2-Slow Relaxation Mode From Solutions
2-Slow Relaxation Mode From Solutions
& The Society of Polymer Science, Japan (SPSJ) All rights reserved 0032-3896/10 $32.00
www.nature.com/pj
INVITED REVIEW
In the past, many laser light-scattering experimental results revealed that besides the fast relaxation mode, there existed an
additional slow mode in semidilute solutions. This slow mode has been assigned to a variety of origins, but there has been
no clear and well-accepted explanation. As the polymer concentration increases, the slow relaxation mode persists in the
concentrated region, in melts and in gels in which polymer chains are crosslinked instead of entangled. The slow relaxation
mode has also been reported for charged macromolecules in aqueous and nonaqueous solutions. However, it is generally thought
to be different in nature from that observed in semidilute neutral polymer solution. In recent years, armed with novel solution
preparation methods and some specially designed polymers, we have reexamined the dynamics of polymer chains, especially
the slow mode, in semidilute neutral polymer solutions, dilute polyelectrolyte solutions and gels, which are reviewed here. Our
results suggest that the slow mode can be qualitatively considered as hindered motions of interacting chains even though the
nature of interaction can be very different; namely, from the weak segment–segment interaction in a less good solvent to strong
electrostatic interaction among polyelectrolyte chains, and even to chemical crosslinking inside gel networks.
Polymer Journal (2010) 42, 609–625; doi:10.1038/pj.2010.59
Keywords: anionic polymerization; laser light scattering; polyelectrolyte; polymer interaction; semidilute; slow relaxation mode;
sol–gel transition
1State Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences, Shanghai, PR China; 2Department of Chemistry, The
Chinese University of Hong Kong, Hong Kong, PR China and 3Hefei National Laboratory for Physical Sciences at Microsacle, Department of Chemical Physics, University of
Science and Technology of China, Anhui, PR China
Correspondence: Dr J Li, State Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences, Lingling Lu, Shanghai, PR China.
E-mail: [email protected] or Professor T Ngai, Department of Chemistry, The Chinese University of Hong Kong, Shatin. N.T., Hong Kong, PR China.
E-mail: [email protected]
Received 28 April 2010; revised and accepted 4 June 2010
The slow relaxation mode
J Li et al
610
remains an important controversial problem in polymer science, it intensity fluctuations can relate to the dynamic pair correlating of the
certainly deserves a careful reexamination so that we can find its scattering species:
physical nature. Moreover, advancements of LLS instrumentation, 2
2 2 ð1Þ
including the full digital time correlator and computer, have made dIq ð0ÞdIq ðtÞ ¼ jSðq; tÞj ¼ hSðqÞi 1+bapp gsolute ðq; tÞ ð1Þ
the dynamic study of slow relaxation mode much easier and reliable. ! 4pn y
In recent years, by designing, synthesizing and using some novel where q ¼ q ¼ l0 sin 2 with n, l0 and y are the solvent refractive
polymers or sample preparation methods, we have reexamined the index, the wavelength of the light in vacuum and scattering angle,
dynamics of slow relaxation mode in polyelectrolyte, semidilute respectively, b is the coherent factor depending on the detection
solutions and gelling systems by a combination of static and dynamic optics, S(q,t) is the dynamic scattering factor, S(q) is the structure
LLS.75,96–107 In this article, we mainly review what has been carried factor, g(1)(q,t) is the normalized field–field time correlation functions,
out in our laboratory during the past 10 years. We will outline the defined as:
details of each system in the following sections by starting with a brief hEð0; qÞE ðt; qÞi=hEð0ÞE ð0Þi; If t ! 1; g ð1Þ ðq; tÞ ¼ 0; and
introduction of LLS. Z
Sðq; t ¼ 1Þ ¼ SðqÞ ¼ exp½i~ q ð~
rmi ~
rnj ÞdNd~
r
LLS
In a broad definition, LLS could be grouped as inelastic (for example, For polymer in dilute solution without long-range pair interaction,
Raman, fluorescence and phosphorescence) and elastic (no adsorp- the structure factor can relate to molar mass and form factor of the
tion) light scattering. However, in polymer and colloid science, light scattering particle by the follow equation:
scattering is normally referred to in terms of static (elastic) or dynamic
q2 2
(quasi-elastic) measurements, or both.108 Static LLS as a classic and SðqÞ / Np NPðqÞ / CMð1 2CMA2 + . . .Þð1 R + . . .Þ ð3Þ
3 g
absolute analytical method measures the time-average intensity and it
has been long and widely used to characterize both synthetic and where Np, N, C, M, A2 and Rg are the number of polymer chain, the
natural macromolecules.109 On the other hand, dynamic LLS mea- number of monomer in each polymer, the concentration, molar mass
sures the intensity fluctuation. This is where the word dynamic comes of polymer, the second virial coefficient and the radius of gyration,
from. The visibility of the scattering objects (macromolecules or respectively. In static LLS, the excess absolute time-averaged scattered
colloidal particles) in LLS depends on the refractive index difference light intensity, known as the excess Rayleigh ratio Rvv(q), of a dilute
between the scattering object and the background (dn¼nn0). polymer solution at concentration C (g ml–1) is related to the weight
In the last two decades, owing to the advance of stable laser, ultra- average molar mass Mw, the root-mean square z-average radius of
fast electronics and personal computers, LLS, especially dynamic LLS, gyration /Rg2Sz1/2 (or written as /RgS), and the scattering vector q as:
has evolved from a very special instrument for physicists and physical KC 1 1
chemists to a routine analytical tool in polymer laboratories or even to ð1+ hR2g iq2 :::Þ+2A2 C :::: ð4Þ
Rvv ðqÞ Mw 3
a daily quality-control device in production lines. Commercially
available research-grade LLS instruments nowadays are normally cap- where K¼4p2n2(dn/dC)2/(NAl04) with NA and dn/dC being the
able of making static and dynamic measurements simultaneously for Avogadro number and the specific refractive index increment, respec-
studies of colloidal particles in suspensions or macromolecules in tively, This is the most basic equation in static LLS. With Rvv(q)
solutions as well as in gels and viscous media. In our laboratory, we measured over a series of C and q, one can obtain /RgS and A2,
have used a slightly modified commercial LLS spectrometer (ALV/DLS/ respectively, from the slopes of [KC/Rvv(q)]C-0 vs q2 and [KC/
SLS-5022f, ALV GmbH, Langen, Germany) equipped with a cylindrical Rvv(q)]q-0 vs C; and Mw from [Kc/Rvv(q)]C-0,q-0. The Zimm plot,
22-mW UNIPHASE He-Ne laser (l0¼632 nm, JDS Uniphase Corp., that is, KC/Rvv(q) vs (q2+kC) with k being an adjustable constant,
Milpitas, CA, USA) as the light source and a sensitive Avalanche Photo allows both q and C extrapolations to be made on a single grid.
Diode (APD) detector. The incident light beam was vertically polarized In dynamic LLS, the intensity–intensity time auto-correlation
with respect to the scattering plane and the intensity was regulated with function G(2)(q,t), defined as /I(q,0)I(q,t)S//I(q)S2, in the
a beam attenuator (Newport Corp., Mountain View, CA, USA, M-925B) homodyne mode or self-beating mode was measured, where t is the
so as to avoid localized heating in the light-scattering cuvette. It is delay time and /I(q)S is the time-average scattering intensity, that is,
worth mentioning that many reviews, books, proceedings and chap- the measured baseline. G(2)(q,t) is related to g(1)(t,q) by the Siegert
ters have been published on the details of LLS instrumentation and relation as:
theory. Herein, we will concentrate only on the relevant details. h 2 i
Gð2Þ ðq; tÞ ¼ A 1+bg ð1Þ ðq; tÞ ð5Þ
When a light beam is incident on a solution, and the solvent
refractive index is different from that of the solute (macromolecules or where A is the measured baseline. For monodispersed spherical
colloidal particles), the incident light is scattered by each illuminated scatters, |g(1)(q,t)| is theoretically represented as an exponential
macromolecule or colloidal particle in all directions.9–11 If all scatter- decay function:
ing species are stationary, the scattered light intensity at each direction
would be a constant, that is, independent of time. However, in reality, jg ð1Þ ðq;tÞj ¼ G expðGtÞ ð6Þ
all the solute in solution are undergoing constant Brownian motions, where G and G are the factor of proportionality and the line width,
which leads to both fluctuations of the scattered intensity pattern on respectively, and G¼1/t, the characteristic decay time representing the
the detection plane and the fluctuations of net scattered light intensity, rate of dynamic relaxation. For a polydispersed dynamic relaxation,
I(t), with time in sufficiently small detector area. The fluctuation rates equation (6) may be generalized as:
can be related to different relaxation processes such as translational
Z1
and rotational diffusions as well as internal motions of the macro- ð1Þ
g ðq; tÞ ¼ GðGÞetG dG ð7Þ
molecules. The faster the relaxation process, the faster the intensity
fluctuations will be. In the homodyne or self-beating system, the 0
Polymer Journal
The slow relaxation mode
J Li et al
611
where G(G) is called the line-width distribution and G(G)dG segment–segment interaction is ‘completely’ screened out in an
is the statistic weight of the dynamic relaxation, which possess line athermal solvent.116,118,119
width or related to the polymer distribution in dilute polymer Besides the fast relaxation mode, an additional slow relaxation
solution. mode has also been observed in the intensity–intensity time correla-
The Laplace inversion of each measured G(2)(q,t) can lead to tion functions of various semidilute polymer solutions.28–39 This slow
one G(t) on the basis of equations 5 and 7. For a diffusive relaxation, relaxation mode was assigned to different physical origins, such as the
G(1/t) is related to the translational diffusion coefficient D by reptation of the entire chain inside a ‘tube’ made of its surrounding
(G/q2)c-0,q-0-D, so that G(G) can be converted into a transitional chains,114,115,117 the scattering vector (q)-independent relaxation of a
diffusion coefficient distribution G(D) or further a hydrodynamic transient network,29,34,35 the q2-dependent translational diffusion of
radius distribution f(Rh) applying the Stokes–Einstein equation, large aggregates or even dust particles,32,33,37,38,112 and internal
Rh¼(kBT/6pZ)/D, where kB, T and Z are the Boltzmann constant, motions of large transient chain clusters.92 Later, it was realized that
the absolute temperature and the solvent viscosity, respectively. Note the reptation is not observable in dynamic LLS because of its nature
that Z is local viscosity. In general, we cannot conclude that the slower for homopolymer system.
mode should be caused by larger scattering object if the local viscosity Theoretically, Brochard and de Gennes40 attributed DC to the
is not uniform. osmotic modulus (Mp), the elastic modulus (M0) and the friction
We should note that the Laplace inversion is an ill-conditioned coefficient (z) as DC¼(Mp+M0)/zC. Adam and Delsanti34,35 allowed
problem because of the bandwidth limitation of photon correlation the transient gel network to relax and assigned the q-independent slow
instruments, some unavoidable noises in the measured time correla- mode with a viscoelastic nature and a characteristic time (tr). On the
tion function, and a limited number of data points. Sometimes we other hand, Wang et al.41–46 related the slow relaxation to the
used a combination of two stretch exponential functions to fit the viscoelasticity under the Y condition from the osmotic pressure
correlation function: fluctuation without using the concept of a transient gel network.
h i n o2 The cooperative diffusion and the viscoelastic effect are generally
Gð2Þ ðq; tÞA =A ¼ bapparent Af exp½ðt=tf Þdf +As exp½ðt=ts Þds mixed, depending on the frequency and a coupling parameter (b).
ð8Þ b¼(C/r)(VSVP)/VS, proportional to the difference between partial
where bapparentpb, an apparent coherent factor, Af and As are specific volumes of polymer and solvent (VSVP) and r is the
intensity contributions of the fast and slow modes, respectively, and solution density. The influence of longitudinal modulus on the
Af+As¼1. The details of the data analysis could be referred to the concentration fluctuation leads to a broad distribution.44–46 If b¼1,
articles.97–107,110,111 there exits only one stress-relaxation mode. DC becomes identical to
that predicted in the transient gel network model.40,114,115
The slow mode in semidilute neutral polymer solutions Experiments around the Y condition indicate that the solvent
When the concentration of polymer is higher than the overlap quality has a profound effect on the slow relaxation.48–54 Actually,
concentration (C*), polymer chains start to ‘touch’ each other and not everyone accepts or recognizes such a slow mode, even though it
the solution enters the semidilute regime. C* has been differently has been repeatedly observed for more than three decades. In the last
defined as 3M/(4pNARg3), or M/(23/2NARg3), and [Z]1, where M, Rg, two decades, advancements of LLS instrumentation, including the full
NA and [Z] are the molar mass, the radius of gyration of polymer digital time correlator and computer, have made the dynamic study of
chains, the Avogadro constant and the intrinsic viscosity, respectively. semidilute solutions much easier and more reliable, but the prepara-
The difference of C* calculated based on these definitions could be as tion of dust-free viscous semidilute solutions still remains an experi-
large as five times.112,113 On the other hand, at the entanglement mental challenge.
concentration (Ce), polymer chains start to interwind with each other, Traditionally, viscous dust-free semidilute solutions are prepared by
where Ce can be estimated from the ratio of the chain length (L) and slow evaporation of solvent from a dust-free dilute solution (the
the entanglement length (Le).114 Ce is higher than C*. Note that for increase of polymer concentration). Figure 1 shows how the polymer
chains shorter than Le, they are not able entangle with each other even concentration (C) of four polystyrene (PS) standards increases with
at higher concentrations. the evaporation time (t). The inset shows that the time-average excess
The mean-field theory can effectively describe some properties of scattering intensity (/ISq-0¼/ISsolution–/ISsolvent) decreases as C
polymer solutions over the whole concentration range, but fails to increases, following a scaling of /ISq-0pC0.25±0.05. For a polymer
explain other static and dynamic properties because of strong intra- solution, /ISq-0p(kBT/V)C/(qp/qC), where V is the volume
and interchain interaction among covalently bonded monomers. and qp/qC is the osmotic compressibility. In dilute solution,
Using the concept of a ‘blob’, the reptation and tube model, de qp/qC¼NAkBT/M and then /ISq-0pC. In semidilute solutions,
Gennes and colleagues developed some scaling laws to predict various qp/qCpC5/4 so that /ISq-0pC1/4. The scaling in the inset agrees
properties of semidilute solutions.114–117 well with the prediction and shows that solvent evaporation gradually
In a transient gel model, there exists only one characteristic length induces the dilute polymer solution to the semidilute regime. As
or one dynamic process in semidilute solutions if the solvent is expected, only one fast relaxation mode is observed over the entire
thermodynamically good or athermal. In other words, the thermal concentration range, consistent with some of the previous
energy (kBT) can only agitate one short segment of the chain so that studies.112,116 However, it is worth mentioning that the whole process
its gravity center undergoes a random Brownian motion inside a lasted for about 1 year. Even using such a painful process, one still
confined volume. The static and dynamic correlation lengths (xS and faces a concentration gradient problem, that is, the upper layer is likely
xD) can be measured from the angular dependence of the average more concentrated than the bottom layer.
scattering intensity and the average cooperative diffusion coefficient On the basis of the definition of C*, one can also change Rg or M to
(DC), respectively. Previous studies are well represented by two master switch a solution from dilute to semidilute for a fixed polymer
curves: xSBC0.72±0.01 and xDBC0.70±0.01, slightly deviated from concentration. One interesting advantage of the in situ change of Rg
xBC0.75 predicted in theories.28,112 For a distance longer than x, the or M or both inside an LLS cell is that the polymer solution can
Polymer Journal
The slow relaxation mode
J Li et al
612
0.18 100
101
<I >q→0 ~ C -0.25 ± 0.05
<I>q→0 (a.u.)
0.12
[G (2)(t) - A]/A
C / g ml –1
20 the delay time (t) with T and the solvent viscosity, which is expected
because benzene is an athermal solvent for PS in the temperature
Γ / 105 s-1
15
range studied.
10 θ° With the same prepared method, we further studied the solution
10-2 15
dynamics of PS in a less good/y solvent, cyclohexane. As expected,
5
60
0
there is only one fast relaxation mode in cyclohexane when the
90
0 2 4 6 8 10 solution is dilute, independent of the scattering angle. As soon as
120
q2 / 1010 cm-2 the solution becomes semidilute, that is, C/C*41, an additional slow
10-3 relaxation mode appears in the measured intensity–intensity time
103 104 105 correlation function. The analysis of such a time correlation function
tq 2 /s cm–2 on the basis of equation (8) leads to two characteristic relaxation times
(/tSf and /tSs). Figure 3 shows that the slow mode becomes more
Figure 2 Scattering vector dependence of intensity–intensity time correlation
functions of PS in benzene, where delay time (t) is scaled with q2. The inset
obvious at smaller scattering angles. Note that in LLS, the observation
shows scattering vector dependence of average line width. Reproduced with length is directly proportional to 1/q, that is, large objects are more
permission from Li et al.75 Copyright (2008) American Chemical Society. obvious at smaller scattering angles. Also note that here C/C* is only
B1.4 at 45 1C and C/CeB0.1 so that most of the chains are not
entangled with each other. In a less good solvent, those entangled
completely avoid possible problems of dust, solubility and inhomo- chains behave like large transient clusters within the delay time
geneity. To achieve that, we used anionic living polymerization75 to window in dynamic LLS due to relatively strong segment–segment
prepare dust-free semidilute/concentrated solutions (up to 30%) of interaction in comparison with in an athermal solvent. Moreover, we
narrowly distributed PS chains directly inside the LLS cell with a high found that the characteristic line width (/GsS) is scaled to q as
vacuum stopcock, starting from a dust-free monomer solution. /GsSBq3, indicating that the slow mode reflects some internal
Subsequently, we comparatively reexamined the solution dynamics motions of large scattering objects. The concentration dependence
of PS chains in both good solvent (benzene) and a less good/y solvent of the scaling exponent (as) between /GsS and q will be discussed
(cyclohexane). more in the follow section.
For the solution studied in Figure 2, we terminated living PS chains As the solvent quality also affects the solution dynamics, we studied
and measured its weight average molar mass (Mw) after the end of the temperature dependence of the slow relaxation mode in cyclo-
polymerization. We found that Mw is B5.3105 g mol–1 and the hexane. Figure 4 shows typical plots of the temperature dependence of
polymer concentration is 0.18 g ml–1, that is, C/C*B30. Importantly, intensity–intensity correlation function for the PS in cyclohexane, and
Figure 2 shows that even with such a high concentration, there is still Figure 5 shows the corresponding plots of characteristic relaxation
only one fast relaxation mode, agreeing well with the prediction in an time distribution function analyzed by the CONTIN method. It clearly
athermal solvent. The scaling of the delay time (t) with q2 makes the reveals that at a low scattering angle, the decreases of the solution
measured time correlation functions at different scattering angles temperature makes the slow mode less obvious and results in a
collapse into one curve. The inset shows that the average line width decrease of its intensity contribution (the peak area). At 37 1C, the
Polymer Journal
The slow relaxation mode
J Li et al
613
100 100
θ°
15
30
60
10-1
[G (2)(t) - A]/A
[G (2)(t) - A]/A
90
10-1 PS in cyclohexane = 15° PS in cyclohexane T = 50.2 °C 150
M w = 4.6 x 104 g mol –1 T / ˚C Mw = 5.5 x 105 g mol–1
M w/M n = 1.07 49.1 M w /Mn = 1.08
C = 0.091 g ml –1 40.8 10-2 C = 0.14 g ml–1
C/C * = 1.4 at 45 °C 37.0 C/C* = 12 at 45 °C
10-2
10-3
103 104 105 106 107 108
10-4 10-3 10-2 10-1 100 101
tT/ / sK/cp tq 2/s cm–2
Figure 4 Temperature dependence of intensity–intensity time correlation Figure 6 Scattering vector dependence of intensity–intensity time correlation
functions of PS in cyclohexane, where delay time (t) is scaled with solution functions of PS in cyclohexane, where delay time (t) is scaled with q2.
temperature (T ) and solvent viscosity (Z). Reproduced with permission from Reproduced with permission from Li et al.75 Copyright (2008) American
Li et al.75 Copyright (2008) American Chemical Society. Chemical Society.
1.2 100
θ°
30
0.9 60
PS in cyclohexane θ = 15°
90
T / ˚C 10-1
104 mol–1
[G (2)(t) - A]/A
Mw = 4.6 x g 120
49.1 150
G()
Mw/Mn = 1.07
0.6 40.8
C = 0.091 g ml–1
37.0 PS in cyclohexane T=49.9 °C
C/C* = 1.4 at 45 °C
10-2 M w = 1.9 x 106 g mol–1
0.3 M w/M n = 1.13
C = 0.22 g ml–1
C /C * = 51 at 45 °C
0.0
10-3
100 101 102
103 104 105 106 107 108
T//sK/cp 2 /s cm–2
tq
Figure 5 Temperature dependence of characteristic relaxation time
Figure 7 Scattering vector dependence of intensity–intensity time correlation
distributions G(t) of PS in cyclohexane, where characteristic relaxation time
functions of PS in cyclohexane (Solution 10 in Li et al.75), where delay time
(t) is scaled with solution temperature (T ) and solvent viscosity (Z).
(t) is scaled with q2. Reproduced with permission from Li et al.75 Copyright
Reproduced with permission from Li et al.75 Copyright (2008) American
(2008) American Chemical Society.
Chemical Society.
slow mode completely disappears in the whole angular range and only
one fast mode is observable. As discussed before, the decrease of the independent of the position of the scattering volume inside the
solution temperature makes the solvent quality poorer and the solution. Figure 7, once again shows that the fast mode collapses
segment–segment interaction stronger so that PS chains contract as together after delay time (t) is scaled by q2, indicating that it remains
the temperature decreases, resulting in an increase of C*, which can diffusive. At the same time, Figure 8 shows that for a given scattering
switch a solution from semidilute to dilute for a given C, that is, from angle, the slow mode become slower and its intensity contribution
C/C*41 to C/C*o1 when C/C* is not higher at higher temperature. (As) increases as the solution temperature decreases. Note that the
Figure 6 shows that for a more concentrated solution, that is, decrease of the solution temperature has two opposite effects on the
C/C*¼12 or more precisely C/Ce41, the fast mode is collapsed slow mode. On the one hand, it increases the segment–segment
together when the decay time (t) is scaled by q2, revealing its diffusive interaction and enhances the slow mode; on the other hand, the
nature. In contrast with Figure 3, here the slow mode becomes more chain contraction disentangles the transient clusters. When the latter is
obvious at larger angles, which is different from those observed in the dominant, the slow mode becomes weak or even disappears at lower
above lower concentration region. It implies that the slow mode is not solution temperatures. However, when C is much greater than Ce, the
related to some large scattering objects. Further increase of the chain contraction cannot switch the solution from semidilute to dilute
polymer concentration makes the slow mode even visible at low so that decreasing the solution temperature only increases the seg-
scattering angles, as shown in Figure 7. Here, C/C*450 and the ment–segment interaction. This presumably is the reason why the
solution appears like a gel without any visible flow within a short time, slow mode becomes slower and more visible at lower temperatures as
but the solution is homogeneous and its scattering intensity is shown in Figure 8.
Polymer Journal
The slow relaxation mode
J Li et al
614
C/C* = 51 at 45 °C
T / °C
10-2 49.9
45.5
38.9
37.0
10-3
3
s
f
2
Polymer Journal
The slow relaxation mode
J Li et al
615
When C4Ce, all the chains inside the solution entangle with each
other to form one huge ‘cluster’. Each chain could be visualized to
arrest inside a ‘tube’ made of its surrounding chains within a short
time and ‘divided’ into a number of segments (blobs) with a dimen-
sion limited by the tube diameter (the correlation length). Within each
blob, the segment is excited by thermal energy (kBT) so that its center
of gravity undergoes a Brownian motion. These thermal blobs have a
similar size. In an athermal solvent, there is no interaction between
them and the tube so that they experience a similar microenviron-
ment, resulting in only one (fast) relaxation mode. In a less good or
y solvent, the segment–segment interaction near the entanglement
points is expected to be stronger. Therefore, the blobs near each
entanglement point should move slower than those in the middle
between two entanglement points; namely, the blobs experience two
different microenvironments, as schematically shown in Scheme 1b.
The increase of polymer concentration makes the tube thinner. Up to
one point, the motions of the segments perpendicular to the tube near
the entanglement point is so limited that the segments can only
randomly oscillate in the tube. This might explain why its character-
istic relaxation time becomes less dependent on or even independent
of the scattering vector.
The temperature dependence studies also support our speculation
about the nature of the slow mode. The decrease of the solution
temperature makes the segment–segment interaction stronger so that
each chain contracts. When C/Ceo1, the chain contraction disen-
tangles the transient clusters so that the slow mode becomes weak or
even disappears at lower solution temperatures. However, when
C4Ce, the chain contraction cannot switch the solution from
semidilute to dilute so that decreasing the solution temperature only
increases the segment–segment interaction. This is why the slow mode
becomes slower and more visible at lower temperatures.
The slow mode in PNIPAM/H2O system Figure 10 (a) Concentration dependence of intensity–intensity time
The solution dynamics in poly(N-isopropylacrylamide) (PNIPAM)/ correlation function [G(2)(t,q)A]/A and (b) their corresponding characteristic
H2O system was selected and examined by dynamic LLS because of its relaxation time distribution function G(t) of PNIPAM/H2O in both dilute
thermal-responsive properties, which exhibits a lower critical solution and semidilute regimes. Reproduced with permission from Yuen et al.96
Copyright (2006) American Chemical Society.
temperature around 32 1C. In such a way, similar to the PS in
cyclohexane, we are able to use temperature dependence of the
solvent–polymer interaction to investiagte effects of the shrinkage of
polymer chains with increasing temperature, and consequently disen- and enters the semidilute regime, the chains begin to overlap and
tanglement of transient network, which in turn can prove the coex- entangle, and new dynamical processes involving interchain interac-
istence of individual chains and large transient network in semidilute tion and disentanglement begin to occur besides the fast relaxation
solutions when C/Ce o1. mode. The slow mode becomes more and more evident, and its
For PNIPAM in semidilute aqueous solutions, Hirotsu and co- characteristic relaxation time shifts toward slow direction with increas-
workers120 found that the correlation function measured by dynamic ing concentrations.
LLS could be fitted to single exponential quite well through the whole Figure 11 shows the temperature dependence of characteristic
temperature range investigated (25–40 1C). They claimed that the relaxation time distribution function obtained from a semidilute
translational motion of polymer molecules is profoundly inhibited solution. The data on the x axis is shifted by (T/Z) so that the peak
due to the relative immobility of densely packed coils and the positions at different temperatures can be directly compared, where Z
intermolecular entanglement effect; the correlation function represents is the solvent viscosity at the absolute temperature. It is clear that the
the relaxation of the internal or pseudogel mode. However, our recent fast mode shifts to the slower direction, whereas the slow mode
result found that the single relaxation mode measured from dilute becomes faster with increasing temperatures when To30.0 1C. The
PNIPAM aqueous solution by dynamic LLS turns into one fast and effects are just contrary to that of increasing concentrations. The
one slow relaxation mode when the concentration enters into semi- solvent quality tends to be poor with increasing temperature, which
dilute regime. Figure 10a shows typical plots of the intensity–intensity will result in fewer entanglement points and, consequently, a loose
correlation function for the PNIPAM/H2O system in both the dilute network with more defects or larger structural fluctuation. However,
and semidilute regimes, and Figure 10b shows the corrsponding plots at T¼30.0 1C, phase separation happens. After keeping the solution at
of characteristic relaxation time distribution function analyzed by the this temperature for 32 h, the characteristic relaxation time distribu-
CONTIN method. When CoC* (C*¼1.9–2.8103 g ml–1 at 25 1C), tion function turns out to be single narrow mode, which is q2
a single relaxation mode is observed, which is related to the mutual dependent and related to collective diffusion of aggregates. The
diffusion of individual polymer chains. As the concentration increases temperature dependence studied for PNIPAM/H2O further supports
Polymer Journal
The slow relaxation mode
J Li et al
616
our speculation about the nature of the slow mode that for C/Ceo1, it
is related some large structures.
Polymer Journal
The slow relaxation mode
J Li et al
617
with that in PNIPAM system. With increasing temperature, it is clear manner so that we are able to investigate the effect of crosslinking on
that an intermediate mode appears accompanied by the disappearance the fast and slow relaxation modes of semidilute solutions at each
of the slow mode. The intermediate mode is evident at high tem- photo-reaction stage of the sol–gel transition.
perature, and its relaxation rate is q2 dependent; therefore it can be Figures 14 shows that in the initial semidilute solution before the
interpreted as collective diffusion of micelles. photo-crosslinking, g(2)(q,t) has two distinct relaxation modes. The
In summary, the special thermal-sensitive properties of PNIPAM/ fast relaxation corresponds to the cooperative diffusion of the sub-
H2O and PEO-PPO-PEO/H2O systems enable us to follow the chains between two entangled points, which is well described by the
temperature-induced shrinkage of polymer chains and, consequently, ‘blob’ or scaling theory,116,117 whereas the controversial slow relaxation
disentanglement of transient network in semidilute solutions formed can be attributed to the relatively stronger segment–segment interac-
by overlapped or entangled polymer chains, without changing the tion near the entanglement points makes the congested polymer chain
overall concentrations. The difference of the nature of slow mode nonuniform.99,102,103 Previously, we also speculated that such slow
between these two systems is that aggregates are formed in the mode might be related to some long-range density fluctuation.102 As
PNIPAM/H2O system but micelles are formed in the PEO-PPO- the crosslinking proceeds, the slow relaxation becomes even slower,
PEO/H2O systems when the solvent becomes poor. and at the same time, its contribution to the total scattering intensity
increases, reflecting in the raise of its amplitude. It clearly shows that
Effect of crosslinking on dynamics during the sol–gel transition the crosslinking enhances the restricted movement of the subchains
Up to this point, our results clearly show that the slow relaxation near the entanglement points, resulting in the slow mode. The flowing
observed in the semidilute solutions is real, not because of previously
suggested artifacts in the solution preparation. The concentration and
temperature dependence of the characteristic relaxation time of slow T = 40 °C = 30°
mode in different systems also suggest that such a slow mode has two 0.8
different natures. When C is lower than the entanglement concentra-
t UV = 14 h gel
tion (Ce), it can be related either to the large temporal clusters formed
0.6
by some entangled chains. However, at higher entanglement concen-
g(2)(q,)
CH3 CH3
( CH2 C ) ( CH C )
m 2 n
C O C O
CH3 CH3
O OCH3
( CH2 C )m ( CH2 C )
n O
O C O O
C = 310 nm CH3
O OCH3 CH3
O
CH3
O
O OCH3
O
O
C O C O
P(MMA-co-AMC)
( CH2 C ) ( CH2 C )
m n
CH3 CH3
Scheme 2 Photodimerization-induced crosslinking of P(MMA-co-AMC) copolymer chains in chloroform. Reproduced with permission from Ngai et al.103
Copyright (2003) American Chemical Society.
Polymer Journal
The slow relaxation mode
J Li et al
618
2.8
t UV = 0 h T = 40 °C tUV = 0 h, semidilute
0.3
0.2
tUV = 6 h, pre-gel
G()
1.4
0.1
0.7
tUV = 14 h, gel
0.0
0.0
10 15 20 25 30 35 40
10-5 10-4 10-3 10-2 10-1 100 101
T / °C
/s
Figure 16 Temperature dependence of intensity weighting (relative
Figure 15 Peak area-normalized intensity distribution of characteristic decay
contribution) of fast relaxation (Af) of crosslinkable P(MMA-co-AMC) chains
time f(t) of crosslinkable P(MMA-co-AMC) chains in chloroform during the
in chloroform during the sol–gel transition. Note that Af+As¼1. Reproduced
sol–gel transition, where f(t) is t-weighted G(t), that is, f(t)tG(t), because
with permission from Ngai et al.102 Copyright (2004) American Chemical
the x axis is logarithmic. Reproduced with permission from Ngai et al.102
Society.
Copyright (2004) American Chemical Society.
Polymer Journal
The slow relaxation mode
J Li et al
619
As for the increase of /ISslow, we should note that in semidilute or low-salt solution; the counterions must be condensed nearby the
solution polymer, chains are highly entangled with each other. Inside chain backbone, and the fluctuation of condensed counterions toward
the volume of one chain, there also exist many other chains, especially and away from each chain backbone induces an electric field that leads
when the chain is long. Assuming that each chain with a radius of to an electrophoresis-mobility-related diffusive relaxation.
gyration (Rg) has ns segments and there are np chains inside its On the other hand, the interpretation of the slow mode, especially
occupied space (BRg)3, the scattering intensity (/IS) from the nsnp for those very slow relaxation modes observed in salt-free or low-salt
segments on the basis of the LLS theory is related the position of each polyelectrolyte solutions, is even more controversial. It has been
segment (r) as follows,108 attributed to dynamics of large multichain domains (transient aggre-
*n n n n + gates or clusters) formed due to electrostatic interaction76–80 or some
X s p X
s p
hIi / ! ! !
expi½ q ð r r Þ ð9Þ insoluble chain clusters, or even a trace amount of large particles
T j k
j¼1 k¼1 introduced during the imperfect preparation of solution.89–93 The
If the motions of nsnp segments are fully correlated, /ISTp(nsnp)2 transient cluster interpretation was supported by studies of poly(styrene-
when qRg{1; although if the motions of nsnp segments are completely sulfonate) with sodium counterions in aqueous solutions under
independent, /ISTpnsnp. In reality, /IST is between nsnp and dialysis. For a long time, this transient cluster has been attributed to
(nsnp)2 in semidilute solution because the segments between different the effective interaction between similarly charged segments, namely,
chains are only partially correlated. The crosslinking makes the the overlapping of the ion clouds of neighboring polyions, so that the
segments between different chains more correlated so that the scatter- more loosely associated small ions are ‘shared’ by two or more
ing intensity increases. On the other hand, the increase of the solution polyions chains, resulting in a fluctuating dipole field that tends to
viscosity at lower temperatures enhances the hydrodynamic inter- retard the relative motions of those participating polyions.
action of the segments between different chains. This is why /ISslow Experimentally, a polymer solution is normally prepared by allow-
increases with a decreasing temperature. ing a macroscopic piece of polymer to dissolve in a solvent. The
so-called permanent chain clusters must be from the incomplete
dissolution in the solution preparation, not the chain association,
The slow mode in salt-free or low-salt polyelectrolytes solution because we do not place individual chains into a solvent. Cong et al.95
Dynamics of polyelectrolytes in salt-free or low-salt solutions has addressed those possible problems and recently prepared a ‘virgin’
attracted much attention after Lin et al.111 reported their dynamic LLS poly(styrenesulfonate) sample directly from 4-styrene-sulfonic sodium
observation of an extremely slow diffusive mode in the low-salt salt in an aqueous medium. They then used dialysis to in situ change
poly(L-lysine) solutions. Later, it was shown that there are actually a the salt concentration. Even in that case, the solution cleanliness still
fast and a slow relaxation mode in dynamic LLS measurements of salt- deteriorates after the dialysis. To avoid some possible problems in
free or low-salt dilute polyelectrolyte solution in comparison with the sample preparation, ideally, one would like to have a polymer solution
translational diffusive relaxation mode of individual neutral chains in which polymer chains can be in situ switched from a completely
with similar lengths. Some reported slow relaxation modes were really neutral to a fully charged state by some chemical reactions under a
slow with a hydrodynamic size of 102–103 nm or larger. Gradually, mild condition. We have searched such a system for a long time.
such two modes have been reported for nearly all charged macro- Recently, using the Staudinger ligation to attach 4-(5¢-(1-(dimethyl-
molecules, including both synthetic and biological polyelectrolytes in amino)-ethylidene-amino)pentyl)-1-methyl-2-(diphenylphosphino)tere-
aqueous and nonaqueous solutions.76–84 Therefore, these two modes phthalate to poly(p-azidomethylstyrene)-co-polystyrene, we successfully
are general features of salt-free or low-salt polyelectrolyte solutions. prepared a novel polymer, poly(4-(4-((5-(1-(dime-thylamino)ethylide-
In the past, the fast mode in polymer dilute and semidilute solutions neamino)pentyl-oxy)carbonyl)-2-(diphenylphosphoryl)benzoyloxy-
has been attributed to different origins, such as propagation of excita- methyl-styrene)-co-poly(styrene) (P(‘amidine’MS)-co-PS), which
tions in a polyelectrolyte pseudolattice or the free diffusion of the undergo a reversible neutral-charged-neutral transition when CO2
noncaged chains/particles. When polymer chains are charged in a salt- and N2 are alternatively bubbled through its N,N-dimethylformamide
free or low-salt solution, their translational diffusive relaxation modes (DMF) solution as illustrated in Scheme 3. Armed with this novel
split into a fast and a slow diffusive relaxation mode. Muthukumar86–88 polymer, we are able to rule out any interference of impurities or
and Lee and colleagues111 suggested that the fast mode, independent of permanent chains clusters on solutions, and in situ study the variation
the chain length, would be related to the coupled diffusion (Df ) of of solution dynamics of polymer chains during the neutral-charged-
polyelectrolyte chains and their small counterions, namely, in a salt-free neutral transition.
* n m * n m
NH N2 NH
Me O Me
O O
O O N N O N N H
P Me P Me
Me O Me
O
HCO3
Scheme 3 Reversible change between P(‘amidine’MS)-co-PS and its charged state in DMF with 0.5% H2O after alternating bubbling of CO2 and N2.
Reproduced with permission from Zhou et al.107 Copyright (2009) American Chemical Society.
Polymer Journal
The slow relaxation mode
J Li et al
620
120 1.8
CO2 CO2 initial neutral state
after CO2 bubbling
Conductivity / (µS cm–1)
G()
C = 1.45 mg ml–1
T = 25 °C
40 0.6
N2 N2 N2
0 0.0
0 20 40 60 10-4 10-3 10-2 10-1
Time / min /s
Figure 18 Conductivity of P(‘amidine’MS)-co-PS in DMF with 0.5% H2O at Figure 20 Characteristic relaxation time distributions in one cycle of CO2
25 1C after alternative bubbling of CO2 and N2, where polymer concentration and N2 bubbling of P(‘amidine’MS)-co-PS solution. Reproduced with
is 7.3 mg ml–1. Insets are photos of P(‘amidine’MS)-co-PS in THF with 0.5% permission from Zhou et al.107 Copyright (2009) American Chemical
H2O in different states before and after CO2 and N2 bubbling, where Society.
polymer concentration is 25.0 mg ml–1. Reproduced with permission from
Zhou et al.107 Copyright (2009) American Chemical Society.
= 15° 2.0
C = 1.45 mg ml–1
40
T = 25 °C during bubbling <>s
CO2 0.0
0.0 2.0 4.0 6.0 8.0
0 100 200 0 300 0 100 200 300
q2 / 1010 cm-2
Time / s
Figure 21 The scattering vector (q) dependence of average characteristic
Figure 19 Time dependence of time-average scattering intensity of line widths (G) in one cycle of CO2 and N2 bubbling of P(‘amidine’MS)-
P(‘amidine’MS)-co-PS in DMF with 0.5% H2O at different states before and co-PS solution. Reproduced with permission from Zhou et al.107 Copyright
after CO2 and N2 bubbling. Reproduced with permission from Zhou et al.107 (2009) American Chemical Society.
Copyright (2009) American Chemical Society.
Figure 18 shows that after CO2 and N2 gases were alternatively It has been well known in LLS that /IS is proportional to (qC/qp)T,
bubbled 30 min through each solution of P(‘amidine’MS)-co-PS in where C and p are polymer concentration and solution osmotic
DMF at 25 1C, the solution conductivity dramatically changes pressure, respectively. When polymer chains are charged, it is more
five times and switches between two constant values. The switching difficult to induce the concentration fluctuation for a given osmotic
is completely reversible. The inset photos in Figure 18 are the pressure change because of electrostatic repulsion. Therefore, such a
appearance of the solution of P(‘amidine’MS)-co-PS in a different decrease of /IS after the chains are charged is understood.
solvent (tetrahydrofuran, THF) with 0.5% H2O, in which the charged Figure 20 shows variation of the characteristic relaxation time
P(‘amidine’MS)-co-PS chains have an extremely low solubility. The distribution G(t) in one cycle of CO2 and N2 bubbling of the
change solution appearance from clear to cloudy and from cloudy to P(‘amidine’MS)-co-PS solution. When the polymer chains are their
clear during the alternative bubbling of CO2 and N2 clearly shows the initial neutral state before the CO2 bubbling, there is only one
expected neutral-charged-neutral transition of P(‘amidine’MS)-co-PS relaxation mode. When each P(‘amidine’MS)-co-PS chain is charged
chains in THF. after the CO2 bubbling, G(t) splits into two peaks. The N2 bubbling
Dynamic LLS was also used to follow such neutral-charged-neutral can return G(t) to its initial state, further indicating that the neutral-
transition of P(‘amidine’MS)-co-PS chains. Figure 19 shows that as charged-neutral transition is completely reversible. The scattering
individual P(‘amidine’MS)-co-PS chains are gradually charged during vector (q) dependence of average characteristic line widths (/GS)
the CO2 bubbling, the time-average scattering intensity (/IS) in the corresponding one cycle of CO2 and N2 bubbling of the
decreases B45% over a period of B500 s. The alternative bubbling P(‘amidine’MS)-co-PS solution is shown in Figure 21. No matter
of CO2 or N2 can reversibly switch /IS between two average values. whether the chains are charged, the average characteristic line width
Polymer Journal
The slow relaxation mode
J Li et al
621
1.0 1.2
T = 25 °C charged state
2 h bubbling N2
C = 1.45 mg ml–1
5 h bubbling N2
0.8 neutral state
As /(As +Af )
G()
0.5
= 15°
0.4 C = 1.45 mg ml–1
T = 25 °C
0.0
0.0 10-4 10-3 10-2 10-1
0.0 2.0 4.0 6.0 8.0
/s
q 2 / 1010 cm-2
Figure 23 Characteristic relaxation time distributions G(t) during the
Figure 22 Scattering vector (q) dependence of intensity contribution of slow charge-to-neutral transition induced by slowly bubbling N2 through
relaxation mode (As) after N2 bubbling of P(‘amidine’MS)-co-PS solution. P(‘amidine’MS)-co-PS solution. Reproduced with permission from Zhou
Note that As+As¼1. Reproduced with permission from Zhou et al.107 et al.107 Copyright (2009) American Chemical Society.
Copyright (2009) American Chemical Society.
=15°
These results are similar to those recently observed by Russo and T = 25 °C
colleagues.95 but they still attributed the slow mode to some temporal 0.4 C = 1.45 mg ml–1
chain aggregates. We found that our results agree well with the CLiBr = 50 mM
interpretation of dynamics of individual polyelectrolyte chains in
salt-free dilute solutions proposed by Muthukumar.86–88 Pusey and
Tough147 attributed the fast and slow modes to the mutual and self-
diffusion, respectively. For example, in a short time, each chain moves 0.0
10-4 10-3 10-2 10-1
over a short distance and the diffusion of counterions makes the
/s
charged particles to diffuse faster. After the relaxation of such
‘collective diffusion’, the ‘self-diffusion’ of individual chains in a long Figure 24 Characteristic relaxation time distributions G(t) of
time or over a long length must be retarded by those surrounding P(‘amidine’MS)-co-PS solution in the presence of 50 mM LiBr before and
after CO2 bubbling. Reproduced with permission from Zhou et al.107
interacting particles and becomes detectable in LLS. Our fast mode
Copyright (2009) American Chemical Society.
agrees well with the interpretation of the coupled diffusion of
individual polyelectrolyte chains and their counterions in salt-free
dilute solutions proposed by Lin et al.111 and Muthukumar,86–88 When the high amount of salt LiBr was added into the solution, as
whereas the slow mode is related to the self-diffusion of the center- expected, we can find that there is only one relaxation mode after the
of-mass of individual chains under long-range electrostatic interac- CO2 bubbling. Figure 24 shows that the CO2 bubbling has no effect on
tion-induced constrains of other surrounding chains, just like a cage, G(t) of charged P(‘amidine’MS)-co-PS chains in dilute solutions;
because our results reveal that the slow mode is related to small but namely, the addition of 50 mM LiBr sufficiently suppress electrostatic
slowly moving subjects, in other words, the long-range electrostatic interaction among charges on the chain backbone so the charged
interaction may slow down the self-diffusion of the center-of-mass of chains behave like in the initial neutral state. The slightly larger /RhS
individual chain. is attributed to the chain swelling because the addition of LiBr makes
Figure 23 shows how G(t) changes in the process of backing to the DMF a better solvent.
neutral state when N2 is slowly bubbled through the solution. It is In summary, by using a novel P(‘amidine’MS)-co-PS that can be in
clear that during the charge-to-neutral transition, /tSfast increases, situ and reversibly switched between a neutral state to a charged state
that is, Dfast decreases, so that the coupled fast relaxation slows down. by alternative bubbling of CO2 and N2, our LLS results reconfirm that
At the same time, the total scattering intensity /IS and the contribu- when individual polymer chains are charged in salt-free dilute solu-
tion from the slow relaxation (As) increase because the electrostatic tions, their translational diffusive relaxation in the neutral state splits
interaction constrained the motion of chains weakly and the intensity into a fast and a slow diffusive mode due to long-range electrostatic
of the fast mode, which is contributed to the coupled diffusive interaction. As expected, adding high amount of salt can suppress the
relaxation of individual polyelectrolyte chains and their counterions, electrostatic interaction and the charged chains behave like the neutral
diminishes as each chain becomes neutral. ones with only one translational diffusive relaxation mode. The fast
Polymer Journal
The slow relaxation mode
J Li et al
622
Scheme 4 Schematic of fast and slow relaxation modes after each P(‘amidine’ MS)-co-PS chain becomes charged after the CO2 bubbling; namely, the fast
mode is because of the coupling of the diffusion of the chain and the wiggling of chain segments induced by fluctuation of counterions condensed nearby
the charged chain backbone in a salt-free or low-salt solution and the slow mode to the self-diffusion of the charged chain with an extended conformation.
Reproduced with permission from Zhou et al.107 Copyright (2009) American Chemical Society.
mode can be attributed to the coupled diffusion originated from a further confirms the coexistence of individual chains and large
convective current generated by an induced electric field arising from temporal chain clusters in the semidilute regimes.
fluctuation of all charged species (charges on the chain and counter- When C4Ce, all the chains are entangled with each other to form
ions) in the solution in a short time or length scale, whereas the slow one huge ‘cluster’. Each chain is confined inside a ‘tube’ made of its
mode is related to slowly moving scattering objects but it is not large, surrounding chains and the gravity center of each chain segment
supporting the assumption of the retarded self-diffusion as Scheme 4. (blob) fluctuates inside the tube within a short time scale (B103 s).
In an athermal solvent, there is no interaction among the chains so
Conclusion that all the blobs are in the same microenvironment, leading to only
Using in situ anionic polymerization to directly prepare macroscopi- one fast relaxation mode. In a less good/y solvent, the relatively
cally homogeneous dust-free semidilute/concentrated PS solutions stronger segment–segment interaction near the entanglement points
inside a LLS cell, we comparatively reexamined solution dynamics makes the tube nonuniform. It can be visualized that each chain is
of PS in an athermal solvent (toluene) and in a less good/y solvent inside a nonuniform tube with a band-like structure. Therefore, the
(cyclohexane in the range 34–50 1C) by a combination of static and blobs experience two different microenvironments. The restricted
dynamic LLS. We confirm that in toluene there is only one fast movement of the blobs near the entanglement points results in the
diffusive relaxation mode in the measured intensity–intensity time slow mode. The use of a special (2+2) photoaddition to induce in situ
correlation function even for the solution with a concentration (C) 30 crosslinking of polymer chains rather than the entanglement in a
times higher than the overlapping concentration (C*). In cyclohexane, semidilute solution further indicates an expected increase of the
an additional slow mode appears as soon as C/C*41. Our results relative contribution of the slow mode as the crosslinking proceeds
show that the slow mode is real, not because of previously suggested because the crosslinking enhances the restricted movement of the
artifacts in the solution preparation. The concentration, and time polymer chains near the crosslinked (entangled) points.
dependence of the characteristic relaxation time of the slow mode By using a novel P(‘amidine’MS)-co-PS that can be in situ and
(/tSs), the scaling exponent (as) between /tSs and the scattering reversibly switched between a neutral to a charged state by alternative
vector (q), and the relative intensity contribution (As) of the slow bubbling of CO2 and N2, we reconfirm that when individual polymer
mode suggests that such a slow mode has two different natures, chains are charged in salt-free dilute solutions, their translational
depending on the polymer concentration. diffusive relaxation in the neutral state splits into a fast and a slow
When C is higher than the overlapping concentration (C*) but diffusive mode (Df and Ds), respectively, due to the long-range
lower than the entanglement concentration (Ce), only some of the electrostatic interaction. The fast mode can be attributed to the
chains are entangled together to form large transient clusters, which coupled diffusion originated from a convective current generated by
coexist with individual (nonentangled) chains in the solution. In such an induced electric field arising from fluctuation of all charged species
a transitional region between dilute and semidilute solutions, the (charges on the chain and counterions) in the solution in a short time
relatively stronger segment–segment interaction in a less good/y or length scale. However, there is still no decisive evidence to
solvent makes the movements of the chains inside each cluster more differentiate whether the slow relaxation is related to large temporal
correlated within a short delay time window (o1 s) and leads to the aggregates formed because of the overlapping of ion clouds among
slow mode. The temperature dependence of the slow relaxation mode different chains or to the self-diffusion of individual chains retarded
in two different systems, PNIPAM/H2O and PEO-PPO-PEO/H2O by surrounding chains (interchain friction) in a long time or length
Polymer Journal
The slow relaxation mode
J Li et al
623
scale. Our results indicate that the slow mode is not related to large 25 Jian, T., Anastasiadis, S. H., Semenov, A. N., Fytas, G., Adachi, K. & Kotaka, T.
but slowly moving scattering objects, supporting the assumption of Dynamics of composition fluctuations in diblock copolymer solutions far from and
near to the ordering transition. Macromolecules 27, 4762–4773 (1994).
the retarded self-diffusion. Consequently, we conclude that the slow 26 Balsara, N. P., Stepanek, P., Lodge, T. P. & Tirrell, M. Dynamic light scattering from
mode observed in the semidilute, polyelectrolyte, or even gelling microstructured block copolymer solutions. Macromolecules 24, 6227–6230 (1991).
system is real and can be generalized as hindered motions of inter- 27 Pan, C., Maurer, W., Liu, Z., Lodge, T. P., Stepanek, P., von Meerwall, E. D. &
Watanabe, H. Dynamic light-scattering from dilute, semidilute, and concentrated
acting polymer chains even though the nature of interaction can be block-copolymer solutions. Macromolecules 28, 1643–1653 (1995).
different, including, the weak segment–segment interaction in a less 28 Chu, B. & Nose, T. Static and dynamical properties of polystyrene in trans-decalin.4.
good solvent to strong electrostatic interaction among polyelectrolyte Osmotic compressibility, characteristic lengths, and internal and pseudogel motions in
the semidilute regime. Macromolecules 13, 122–132 (1980).
chains, and even to chemical crosslinking inside gel networks. 29 Ewen, B., Richter, D., Farago, B. & Stühn, B. Neutron spin-echo investigations on the
segmental dynamics in semidilute polymer-solutions under theta-solvent and good
solvent conditions. J. Non-Cryst. Solids 172–174, 1023–1027 (1994).
ACKNOWLEDGEMENTS 30 Rital, A., Belkoura, L. & Woermann, D. Static and dynamic light scattering experi-
The financial support of the National Natural Scientific Foundation of China ments on semidilute solutions of polystyrene in cyclohexane between the Theta-
(NNSFC) Project (20904049, 20934005 and 50773077), and the financial temperature and the binodal curve. Phys. Chem. Chem. Phys. 1, 1947–1955 (1999).
support of the Hong Kong Special Administration Region General Research 31 Wiltzius, P., Hans, H. R., Cannell, D. S. & Schaefer, D. W. Universality for
static properties of polystyrenes in good and marginal solvents. Phys. Rev. Lett. 51,
Fund (CUHK402506, 2160291; and CUHK403908, 2160361) is gratefully
1183–1186 (1983).
acknowledged. 32 Amis, E. J. & Han, C. C. Cooperative and self-diffusion of polymers in semi-dilute
solutions by dynamic light-scattering. Polymer 23, 1403–1406 (1982).
33 Melnichenko, Y. B., Brown, W., Rangelov, S., Wignall, G. D. & Stamm, M.
Dynamic acid static correlations in solutions of polymers in liquid and supercritical
solvents: dynamic light scattering and small angle neutron scattering. Phys. Lett.
1 Rouse, P. E. J. A theory of the linear viscoelastic properties of dilute solutions of A 268, 186–194 (2000).
coiling polymers. J. Chem. Phys. 21, 1272–1280 (1953). 34 Adam, M. & Delsanti, M. Dynamical properties of polymer-solutions in good solvent by
2 Zimm, B. H. Dynamics of polymer molecules in dilute solution—viscoelasticity, flow Rayleigh-scattering experiments. Macromolecules 10, 1229–1237 (1977).
birefringence and dielectric loss. J. Chem. Phys. 24, 269–278 (1956). 35 Adam, M. & Delsanti, M. Dynamical behavior of semidilute polymer-solutions
3 Pecora, R. Spectral distribution of light scattered from flexible-coil macromolecules. in a theta-solvent—quasi-elastic light-scattering experiments. Macromolecules 18,
J. Chem. Phys. 49, 1032–1035 (1968). 1760–1770 (1985).
4 de Gennes, P. G. Quasi-elastic scattering of neutrons by dilute polymer solutions: 36 Brown, W. & Stepanek, P. Distribution of relaxation-times from dynamic light-scatter-
I free-draining limit. Physics 3, 37–45 (1967). ing on semidilute solutions—polystyrene in ethyl-acetate as a function of temperature
5 Dubois-Violette, E. & de Gennes, P. G. Quasi-elastic scattering by dilute, ideal, from good to theta-conditions. Macromolecules 21, 1791–1798 (1988).
polymer solution: II effects of hydrodynamic interactions. Physics 3, 181–198 37 Brown, W. & Stepanek, P. Dynamic behavior in concentrated polystyrene cyclohexane
(1967). solutions close to the theta-point—relaxation-time distributions as a function of
6 Akcasu, A. Z. & Gurol, H. Quasi-elastic scattering by dilute polymer-solutions. concentration and temperature. Macromolecules 25, 4359–4363 (1992).
J. Polym. Sci.: Polym. Phys. Ed. 14, l–10 (1976). 38 Brown, W. & Stepanek, P. Viscoelastic relaxation in semidilute and concentrated
7 Akcasu, A. Z. & Higgins, J. S. Quasi-elastic scattering of neutrons from freely jointed polymer-solutions. Macromolecules 26, 6884–6890 (1993).
polymer-chains in dilute-solutions. J. Polym. Sci.: Polym. Phys. Ed. 15, 1745–1756 39 Konak, C., Mrkviekova, L. & Bansil, R. Dynamics of pregel solutions and gels in a
(1977). Theta-solvent near a spinodal. Macromolecules 29, 6158–6164 (1996).
8 Benmouna, M. & Akcasu, A. Z. Temperature effects on dynamic structure factor in 40 Brochard, F. & de Gennes, P. G. Dynamical scaling for polymers in theta-solvents.
dilute polymer-solutions. Macromolecules 11, 1187–1192 (1978). Macromolecules 10, 1157–1161 (1977).
9 Chu, B. Laser Light Scattering 291–293 (Academic Press, New York, 1974). 41 Wang, C. H. Dynamic light-scattering, mutual diffusion, and linear viscoelasticity of
10 Berne, B. J. & Percora, R. Dynamic Light Scattering (Wiley Interscience, New York, polymer-solutions. J. Chem. Phys. 95, 3788–3797 (1991).
1976). 42 Wang, C. H. Dynamic light-scattering and viscoelasticity of a binary polymer-solution.
11 Pecora, R. Dynamic Light Scattering: Applications of Photon Correlation Spectroscopy
Macromolecules 25, 1524–1529 (1992).
(Plenum Press, New York, 1985). 43 Wang, C. H., Fytas, G. & Fischer, E. W. Determination of the longitudinal compliance
12 Han, C. C. & Akcasu, A. Z. Dynamic light-scattering of dilute polymer-solutions in the
of polyvinyl acetate) by using the dynamic light-scattering technique. J. Chem. Phys.
non-asymptotic q-region. Macromolecules 14, 1080–1084 (1981).
82, 4332–4336 (1985).
13 Lodge, T. P., Han, C. C. & Akscasu, A. Z. Temperature-dependence of dynamic
44 Wang, C. H. & Zhang, X. Q. Quasi-elastic light-scattering and viscoelasticity of
light-scattering in the intermediate momentum-transfer region. Macromolecules 16,
polystyrene in diethyl phthalate. Macromolecules 26, 707–714 (1993).
1180–1183 (1983).
45 Wang, C. H. & Zhang, X. Q. Quasi-elastic light-scattering investigation of concentra-
14 King, T. A., Knox, A. & McAdam, J. D. G. Internal motion in chain polymers. Chem.
tion fluctuations and coupling to stress-relaxation in a polymer-solution—polystyrene
Phys. Lett. 19, 351–354 (1973).
in CCl4. Macromolecules 28, 2288–2296 (1995).
15 Huang, W. N. & Frederick, J. E. Determination of intramolecular motion in a random-
46 Wang, C. H., Sun, Z. & Huang, Q. R. Quasielastic light scattering of polystyrene in
coil polymer by means of quasi-elastic light-scattering. Macromolecules 7, 34–39
diethyl malonate in semidilute concentration. J. Chem. Phys. 105, 6052–6059 (1996).
(1974).
47 Adam, M., Farago, B., Schleger, P., Raspaud, E. & Lairez, D. Binary contacts in
16 Tsunashima, Y., Nemoto, N. & Kurata, M. Dynamic light-scattering-studies of polymer-
semidilute solution: good and theta solvents. Macromolecules 31, 9213–9223
solutions .1. Histogram analysis of internal motions. Macromolecules 16, 584–589
(1983). (1998).
17 Nemoto, N., Makita, Y., Tsunashima, Y. & Kurata, M. Dynamic light scattering studies 48 Colby, R. H. & Rubinstein, M. 2-parameter scaling for polymers in theta-solvents.
of polymer-solutions. 3. Translational diffusion and internal motion of high molecular- Macromolecules 23, 2753–2757 (1990).
weight polystyrenes in benzene at infinite dilution. Macromolecules 17, 425–430 49 Nicolai, T., Brown, W., Johnsen, R. M. & Stepanek, P. Dynamic behavior of o solutions
(1984). of polystyrene investigated by dynamic light-scattering. Macromolecules 23, 1174
18 Kim, S. H., Ramsay, D., Patterson, G. D. & Selser, J. C. Static and dynamic light- (1990).
scattering of poly-(alpha-methyl styrene) in toluene in the dilute region. J. Polym. 50 Nicoclai, T. & Brown, W. Static and dynamic light-scattering-studies on semidilute
Sci.Polym. Phys. 28, 2023–2056 (1990). solutions of polystyrene in cyclohexane as a function of temperature. Macromolecules
19 Chirico, G. & Baldini, G. Dynamic light-scattering from DNA plasmids—diffusional 23, 3150–3155 (1990).
and internal motion. J. Mol. Liq. 41, 327–345 (1989). 51 Nicoclai, T. & Brown, W. Cooperative diffusion of concentrated polymer solutions:
20 Chu, B., Wang, Z. & Yu, J. Dynamic light-scattering study of internal motions of a static and dynamic light scattering study of polystyrene in DOP. Macromolecules 29,
polymer coils in dilute-solution. Macromolecules 24, 6832–6838 (1991). 1698–1704 (1996).
21 Sorlie, S. S. & Pecora, R. A dynamic light-scattering study of a 2311 base pair DNA 52 Brown, W. & Johnsen, R. M. Diffusion-coefficients in semidilute solutions evaluated
restriction fragment. Macromolecules 21, 1437–1449 (1988). from dynamic light-scattering and concentration gradient measurements as a function
22 Nicolai, T., Brown, W. & Johnsen, R. The internal-modes of polystyrene single coils of solvent quality.1.Intermediate molecular-weights. Macromolecules 18, 379–387
studied using dynamic light-scattering. Macromolecules 22, 2795–2801 (1989). (1985).
23 Wu, C. & Zhou, S. Internal motions of both poly(N-isopropylacrylamide) linear chains 53 Xie, Y., Ludwing, K. F., Bansil, R., Gallagher, P. D., Cao, X. & Morales, G. Small-angle
and spherical microgel particles in water. Macromolecules 29, 1574–1578 (1996). X-ray scattering studies of semidilute polystyrene-cyclohexane solutions. Physica A
24 Tsunashima, Y. & Kawamata, Y. Dynamics of polystyrene subchains of a styrene- 232, 94–108 (1996).
methyl methacrylate diblock copolymer in solution measured by dynamic 54 Kostko, A. F., Anisimov, M. A. & Sengers, J. V. Dynamic crossover to tricriticality and
light-scattering with an isorefractive index solvent.1. dilute-solution region. Macro- anomalous slowdown of critical fluctuations by entanglements in polymer solutions.
molecules 26, 4899–4909 (1993). Phys. Rev. E 66, 020803–020806 (2002).
Polymer Journal
The slow relaxation mode
J Li et al
624
55 Brown, W., Johnsen, R. M., Konak, C. & Dvoranek, L. Dynamics in concentrated 84 Valachovic, D. E. & Amis, E. J. Polyelectrolyte dendrimers in low-salt solutions. Abstr.
polymer-solutions by polarized Rayleigh-Brillouin scattering and dynamic light-scat- Pap. Am. Chem. Soc. 209, 58 (1995).
tering. J. Chem. Phys. 95, 8568–8577 (1991). 85 Dobrynin, A. V. & Rubinstein, M. Theory of polyelectrolytes in solutions and at
56 Nose, T. & Chu, B. Static and dynamical properties of polystyrene in trans decalin. 1. surfaces. Prog. Polym. Sci. 30, 1049–1118 (2005).
NBS 705 standard near h conditions. Macromolecules 12, 590–599 (1979). 86 Muthukumar, M. Polyelectrolyte dynamics. Adv. Chem. Phys. 131, 1–60 (2005).
57 Sawatari, N., Yoshizaki, T. & Yamakawa, H. Dynamic structure factor of polystyrene 87 Muthukumar, M. Double screening in polyelectrolyte solutions: limiting laws and
and poly(methyl methacrylate) in Theta solvents. Macromolecules 31, 4218–4222 crossover formulas. J. Chem. Phys. 105, 5183–5199 (1996).
(1998). 88 Muthukumar, M. Dynamics of polyelectrolyte solutions. J. Chem. Phys. 107,
58 Schröder, J. M., Wiegand, S., Aberle, L. B., Kleemeier, M. & Schroer, W. Experimental 2619–2635 (1997).
determination of singly scattered light close to the critical point in a polystyrene- 89 Li, X. & Reed, W. F. Polyelectrolyte properties of proteoglycan monomers. J. Chem.
cyclohexane mixture. Phys. Chem. Chem. Phys. 1, 3287–3292 (1999). Phys. 94, 4568–4580 (1991).
59 Appelt, B. & Meyerhoff, G. Characterization of polystyrenes of extremely high 90 Reed, W. F., Ghosh, S., Medjahdi, G. & Francois, J. Dependence of polyelectrolyte
molecular-weights. Macromolecules 13, 657–662 (1980). apparent persistence lengths, viscosity, and diffusion on ionic-strength and linear
60 Brown, W. Slow-mode diffusion in semidilute solutions examined by dynamic light- charge-density. Macromolecules 24, 6189–6198 (1991).
scattering. Macromolecules 17, 66–72 (1984). 91 Ghosh, S., Li, X., Reed, C. E. & Reed, W. F. Apparent persistence lengths and diffusion
61 Brown, W. Dynamical properties of high molecular-weight polystyrene in tetrahydro- behavior of high-molecular-weight hyaluronate. Biopolymers 30, 1101–1112 (1990).
furan in the dilute semidilute transition region. Macromolecules 18, 1713–1719 92 Peitzsch, R. M., Burt, M. J. & Reed, W. F. Evidence of partial draining for linear
(1985). polyelectrolytes—heparin, chondroitin 6-sulfate, and poly(styrenesulfonate). Macro-
62 Bastide, J. & Leibler, L. Large-scale heterogeneities in randomly cross-linked net- molecules 25, 806–815 (1992).
works. Macromolecules 21, 2647–2649 (1988). 93 Ghosh, S., Peitzsch, R. M. & Reed, W. F. Aggregates and other particles as the origin of
63 Brown, W. & Mortensen, K. Dynamical properties of high molecular-weight polystyrene the extraordinary diffusional phase in polyelectrolyte solutions. Biopolymers 32,
in tetrahydrofuran in the dilute semidilute transition region. Macromolecules 21, 1105–1122 (1992).
420–425 (1988). 94 Ermi, B. D. & Amis, E. J. Domain structures in low ionic strength polyelectrolyte
64 Colby, R. H., Fetters, L. J., Funk, W. G. & Graessley, W. W. Effects of concentration and solutions. Macromolecules 31, 7378–7384 (1998).
thermodynamic interaction on the viscoelastic properties of polymer-solutions. Macro- 95 Cong, R., Temyanko, E., Russo, P. S., Edwin, N. & Uppu, R. M. Dynamics of poly
molecules 24, 3873–3882 (1991). (styrenesulfonate) sodium salt in aqueous solution. Macromolecules 39, 731–739
65 Nicolai, T., Brown, W., Hvidt, S. & Heller, K. A comparison of relaxation-time (2006).
distributions obtained from dynamic light-scattering and dynamic mechanical mea- 96 Yuan, G. C., Wang, X. H., Han, C. C. & Wu, C. Reexamination of slow dynamics in
surements for high-molecular-weight polystyrene in entangled solutions. Macromole- semidilute solutions: from correlated concentration fluctuation to collective diffusion.
cules 23, 5088–5096 (1990). Macromolecules 39, 3642–3647 (2006).
66 Vshivkov, S. A. & Safronov, A. P. The conformational coil-globule transition of 97 Yuan, G. C., Wang, X. H., Han, C. C. & Wu, C. Reexamination of slow dynamics in
polystyrene in cyclohexane solution. Macromol. Chem. Phys. 198, 3015–3023 semidilute solutions: temperature and salt effects on semidilute poly(N-isopropyl-
(1997). acrylamide) aqueous solutions. Macromolecules 39, 6207–6209 (2006).
67 Liu, Z., Pan, C., Lodge, T. P. & Stepanek, P. Dynamic light-scattering from block- 98 Li, W., Hong, L. Z., Ngai, T., Huang, H. Y., He, T. B. & Wu, C. A comparative study of
copolymer solutions under the zero average contrast condition. Macromolecules 28, chain dynamics of di- and tri-block copolymers in semidilute solution in a non-
3221–3229 (1995). selective solvent. Chin. J. Polym. Sci. 22, 589–598 (2004).
68 Borsali, R., Fischer, E. W. & Benmouna, M. Dynamic light-scattering from polystyrene- 99 Li, J. F., Lu, Y. J., Zhang, G. Z., Li, W. & Wu, C. A slow relaxation mode of polymer
poly(methylmethacrylate) diblock copolymer in toluene. Phys. Rev. A 43, 5732–5735 chains in a semidilute solution. Chin. J. Polym. Sci. 26, 465–473 (2008).
(1991). 100 Wu, C. & Ngai, T. Reexamination of slow relaxation of polymer chains in sol-gel
69 Benmouna, M., Benoit, H., Borsali, R. & Duval, M. Theory of dynamic scattering from transition. Polymer 45, 1739–1742 (2004).
copolymer solutions using the random phase approximation. Macromolecules 20, 101 Ngai, T., Wu, C. & Chen, Y. Origins of the speckles and slow dynamics of polymer gels.
2620–2624 (1987). J. Phys. Chem. B. 108, 5532–5540 (2004).
70 Jian, T., Anastasiadis, S. H., Semenov, A. N., Fytas, G., Adachi, K. & Kotaka, T. 102 Ngai, T., Wu, C. & Chen, Y. Effects of temperature and swelling on chain dynamics
Dynamics of composition fluctuations in diblock copolymer solutions far from and during the sol-gel transition. Macromolecules 37, 987–993 (2004).
near to the ordering transition. Macromolecules 27, 4762–4773 (1994). 103 Ngai, T. & Wu, C. Effect of cross-linking on dynamics of semidilute copolymer
71 Semenov, A. N., Anastasiadis, S. H., Boudenne, N., Fytas, G., Xenidou, M. & solutions: poly(methyl methacrylate-co-7-acryloyloxy-4-methylcoumarin) in chloro-
Hadjichristidis, N. Dynamic structure factor of diblock copolymers in the ordering form. Macromolecules 36, 848–854 (2003).
regime. Macromolecules 30, 6280–6294 (1997). 104 Zhao, Y. & Wu, C. A hybrid polymer gel and its static nonergodicity. Chin. J. Polym.
72 Chrissopoulou, K., Pryamitsyn, V. A., Anastasiadis, S. H., Fytas, G., Semenov, A. N., Sci. 20, 269–274 (2002).
Xenidou, M. & Hadjichristidis, N. Dynamics of the most probable composition 105 Zhao, Y., Zhang, G. Z. & Wu, C. Nonergodic dynamics of a novel thermally sensitive
fluctuations of ‘real’ diblock copolymers near the ordering transition. Macromolecules hybrid gel. Macromolecules 34, 7804–7808 (2001).
34, 2156–2171 (2001). 106 Wang, C. Q., Jiang, S. H. & Wu, C. Application of the temperature-ramped holographic
73 Anastasiadis, S. H., Fytas, G., Vogt, S. & Fisher, E. W. Breathing and composition relaxation spectroscopy in the investigation of physically cross-linked gels. Macro-
pattern relaxation in homogeneous diblock copolymers. Phys. Rev. Lett. 70, molecules 34, 6737–6741 (2001).
2415–2418 (1993). 107 Zhou, K. J., Li, J. F., Lu, Y. J., Zhang, G. Z. & Wu, C. Re-examination of dynamics of
74 Boudenne, N., Anastasiadis, S. H., Fytas, G., Xenidou, M., Hadjichristidis, N., polyeletrolytes in salt-free dilute solutions by designing and using a novel neutral-
Semenov, A. N. & Fleischer, G. Thermodynamic effects on internal relaxation in charged-neutral reversible polymer. Macromolecules 42, 7146–7154 (2009).
diblock copolymers. Phys. Rev. Lett. 77, 506–509 (1996). 108 Chu, B. Laser Light Scattering 2nd edn. 84–85 (Academic Press, New York, 1991).
75 Li, J. F., Li, W., Huo, H., Luo, S. Z. & Wu, C. Reexamination of the slow mode in 109 Huglin, M. B. Light Scattering from Polymer Solution (Academic Press, New York,
semidilute polymer solutions: the effect of solvent quality. Macromolecular 41, 1972).
901–911 (2008). 110 Kuznetsova, N. A. & Kaliya, O. L. Photochemistry of coumarins. Uspekhi Khimii 61,
76 Sedlak, M., Konak, C., Stepanek, P. & Jakes, J. Semidilute solutions of poly 1243–1267 (1992).
(methacrylic acid) in the absence of salt—dynamic light-scattering study. Polymer 111 Lin, S. C., Lee, W. I. & Schurr, J. M. Brownian-motion of highly charged poly
28, 873–880 (1987). (l-lysine)—effects of salt and polyion concentration. Biopolymers 17, 1041–1064
77 Schmitz, K. S. & Yu, J. W. On the electrostatic contribution to the persistence length (1978).
of flexible poly-electrolytes. Macromolecules 21, 484–493 (1988). 112 Brown, W. & Nicolai, T. Static and dynamic behavior of semidilute polymer-solutions.
78 Mattoussi, M., Karasz, F. E. & Langley, K. H. Electrostatic and screening effects Colloid Polym. Sci. 268, 977–990 (1990).
on the dynamic aspects of polyelectrolyte solutions. J. Chem. Phys. 93, 3593–3603 113 Teraoka, I. Polymer Solution (John Wiley Sons: Inc., New York, 2002).
(1990). 114 de Gennes, P. G. Dynamics of entangled polymer solutions. I. The Rouse Model.
79 Forster, S., Schmidt, M. & Antonietti, M. Static and dynamic light-scattering by Macromolecules 9, 587–593 (1976).
aqueous polyelectrolyte solutions—effect of molecular-weight, charge-density and 115 de Gennes, P. G. Dynamics of entangled polymer solutions. II. Inclusion of hydro-
added salt. Polymer 31, 781–792 (1990). dynamic interactions. Macromolecules 9, 594–598 (1976).
80 Sedlak, M. & Amis, E. J. Dynamics of moderately concentrated salt-free polyelec- 116 Daoud, M., Cotton, J. P., Farnoux, B., Jannink, G., Sarma, G., Benoit, H., Duplessix,
trolyte solutions—molecular-weight dependence. J. Chem. Phys. 96, 817–825 R., Picot, C. & de Gennes, P. G. Solutions of flexible polymers—neutron experiments
(1992). and interpretation. Macromolecules 8, 804–818 (1975).
81 Ermi, B. D. & Amis, E. J. Model solutions for studies of salt-free polyelectrolytes. 117 de Gennes, P. G. Scalling Concenpts in Polymer Physicas (Cornell University Press,
Macromolecules 29, 2701–2703 (1996). Ithaca, NY, 1979).
82 Topp, A., Belkoura, L. & Woermann, D. Effect of charge density on the dynamic 118 Koike, A., Nemoto, N., Inoue, T. & Osaki, K. Dynamic light-scattering and dynamic
behavior of polyelectrolytes in aqueous solution. Macromolecules 29, 5392–5397 viscoelasticity of poly(vinyl alcohol) in aqueous borax solutions. 1. Concentration-
(1996). effect. Macromolecules 28, 2339–2344 (1995).
83 Nierling, W. & Nordmeeir, E. Studies on polyelectrolyte solutions. 7. Fast, hetero- 119 Tao, H., Huang, C. & Lodge, T. P. Correlation length and entanglement spacing
geneous, and slow diffusion modes of poly(diallyl-N,N-dimethylammonium chloride) in concentrated hydrogenated polybutadiene solutions. Macromolecules 32,
in aqueous alcoholic salt solvents. Pol. J. 29, 795–806 (1997). 1212–1217 (1999).
Polymer Journal
The slow relaxation mode
J Li et al
625
120 Yamamoto, I., Iwasaki, K. & Hirotsu, S. Light-scattering study of condensation of poly 135 Chrissopoulou, K., Pryamitsyn, V. A., Anastasiadis, S. H., Fytas, G., Semenov, A. N.,
(N-isopropylacrylamide) chain. J. Phys. Soc. Jpn. 51, 210–215 (1989). Xenidou, M. & Hadjichristidis, N. Dynamics of the most probable composition
121 Di Marzio, E. A. The ten classes of polymeric phase transitions: their use as models for fluctuations of ‘real’ diblock copolymers near the ordering transition. Macromolecules
self-assembly. Prog. in Polym. Sci. 24, 329–377 (1999) and the references in it. 34, 2156–2171 (2001).
122 Smart, T., Lomas, H., Massignani, M., Flores-Merino, M. V., Perez, L. R. & Battaglia, 136 Jian, T., Anastasiadis, S. H., Semenov, A. N., Fytas, G., Adachi, K. & Kotaka, T.
G. Block copolymer nanostructures. Nano Today. 3, 38–46 (2008) and the references Dynamics of composition fluctuations in diblock copolymer solutions far from and
in it. near to the ordering transition. Macromolecules 27, 4762–4773 (1994).
123 Tanaka, T., Hocker, L. O. & Benedek, G. B. Spectrum of light scattered from a 137 Konak, C., Helmstedt, M. & Bansil, R. Temperature dependence of dynamics of
viscoelastic gel. J. Chem. Phys. 59, 5151–5159 (1973). solutions of triblock copolymer in a selective solvent. Polymer 41, 9311–9315
124 Panyukov, S. & Rabin, Y. Polymer gels: frozen inhomogeneities and density fluctua- (2000).
tions. Macromolecules 29, 7690–7975 (1996). 138 Nystrom, B. & Kjoniksen, A. L. Dynamic light scattering of a poly(ethylene oxide)
125 Pusey, P. N. Dynamic light-scattering by nonergodic media. Macromol. Symp. 79, poly(propylene oxide) poly(ethylene oxide) triblock copolymer in water. Langmuir 13,
17–30 (1994). 4520–4526 (1997).
126 Wu, C., Zuo, J. & Chu, B. Molecular-weight distribution of a branched epoxy polymer— 139 Konak, C., Fleischer, G., Tuzar, Z. & Bansil, R. Dynamics of solutions of triblock
1,4-butanediol diglycidyl ether with cis-1,2-cyclohexanedicarboxylic anhydride. copolymers in a selective solvent: effect of varying copolymer concentration. J. Poly.
Sci.B Polym. Phys. 38, 1312–1322 (2000).
Macromolecules 22, 633–639 (1989).
140 Raspaud, E., Lairez, D., Adam, M. & Carton, J. P. Triblock copolymers in a selective
127 Wu, C., Ju, Z. & Chu, B. Laser-light scattering studies of epoxy polymerization of 1,4-
solvent. 2. Semidilute solutions. Macromolecules 29, 1269–1277 (1996).
butanediol diglycidyl ether with cis-1,2-cyclohexanedicarboxylic anhydride. Macro-
141 Konak, C., Helmstedt, M. & Bansil, R. Temperature dependence of dynamics in
molecules 22, 838–842 (1989).
solutions of associating random copolymers. Macromolecules 31, 4639–4641
128 Shibayama, M. Spatial inhomogeneity and dynamic fluctuations of polymer gels.
(1998).
Macromol. Chem. Phys. 199, 1–30 (1998) and the references in it. 142 Alexandridis, P., Holzwarth, J. F. & Hatton, T. A. Micellization of poly(ethylene oxide)-
129 Panyukov, S. & Rabin, Y. Statistical physics of polymer gels. Phys. Rep. 269, 1–131 poly(propylene oxide)-poly(ethylene oxide) triblock copolymers in aqueous solutions—
(1996). thermodynamics of copolymer association. Macromolecules 27, 2414–2425 (1994).
130 Onuki, A. Theory of phase-transition in polymer gels. Adv. Polym. Sci. 109, 63–121 143 Moussaid, A., Munch, J. P., Schosseler, F. & Candau, S. J. Light scattering study of
(1993). partially ionized poly(acrylic acid) systems : comparison between gels and solutions.
131 Konak, C., Helmstedt, M. & Bansil, R. Dynamics in solutions of associating statistical J. Phys. II. 1, 637–650 (1991).
copolymers. Macromolecules 30, 4342–4346 (1997). 144 Krall, A. H., Huang, Z. & Weitz, D. A. Dynamics of density fluctuations in colloidal
132 Bansil, R., Nie, H. F., Konak, C., Helmstedt, M. & Lal, J. Structure and dynamics of gels. Physica A 235, 19–33 (1997).
solutions of a polystyrene-polybutadiene pentablock copolymer in a styrene-selective 145 Norisuye, T., Inoue, M., Shibayama, M., Tamaki, R. & Chujo, Y. Time-resolved dynamic
solvent. J. Poly. Sci.B Polym. Phys. 40, 2807–2816 (2002). light scattering study on the dynamics of silica gels during gelation process. Macro-
133 Zheng, H. & Teraoka, I. Dynamic light scattering from semidilute solutions of a molecules 33, 900–905 (2000).
styrene-acrylonitrile random copolymer. Macromolecules 34, 6074–6082 (2001). 146 Shibayama, M. & Norisuye, T. Gel formation analyses by dynamic light scattering.
134 Vogt, S., Anastasiadis, S. H., Fytas, G. & Fischer, E. W. Dynamics of composition Bull. Chem. Soc. Jpn. 75, 641–659 (2002).
fluctuations in diblock copolymer melts above the ordering transition. Macromolecules 147 Pusey, P. N. & Tough, R. J. A. Dynamic light-scattering, a probe of Brownian particle
27, 4335–4343 (1994). dynamics. Adv. Colloid Interface Sci. 16, 143–159 (1982).
Junfang Li was born in Hubei province of China in 1979. He received a BS in physical chemistry from the Department of Chemical
Physics at the University of Science and Technology of China (USTC) in 2002, and a PhD in the same university in 2007. His PhD thesis
work, under the direction of Professor Chi Wu, was about the using a combination of synthetic chemistry and laser light scattering to
study the dynamics and structures of polymer solutions and gel networks. After a 2-year postdoctoral fellowship in the Department of
Mechanics at USTC (2007-2009), he joined the Shanghai Institute of Organic Chemistry (SIOC), Chinese Academy of Sciences(CAS), as
a research assistant in the State Key Laboratory of Organometallic Chemistry. His current research interests include polyelectrolyte,
coordination polymerization, immobilizing of polyolefin catalyzer and characterizing of polyolefin synthesized by new type single-site
catalyst.
To Ngai was born in Fujian province of China in 1976. He received his BS in chemistry at the Chinese University of Hong Kong in 1999.
In 2003, he obtained his PhD in chemistry in the same university under the supervision of Professor Chi Wu. He moved to BASF
(Ludwigshafen, Germany) in 2003; first as a postdoctoral fellow for 1 year, then as the chemist in polymer physics division. In July 2005,
he went to Professor Lodge’s group in the Chemistry Department of the University of Minnesota, working on polymer blending. He
returned Hong Kong in 2006 and joined the Chemistry Department of the Chinese University of Hong Kong as a research assistant
professor and is now assistant professor. Currently, his research interests center on the designing, synthesizing and measuring
interactions between polymers, particles and soft materials.
Chi Wu was born in Wuhu, Anhui, China (1955). He graduated from the Department of Chemical Physics at the University of Science
and Technology of China in 1982. It was there that he received a rigorous training in physics, but in a chemistry department. After
obtaining his PhD in 1987 followed by a 2-year postdoctoral experience, both under the supervision of Professor Benjamin Chu in the
State University of New York at Stony Brook, he moved to BASF (Ludwigshafen, Germany) in 1989; first as an Alexander von Humboldt
Fellow for 1 year, then as the supervisor of BASF’s laser light-scattering laboratory. He joined the Chinese University of Hong Kong in
1992 and is now Professor of Chemistry. His research combines synthetic chemistry, polymer physics and molecular biology to design
and execute decisive experiments to answer important questions in macromolecules, colloids and biology. Among these, include the
development of nonviral vectors for molecular medicines; the nucleation of neuron-degenerative disease-related protein aggregation; the
design, synthesis and assembly of functional macromolecules; the dynamics and structures of polymer solutions and gel networks; and
molecular characteristic properties of intractable and special polymers.
Polymer Journal