
two-frequency-addressing
282
153
This page has been reformatted by Knovel to provide easier navigation.
153
400
Index terms
Links
U uniaxial orientation
174
urethane-diacetylene copolymers
219
UV light
351
V vinyl acetate copolymer vinyl chloride copolymers voxel
279 67 155
W WAAS
209
Wigner rotation matrix
286
WISE NMR
141
WLF (Williams Landel Ferry)
241
Y Young's modulus
206
Z Zeigler-Natta polymerization
31
This page has been reformatted by Knovel to provide easier navigation.