Slow Mode Plateau in DLS PEG
Slow Mode Plateau in DLS PEG
A low-frequency plateau is often found in the rheological spectra of various kinds of semidilute
solutions of polymers and other colloids; also, many such solutions have been reported to show slow-
modes in their dynamic light scattering autocorrelation functions. Both these observations may lead to
the hypothesis of weak associative network structures built by the dissolved polymer chains or colloidal
building blocks. To challenge this hypothesis, we conduct a series of comparative studies on semidilute
solutions of poly(ethylene glycol) by using classical rheology as well as passive microrheology based on
dynamic light scattering, along with structural studies using static light scattering. Although we indeed
find a low-frequency plateau using classical shear rheology, even at elevated temperatures where
potential polymer aggregates should be broken, no such plateau is observed in any of our microrheology
experiments. Also, dynamic and static light scattering studies on the polymer solutions do not confirm the
presence of larger structural entities: no slow mode can be detected in the autocorrelation function of the
scattering intensity signal, and this signal is angle independent if the samples are purified by a thorough
Received 5th November 2018, procedure of filtration. Based on these findings, we conclude that the low-frequency plateau in classical
Accepted 23rd February 2019 rheology results is an instrument effect caused by erroneous recording of the phase angle, although the
DOI: 10.1039/c8sm02263a magnitude of the torque lies well within the resolution of the rheometer. We also conclude that slow
modes in dynamic light scattering on solutions of poly(ethylene glycol) are impurity-based artifacts rather
rsc.li/soft-matter-journal than due to actual associated structures.
2666 | Soft Matter, 2019, 15, 2666--2676 This journal is © The Royal Society of Chemistry 2019
View Article Online
Experimental section
stabilized by surface-adsorbed amphiphilic PEG chains, Materials and sample preparation
thereby acting as weak network nodes. In an earlier study, The material basis for our investigations is narrowly dispersed
evidence for such bubbles has been found in dilute PEG PEG samples with weight averaged molar masses of 1860 g mol1
solutions,12 and in the semidilute regime, these may become (Carl Roth, denoted PEG-2k) and 23 540 g mol1 (Sigma Aldrich,
a cause for weak network formation. Furthermore, the forma- denoted PEG-20k); in addition, we partly use a PEG sample with
tion of large PEG chain clusters due to interactions between the a nominal molar mass of 100 000 g mol1 (Alfa Aesar, denoted
hydrophobic CH2–CH2 units has been reported, and also the PEG-100k). These polymers are first purified by dissolution in
This journal is © The Royal Society of Chemistry 2019 Soft Matter, 2019, 15, 2666--2676 | 2667
View Article Online
dichloromethane and subsequently precipitated in cold diethyl aqueous solution is examined by DLS, yielding a value of
ether. After drying in a high vacuum, the PEG is dissolved in water, RH = 23.7 nm based on an angular-independent self-diffusion
filtered through a Whatman Anotop filter (25 mm, 0.02 mm pore coefficient of Ds = 9.16 108 cm2 s1. DLS-based microrheol-
size) and a Millex-LGs filter (200 nm pore size), and dried again. To ogy is then performed by loading the polymer solutions with
characterize the molar mass of the polymers and their dispersity, the gold–PEG core–shell nanoprobes at a low gold concen-
SEC measurements are carried out with a 1260 Infinity GPC/SEC- tration of 5 mg L1 to avoid multiple scattering and inter-
system (Agilent) equipped with an RI-detector and a Shodex pre- particle interactions. The microrheology samples are prepared
Published on 26 February 2019. Downloaded by North Carolina State University on 11/6/2023 4:03:42 PM.
column, along with Shodex OHpak SB 804, Shodex OHpak SB by dissolving the polymer in water, adding the gold, homo-
803,and Shodex OHpak SB 802.5 columns, with an eluent mixture genization, freeze drying, and redissolution of the sample in
of water/acetonitrile (70 : 30) at a flow rate of 1 mL min1. SEC 2 mL of water to obtain a homogenous distribution of the
calibration is carried out with poly(ethylene glycol) standards from nanoparticles in the viscous polymer solutions. All light scatter-
PSS, Germany. We determined the polydispersity index (PDI) of ing quartz glass cuvettes are thoroughly pre-cleaned with
1.03 (PEG-2k) and 1.07 (PEG 20-k). Aqueous solutions of the peroxymonosulfuric acid and aqua regia, and dust particles
poly(ethylene glycol)s with different concentrations below, at, and are removed by rinsing with hot acetone. The polymer solutions
above the overlap concentration (100, 200, 300, and 400 g L1 are then poured into clean cuvettes by slow filtration through
for PEG-2k; 35, 70, 150, and 200 g L1 for PEG-20k; 200 g L1 for syringe filters (Millex-LGs, 200 nm pore size and Whatman
PEG-100k) are prepared by dissolving the respective amount of PEG Anotop, 25 mm and 20 nm pore size). This filtration step is
in deionized water. The solutions are gently shaken at 40 1C for performed in a dust-free laminar flow box to obtain dust-free
three days to allow for complete dissolution of the polymer. The polymer solutions. Dynamic light scattering measurements are
overlap concentration c* is calculated as c* = (3Mw)/(4pNARG3) E performed on a Multi goniometer ALV-CGS-8F SLS/DLS 5022F
[Z]1,21 with the polymer molar mass Mw, the Avogadro number NA, with a Uniphase He/Ne Laser (25 mW, l = 632.8 nm) and a
the polymer-coil radius of gyration RG, and the intrinsic viscosity vertically polarized laser beam, with eight simultaneously working
[Z] = 12.5 103 Mw0.78.22 We calculated radii of gyration of ALV-7004 multi-tau correlators at eight fiber optical ALV/High QE
1.62 nm for PEG-2k, 5.13 nm for PEG-20k, and 11.46 nm for APD avalanche photodiodes (all components: ALV-Laser
PEG-100k, using the formula hRG2i1/2 = 61/2 8.88 102 Mw1/2.23 vertriebsgesellschaft mbH, Langen, Germany). A constant tem-
With these values, the overlap concentration c* was determined to perature of 20 1C for all measurements is assured (0.1 1C) by a
be 200 g L1 (PEG-2k), 35 g L1 (PEG-20k), and 11 g L1 (PEG-100k). Lauda ultra thermostat RKS C6 (Lauda Dr Worbser GmbH,
As a further characteristic quantity, the semidilute-solution correla- Königshofen, Germany). A Glan–Thompson polarizer with
tion length x, which can be seen as a mesh size of the transient vertical polarization (VV-mode) is installed in front of the detector
solution-network, is appraised as x = hRG2i1/2(c*/c)0.75 with the to block any light depolarized by the gold nanoparticles.26 The
polymer concentration c.24 For microrheology, probe colloids with measured and normalized intensity autocorrelation function,
a core–shell structure are used. The cores are spherical gold
hI ðq; tÞi hIðq; t þ tÞi
nanoparticles with a diameter of 15 nm purchased from Nano- gð2Þ ðq; tÞ ¼ (1)
hIðq; tÞi2
partzt, whereas the shell is a PEG (Mw = 5000 g mol1) layer. With
that, the probe colloids are significantly larger than the estimated is converted into the amplitude autocorrelation function Fs(q,t)
mesh size of the semidilute solution-network, so we expect compar- via the Siegert relation
ability between the microrheology and the macrorheology results.25
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Table 1 summarises the measured and calculated molecular and Fs ðq; tÞ ¼ gð1Þ ðq; tÞ ¼ gð2Þ ðq;tÞ 1 (2)
structural properties of the PEG samples.
with the angular- and time-dependent intensity I(q,t), and the lag
Dynamic light scattering (DLS) and microrheology
time t. In these equations, the scattering vector q is defined as
Before the actual microrheology experiments are performed, q = (4pnD/l)sin(y/2), where nD, y, and l, are the refractive index
the hydrodynamic radius of the gold nanoparticles in an of the solvent, the scattering angle, and the laser wavelength.
The amplitude autocorrelation function Fs(q,t) is the Fourier
transform of the van-Hove-autocorrelation function, describing
Table 1 Molecular and structural characteristics of the PEG-2k and
PEG-20k samples used in this work the time- and space-dependent particle motion as an ensemble.
The experimental autocorrelation data are fitted by an exponential
Mw Mn RG c* c xcalcd function g(1)(q,t) = Aexp(t/tR) for ideal diluted solutions of
Sample (g mol1) (g mol1) PDI (nm) (g L1) (g L1) (nm)
dispersed spherical particles, with the amplitude A and the
PEG-2k 1810 1860 1.03 1.62 200 100 2.72 characteristic relaxation time tR, and the diffusion coefficient D
200 1.62
300 1.20 is computed as D = (tRq2)1. The hydrodynamic correlation
400 0.96 length xH for polymer solutions in the semidilute concentration
regime can be calculated by a Stokes–Einstein analog relation
PEG20-k 23 540 22 080 1.07 5.13 35 35 5.13
70 3.05 kB T
150 1.72 xH ¼ (3)
200 1.39 6pZD
2668 | Soft Matter, 2019, 15, 2666--2676 This journal is © The Royal Society of Chemistry 2019
View Article Online
with the Boltzmann constant kB, temperature T, and the viscosity purified as described above. First, we investigate the linear
Z. For the microrheology conversion process, the time-dependent viscoelastic regime boundaries of every sample by implement-
mean-square displacement hDr(t)2i (MSD) of the probe colloids ing an amplitude sweep at a constant frequency of o = 1 rad s1
in the polymer solutions is extracted from a stretched biexponen- and amplitudes in the range of 0.01–100% deformation. Then,
tial fit function using the Einstein–Smoluchowski identity frequency sweeps are carried out in a frequency range of
hDr(t)2i = 6Dt and o = 0.1–100 rad s1, performed from high to low frequencies,
2 with a fixed amplitude within the linear viscoelastic regime
2 DrðtÞ
Published on 26 February 2019. Downloaded by North Carolina State University on 11/6/2023 4:03:42 PM.
Rheology
To compare the microrheology results to analogues obtained by
classical macrorheology, we performed classical shear experi-
ments using an Anton Paar modular compact rheometer of type
MCR 302 (Anton Paar, Graz, Austria) equipped with a cone–
plate measuring system, CP50-1/TG, with a cone radius of
Fig. 2 Frequency-dependent storage (squares) and loss moduli (circles)
24.983 mm and a cone angle of 0.9961 at 293 K. For all
of a PEG-20k solution (red) and a PEG-100k solution (blue), each at a
measurements, a solvent trap is used to suppress evaporation concentration of 200 g L1, measured macroscopically with a rheometer
of the solvent. To assure comparability, the same PEG samples at an amplitude well within the linear viscoelastic regime. The dashed line
as in the light scattering studies are used, prepared and represents the calculated low-torque limit of the instrument.
This journal is © The Royal Society of Chemistry 2019 Soft Matter, 2019, 15, 2666--2676 | 2669
View Article Online
determination of low torque values, the unexpected curve- rates, these network structures may be destroyed, leading to a
shape of the storage modulus G 0 at low frequencies can be transition to the normal terminal flow regime in the viscoelastic
caused by phase-angle uncertainties as an effect of the instru- spectrum.
ment, as comprehensively described by Velankar and Giles.30 To examine whether the low-frequency plateau has its origin
According to their work, the total error in G 0 can be expressed as in inaccurate phase-angle values or appears as a consequence
a combination of the error in the complex modulus |G*| and of a real underlying structural feature in the samples, and to get
the error in the phase angle: a closer insight into this structure, we performed microrheology
Published on 26 February 2019. Downloaded by North Carolina State University on 11/6/2023 4:03:42 PM.
Fig. 3 (A) Normalized DLS autocorrelation function g(1)(t) of an aqueous solution of PEG-20k (c = 70 g L1), containing PEG-coated gold nanoparticles
(d = 20 nm, Mw,PEG–corona = 5000 g mol1) as probe colloids observed at scattering angles of 301, 601, 901, and 1201 (black, red, green, and blue symbols),
along with fits and corresponding residuals to a bi-Kohlrausch–Williams–Watts (bi-KWW) function. (B) Angular dependence of the inverse relaxation
times, showing two different diffusive processes.
2670 | Soft Matter, 2019, 15, 2666--2676 This journal is © The Royal Society of Chemistry 2019
View Article Online
colloidal gold tracer particles. The experimental autocorrelation The second observed process is slower than the fast one by
data are all well fitted by a stretched biexponential function of two orders of magnitude and corresponds to the hindered gold
Kohlrausch–Williams–Watts type. tracer diffusion. With an increase in the polymer concentration
C ! E ! and therefore the solution viscosity, the velocity of the tracer
ð1Þ t t decreases due to greater obstruction by the surrounding med-
g ðtÞ ¼ A þ B exp þ D exp (8)
t1 t2 ium. Based on this principle, the viscoelastic properties of that
medium can be quantified from the tracer motion. As it is
Published on 26 February 2019. Downloaded by North Carolina State University on 11/6/2023 4:03:42 PM.
From this fit, two relaxation times, tslow and tfast, of the diffusion driven by the thermal energy kBT only, no external forces are
processes can be computed as applied, and hence, measurements in the linear viscoelastic
regime are always ensured; also, the risk of measurement-
t1 1 t2 1
tslow ¼ G and tfast ¼ G (9) induced destruction of the relevant microscopic sample struc-
C C E E
tures is excluded. The tracer mean-square displacement (MSD)
with the Gamma function G and the fit-function exponents C and is extracted from the DLS autocorrelation fit function and
E. Plotting the inverse relaxation times tslow1 and tfast1 versus converted into the complex shear modulus by a numerical
the square of the scattering vector q2 yields two straight lines inversion method that works without any need for a Fourier
through the origin, with slopes representing two different diffu- transform of the experimental data and thereby minimizes
sion coefficients, as can be seen in Fig. 3B. Fig. 4 shows the fast transformation errors (see Experimental part).
(filled squares) and the slow (open circles) diffusion coefficients Fig. 5 shows the resulting frequency-dependent data for the
depending on the polymer concentration of the PEG-20k samples. loss and storage modulus of a PEG-20k solution at c = 150 g L1,
We attribute the fast coefficient Dfast to the cooperative in comparison with the corresponding data from classical
motion of the overlapping PEG chain segments. This coopera- macroscopic shear rheology. We achieve good agreement
tive diffusion coefficient increases with the increase of the between the loss modulus data from both, macro- and micro-
polymer concentration in the semidilute region (c 4 35 g L1 rheology, in the overlapping frequency range from 101 to
for PEG-20k and c 4 200 g L1 for PEG-2k), as the freely 60 rad s1. In contrast, the low-frequency plateau in the storage
movable polymer segments between overlap points of the modulus G 0 does not appear in the microrheology data. As the
semidilute solution-network decrease in length if the polymer shear forces in microrheology are as weak as possible (only
concentration increases,24 thereby leading to a faster relaxa- thermal energy kBT drives the tracer particles), according to our
tion, whereas in the dilute regime, the fast diffusion coefficient above hypothesis of weak network structures in the PEG solu-
is nearly independent of the polymer concentration and can be tions, these should be more likely to be destroyed in macro-
described by the following equation:31 rheology than in microrheology, such that the latter method
should show them to be even more extended than the first
Dapp = Ds(1 + kcD)(1 + hRG2izq2) (10) method. The opposite is found, though. A possible reason for
the discrepancies in micro- and macrorheology could be inter-
with kD = 2A2Mw (RH3/Mw), the mean-square radius of
action effects between the colloidal gold tracers and the polymer
gyration hRG2iz, the second virial coefficient A2, the hydro-
chains, in particular depletion effects, whereby the gold tracers
dynamic radius RH, and the molar mass Mw. Extrapolation of
are able to create their own microenvironment resulting in a
the concentration c and the scattering vector q2 to zero results
in the self-diffusion coefficient Ds of the single polymer coils.
Fig. 4 Concentration dependence of the fast (filled squares) and the slow Fig. 5 Frequency dependence of storage and loss modulus (black and
(open circles) diffusion coefficients determined by DLS on aqueous solu- red) for a PEG-20k solution (c = 150 g L1), obtained by DLS-based
tions of PEG-20k. The grey bar marks the polymer overlap concentration c*. microrheology (lines) and by classical shear rheology (open symbols).
This journal is © The Royal Society of Chemistry 2019 Soft Matter, 2019, 15, 2666--2676 | 2671
View Article Online
2672 | Soft Matter, 2019, 15, 2666--2676 This journal is © The Royal Society of Chemistry 2019
View Article Online
scattering intensity Itotal consists of the sum of the scattering species (RH,3 = 540 nm), but this 0.79% still has a contribution
contributions of i species Ii, and via the normalized amplitude to the overall scattering intensity of about 62%. This estimate
ratios Ai, the intensity proportion of each species can be shows how crucial it is to thoroughly filter samples in DLS
estimated separately. analytics. Such filtration, however, removes material from
the sample.
X
N
Itotal ¼ Ii (13) To assure that a desired polymer concentration is used in
i¼1 DLS experiments along with assuring good sample purity, we
Published on 26 February 2019. Downloaded by North Carolina State University on 11/6/2023 4:03:42 PM.
1.78RH B M0.6 (18) Table 2 Cooperative diffusion coefficients and correlation lengths of
differently concentrated semidilute solutions of the polymers PEG-2k
The concentration portion xi of the i th species is then and PEG-20k, as obtained by DLS
obtained by
Polymer c (g L1) Dcoop (cm2 s1) xH (nm)
Ai ci
ci 0:6
and xi ¼ N (19) PEG-2k 200 2.09 10 6
1.03
1:78 RH;i P
ci 300 2.21 106 0.97
i¼1 400 2.34 106 0.92
From eqn (13)–(19) and the three hydrodynamic radii, we PEG-20k 35 7.57 107 2.84
estimate that only 0.79% of the sample consists of the largest 70 9.94 107 2.16
Fig. 8 (A) Experimental normalized amplitude autocorrelation function g(1)(t) of a PEG-20k solution (c = 70 g L1), recorded at eight different scattering
angles (301, 561, 641, 811, 981, 1151, 1241, and 1581), and residuals between a biexponential fit function and the raw data at the bottom. (B) The cooperative
diffusion coefficient Dcoop versus the square of the scattering vector q2 shows no angular dependence and yields a hydrodynamic correlation length of
xH = 2.16 nm.
This journal is © The Royal Society of Chemistry 2019 Soft Matter, 2019, 15, 2666--2676 | 2673
View Article Online
Fig. 9 Static light scattering results for PEG-2k (A) and PEG-20k (B) at three different concentrations, measured at 293 K with the experimental
parameters l = 632.8 nm, dn/dc = 0.135 mL g1 (PEG-2k) and 0.139 mL g1 (PEG-20k), and n = 1.3288.
2674 | Soft Matter, 2019, 15, 2666--2676 This journal is © The Royal Society of Chemistry 2019
View Article Online
This journal is © The Royal Society of Chemistry 2019 Soft Matter, 2019, 15, 2666--2676 | 2675
View Article Online
25 E. C. Cooper, P. Johnson and A. M. Donald, Polymer, 1991, 31 Light Scattering from Polymer Solutions and Nanoparticle Dis-
32, 2815–2822. persions, ed. W. Schärtl, Springer-Verlag, Berlin Heidelberg,
26 S. Balog, L. Rodriguez-Lorenzo, C. A. Monnier, B. Michen, 1st edn, 2007.
M. Obiols-Rabasa, L. Casal-Dujat, B. Rothen-Rutishauser, 32 (a) W. Burchard, M. Schmidt and W. H. Stockmayer, Macro-
A. Petri-Fink and P. Schurtenberger, J. Phys. Chem. C, 2014, molecules, 1980, 13, 1265–1272; (b) W. Burchard and
118, 17968–17974. W. Richtering, Prog. Colloid Polym. Sci., 1989, 80, 151–163.
27 J. J. Duffy, C. A. Rega, R. Jack and S. Amin, Appl. Rheol., 2016, 33 K. Devanand and J. C. Selser, Nature, 1990, 343, 739–741.
Published on 26 February 2019. Downloaded by North Carolina State University on 11/6/2023 4:03:42 PM.
2676 | Soft Matter, 2019, 15, 2666--2676 This journal is © The Royal Society of Chemistry 2019