Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

http://www.e-polymers.

org
e-Polymers 2013, no. 020
ISSN 1618-7229

Rheological measurement of molecular weight distribution


of polymers
Maryam Khak, Ahmad Ramazani S.A.*

Chemical and Petroleum Engineering Department, Sharif University of Technology,


Tehran, Iran; fax: (+98) 2166022853; e-mail: ramazani@sharif.edu
(Received: 02 November, 2009; published: 30 May, 2013)

Abstract: This paper has described a method to obtain the molecular weight
distribution (MWD) of polymeric materials from their rheological data. The method
has been developed for linear polymers with log normal molecular weight
distribution. The rheological data required to obtain the molecular weight
distribution are the shear storage modulus, G ' (ω) , and shear loss modulus, G " (ω ) ,
extending from the terminal zone to the plateau region. For determining the
molecular weight average, the method uses the relationship between stress moduli
and relaxation spectrums, with the equation that connects dynamic rheological data
with molecular weight distribution, and so it is not necessary to achieve the
relaxation spectrums and the molecular weight distribution is obtained directly from
dynamic shear experiments and it is one of the main advantageous of the
proposed method. Comparison of calculated and experimental data obtained by
GPC for five polypropylene samples produced in different conditions show that
model can correctly predict molecular weight distribution for these types of
polymers.
Keywords: Molecular weight distribution; Rheology; Polypropylene; average
molecular weight; GPC

Introduction
Molecular weight distribution (MWD) is one of the most important variables affecting
practical processing and performance of polymeric materials. Gel permeation
chromatography (GPC), light scattering and osmotic pressure are various methods
for molecular weight determination of polymers, which require dissolution of the
sample. However, many commercial polymers are only slightly or not soluble in
common solvents, and that can restrict the use of conventional analytical methods [1-
3]. In additional, the traditional analytical GPC is known to be insensitive to small
amounts of high molecular weight material that can have important effects on product
properties [4]. On the other hand, it has long been realized that rheological or
mechanical properties of polymer melts are strongly dependant on MWD with special
sensitivity to the high end of molecular weight. Therefore analytical methods based
on viscoelastic rheological properties of polymer melts may be more effective
methods compared to conventional ones [5-6]. Even for polymers which can be
dissolved easily, rheological measurements are capable of detecting precise
structural differences between samples that GPC is not capable of, such as long
chain branching [7]. In recent years much research has aimed to define the
relationship between polymer rheological information and MWD [8-18].
Two general models have been used for MWD determinations: the “Viscosity model”
and the “Modulus model”. Viscosity model [19-20] was based on the work of Bersted

1
and Slee [21] as well as Malkin and Teishev [22]. The model is derived from some
simple assumptions relating molecular structure to the viscosity flow curve.
Modulus model is more sensitive to low concentrations of high MW components
which usually dominate melt processability. The latter is especially true when
extensional flows are involved [23]. A generalized form of this model that represented
in terms of the relaxation modulus, G (t ) , is offered by Meier, et. al [24]:
β
G (t ) ⎡ ⎤
∞ 1


β
= ⎢ F F ( t , M )W ( M ) d (ln M ) ⎥ (1)
G No ⎢⎣ln( M e ) ⎥⎦
In this model G (t ) is normalized by the plateau modulus, G N0 . F (t , M ) is a kernel
function describing the relaxation behavior of a Monodisperse component of
molecular weight, M . Where the time is related to M by below equation:
t = k ( M ) 3.4 (2)
W ( M ) is the weight fraction MWD function. The exponent β , is a parameter which
characterizes the mixing behavior of the chains. For example β is 1 for simple
reptation and 2 for double reptation theory. M e is the entanglement MW. Other
authors have also used this model for calculating the MWD in simple reptation and
double reptation cases.
Wu was one of the first to attempt molecular weight distribution calculations from
rheological data [25-26]. He assumed single reptation and single exponential
behavior for relaxation function. Thus, β = 1 and F (t , M ) = exp[− t τ o ( M )] , the single
exponential kernel, were used in equation (1). This led to an adequate representation
of nearly Monodisperse samples but very poor representation of bimodal blends.
Tuminello assumed double reptation (β = 2) and oversimplifying step function
behavior for the kernel [27].
F (t , M ) = 1 If t p τ 0 ( M )
(3)
F (t , M ) = 0 For all other t
Although these assumptions predicted a falsely broad distribution for nearly
Monodisperse melts, the predictions were reasonably accurate for bimodal blends
and broad distribution polymers.
Mead improved the representation of nearly Monodisperse, broad and bimodal
distribution melts [28]. He used the double reptation assumption with a single
exponential kernel β = 2 .
One of the other works in this field was the monodisperse model of Doi and Edwards
based on the assumption that the surrounding matrix of polymer constitutes a time
invariant that can be approximated as a fixed tube where the polymer chain reptates
[29]. The relaxation modulus is proportional to the fraction of all chain segments not
relaxed after an instantaneous deformation.
G (t ) = G No .F (t ,τ e ) (4)
F (t ,τ e ) is the relaxation function and τ e is the reptation time.

2
Tsenoglou [30] and des Cloiseaux [31-33] have introduced the concept of double
reptation to model polydisperse systems. This model does not make distinction
between the reptation chain and the surroundings. It is based on the dynamics of
stress points formed by the entanglements. In this model, the relaxation modulus was
given by the following expression:
2
⎡ ⎤
G (t ) = G .⎢∑ wi .F (t , M i )1 / 2 ⎥
o
N (5)
⎣ i ⎦
where F (t , M i ) is the relaxation function for a Monodisperse system and wi is the
weight fraction of chains with molecular weight M i .
The objective of this work is to determine the molecular weight distribution of linear
polymers from rheological data. The model is based on a chain relaxation time
composed with an average molecular weight .

Theory
To determine the molecular weight distribution, the average molecular weights, M n ,
M w and M z , have been calculated using the existing relation between stress moduli
and relaxation spectrums [34]:
+∞ ω 2τ 2 + ωτ
G ' (ω ) + G " (ω ) = ∫ H (τ ) d ln τ (6)
−∞ 1 + ω 2τ 2
The above equation can be also presented as the following form:
H (τ ) ω 2τ 2 + ωτ
+∞ +∞ ω 2τ + ω
G (ω ) + G (ω ) = ∫
' "
. dτ = ∫ H (τ ) dτ (7)
−∞ τ 1 + ω 2τ 2 −∞ 1 + ω 2τ 2
Distribution of relaxation times or relaxation spectrums, H (τ ) , can be related to the
molecular weight distribution by following equation [16]:
G No
H (τ i ) = .M i .W ( M i ) (8)
β
where, G No is the plateau modulus and β is a parameter which determines the
contribution of the molecular weight of a chain in its relaxation time.
The relaxation time, τ i , of an i chain can be expressed as [16]:
α −β β
τ i = k .M .M i (9)

where M is an average molecular weight and α must be close to 3.4. For a


polymeric system with a definite molecular weight distribution, the relaxation time of
an i chain can be expressed as [35]:
τ i = k l M iβ (10)

k l is a parameter which will be presented in the following. Differentiation of the above


equation gives:
dτ = k l .β .M β −1 .dM (11)

3
l
Method of calculating k
After consideration of number average molecular weight, M n , instead of average
molecular weight, M , in Eq. (9) we have [16]:
τ i = k .M n α − β .M i β (12)
It is well known that the zero shear viscosity,η 0 , in polydisperse systems correlates
with the weight average molecular weight, M w , as follows [16]:
α
η o = G No .k .M w (13)
From above equation, k , equals to:
ηo
k= α (14)
G No .M w
After substituting the Eq. (14) in Eq. (12) we have:
α −β ηo α −β β
τ i = k .M n .M iβ = α
M n .M i (15)
G No .M w

From comparison Eq. (15) with Eq. (10), the parameter k l has been calculated finally.
α
ηo α −β ηo Mn
k =
l
α
.M n = α
. β (16)
G No .M w G No .M w M n
Whereas:

α α α
Mn ⎛Mn ⎞ ⎛ 1 ⎞
α
= ⎜⎜ ⎟ =⎜
⎟ ⎜ I ⎟⎟ (17)
Mw ⎝ M w ⎠ ⎝ P1 ⎠
As a result:
ηo
kl = α β (18)
G No .I P1 .M n

Distribution function
The molecular weight distribution, W (M ) can be shown in different distribution
functions such as Gaussian distribution (Normal distribution), Logarithmic normal
distribution (LN distribution) and Schulz-Flory distribution (Flory distribution) [36].
Gaussian distributions are symmetric about the median μ which equals the number-
average molecular weight. The differential number-distribution function of a Gaussian
distribution is:
1 ⎡ (M − μ ) 2 ⎤
W (M ) = exp ⎢− ⎥ (19)
σ 2π ⎣ 2σ 2 ⎦

4
Gaussian distributions are important for many polymer properties. Since they allow
negative values of properties, and negative degrees of polymerization do not exist,
they should not be applied to molar mass distributions. It does not matter, however,
for high degrees of polymerization and narrow distributions.
Logarithmic normal distributions (LN distribution) are Gaussian distributions with
natural logarithm of the property as variable instead of the property itself, it was
shown as:

1 ⎡ (ln M − ln μ )2 ⎤
W (M ) = exp ⎢− ⎥ (20a)
σ 2π ⎣ 2σ 2 ⎦
Or:
⎡ (ln M − μ )2 ⎤
1
W (M ) = exp ⎢− ⎥ (20b)
M 2πσ 2 ⎣ 2σ 2 ⎦
In fact, the logarithm normal distributions are a kind of normal distribution with
characteristics of the natural logarithm (ln). In other words, ln values of M and μ are
put instead of these parameters. LN distributions are skewed and not symmetric
about the median if μ instead of ln μ is chosen as the abscissa. In contrast to some
other types of distribution, they posses a maximum in the differential number
distribution. Therefore, in calculation of molecular weights, the distribution function is
assumed as a logarithmic normal distribution.
In these functions, σ 2 is variance and defined as below:
2
σ 2 = 〈 M 2 〉 − 〈 M 1 〉 2 = M w .M n − M n
2
σ 2 = M n ( I P1 − 1)
As mentioned above, μ is median and is defined as the first moment:

μ = 〈M 1 〉 = M n
Using Eqs. (8), (10), (11) and (20b) in Eq. (7) we have

+∞ G No 1 ⎡ (ln M − μ )2 ⎤ ω 2 k l M β + ω l
G ' (ω ) + G " (ω ) = ∫ .M . exp ⎢− ⎥. k β M β −1 dM
−∞ β M 2πσ 2
⎣ 2σ 2
⎦ 1 + ω 2 2l
k M 2β

(21)
+∞ 1 ⎡ (ln M − μ )2 ⎤ ω.k l .M β + 1 l β −1
= ∫ G No .ω. exp ⎢− ⎥. k M dM
−∞
2πσ 2 ⎣ 2σ 2
⎦ 1 + ω 2 2l
k M 2β

Then by putting equations of k l , σ 2 , and μ in Eq. (21), the below equation is derived
after simplification.

exp⎢−
(
⎡ ln M − M n 2 ⎤ ) ⎥ 1 + ω

⎜ ηo ⎞
⎟.M β
⎢⎣ 2 M n (I P1 − 1) ⎥⎦
2 ⎜ G o .I α .M nβ ⎟
+∞ ηo ⎝ N P1 ⎠
G ' (ω ) + G" (ω ) = ∫ ω. β
. . .M β −1 dM (22)
−∞ α 2
I P1 .M n 2
2π M n ( I P1 − 1) ⎛ ηo ⎞
1+ ω2⎜ ⎟ .M 2 β
⎜ G o .I .M nβ ⎟
⎝ N P1 ⎠

5
For normal distribution, following equation can be derived for determination of
average molecular weight:

(
⎡ M −Mn 2 ⎤
exp ⎢−
) ⎥ 1 + ω⎜
⎛ ηo ⎞
⎟.M β
⎢⎣ 2 M n (I P1 − 1) ⎥⎦
2 ⎜ α β ⎟
+∞ ηo o
⎝ G N .I P1 .M n ⎠
G ' (ω ) + G" (ω ) = ∫ ω. β
. . .M β dM
−∞ α
I P1 .M n
2
2π M n ( I P1 − 1) ⎛ ηo ⎞
2
(23)
1+ ω2⎜ ⎟ .M 2 β
⎜ G o .I .M β ⎟
⎝ N P1 n ⎠

Like the relationship mentioned in equation (6), there are separate equations for
storage and loss modulus which are:
+∞ ω 2τ 2
G (ω ) = ∫
'
H (τ ) d ln τ (24a)
0 1 + ω 2τ 2
+∞ ωτ
G " (ω ) = ∫ H (τ ) d ln τ (24b)
0 1 + ω 2τ 2
By using the previous equations and distribution function in Eqs. (24a) and (24b), a
similar equation can be derived for determination of number average molecular
weight.
Therefore, data of G ' (ω ) or G " (ω ) can be used separately for calculation of number
average molecular weight, but by using the totality of them more accurate results can
be obtained.

Experimental
Five polypropylene samples have been analyzed with the proposed method. The
samples were prepared in our laboratory with different catalysts and co catalysts. The
average molecular weights of all samples studied in this work are summarized in
Table 1. They were measured by PL-GPC 220 device in Arak polymer laboratories.

Tab. 1. Average molecular weight of analyzed samples.

sample Mn Mw Mz M w Mn
PP1 46100 251300 841000 5.45
PP 2 33000 191500 788200 5.80
PP3 60500 290500 974900 4.80
PP 4 90900 476200 1498000 5.24
PP5 42300 219500 866000 5.20

The dynamic shear storage and loss moduli of samples were measured by dynamic
rheometer from Para Physica using frequencies ranging from 0.1 to 100 Hz at 170
o
C. All measurements were performed into the linear viscoelastic region.
Figures (1-5) show the samples shear storage, loss moduli and shear rate viscosity
curves respectively.
The zero shear rate viscosity, η o , and the plateau moduli, G No , of all the studied
samples are shown in Table 2.

6
Fig. 1. shear storage G ' and loss moduli G " master curves for sample PP1 .

Fig. 2. shear storage G ' and loss moduli G " master curves for sample PP 2 .

Fig. 3. shear storage G ' and loss moduli G " master curves for sample PP3 .

7
Fig. 4. Shear storage G ' and loss moduli G " master curves for sample PP 4

Fig. 5. shear storage G ' and loss moduli G " master curves for sample PP5 .

Tab. 2. zero shear rate viscosity,η o , and plateau moduli , G No ,of samples.

sample η o ·10-2 G No ·10-4


PP1 3.1 3.98
PP 2 4.3 4.74
PP3 2.65 2.87
PP 4 4.1 3.89
PP5 3.9 4.10

Results and discussion


Computed master curves of G ' (ω ) + G " (ω ) for the five samples obtained by numerical
integration of equation (17) with the appropriate rheological parameters η o and
G No and accurate choice of parameters α , β . α = 3.4 and β = 1.9 were chosen as a
proper value for the analyzed samples. The zero shear rate viscosity, η o , and plateau

8
moduli , G No , values of samples are listed in Table 2. All numerical integrations were
performed by Matlab and Mathematica programs.

Tab. 3. Calculated Average molecular weight.

sample Mn Mw Mz M w Mn
PP1 51000 271000 1246000 5.32
PP 2 37000 205700 985000 5.56
PP3 56000 280500 1200000 5.01
PP 4 98000 503000 1650000 5.13
PP5 39200 212000 1100000 5.41

Fig. 6. Comparison of calculated MWD and GPC for PP1 .

Fig. 7. Comparison of calculated MWD and GPC for PP 2 .

It was possible to use G ' (ω ) or G " (ω ) data alone in calculations, but the sum was
more reliable. at low frequency range the sum remains correct because G " (ω ) is
accurate and G " (ω ) ff G ' (ω ) , at moderate frequency range both G ' (ω ) and

9
G " (ω ) remain reasonably accurate, moreover at high frequency range the sum also
remains correct because G ' (ω ) is now accurate and G ' (ω ) ff G " (ω ) .

Fig. 8. Comparison of calculated MWD and GPC for PP3 .

Fig. 9. Comparison of calculated MWD and GPC for PP 4 .

Fig. 10. Comparison of calculated MWD and GPC for PP5 .

10
The obtained values of the average molecular weights of samples are listed in table
3.
Eventually, performance of molecular weight distribution of samples in log-normal
case was done, according to relations between stress moduli and relaxation
spectrums with the equation that connects dynamic rheological data with molecular
weight distribution that was described before.
Molecular weight distributions of all samples are shown in Figures (6-10).

Conclusions
This article has described a molecular model to obtain the molecular weight
distribution of linear polymers from rheological data. The model is based on a chain
relaxation time composed with an average molecular weight. For determining the
molecular weight average, the method uses the relationship between stress moduli
and relaxation spectrums, with the equation that connects dynamic rheological data
with molecular weight distribution. This assumption yields that the relaxation
spectrum is directly proportional to M .W ( M ) . At the same time, it is assumed that
the MWD of polymers may be approximated with a log-normal distribution. Then, we
have an equation that connects dynamic rheological data with MWD and we do not
need to achieve the relaxation spectrum. The molecular weight distribution obtained
directly from dynamic shear experiments and it is one of the main features of this
method.
The model has been tested with five polypropylenes prepared in our laboratory in
different conditions. Shear storage and loss moduli master curves from terminal to
rubbery plateau region data has been used to adjust the polydispersity index.
Comparison of calculated and experimental data obtained by GPC for five
polypropylene samples, show that model can correctly predict molecular weight
distribution for these types of polymers.

References
[1] Tuminello, W. H.; Treat, T.A.; English, A.D. Macromolecules 1988, 21, 2606.
[2] Tuminello, W. H. Polym.Eng.Sci. 1989, 29, 645.
[3] Tuminello, W. H.; Buck, W. H.; Kerbow, D. L. Macromolecules 1993, 26, 499.
[4] Sugimoto, M.; Masubuchi, Y.; Takimoto, J.; Koyama, K. J. Polym. Sci. Phys.
2001, Ed 39, 2692.
[5] Van Ruymbeke, E.; Keunings, R.; Baily, C. Non-Newtonian Fluid Mech. 2002,
105, 153.
[6] Gahleitner, M. Polym. Sci. 2001, 26, 895.
[7] Janzen, J.; Colby, R. H. J. Mol. Structure, 1999, 485.
[8] Wu, S. Polymer Engng Sci.1985, 25, 122.
[9] Tuminello, W.H. Polym Engng Sci. 1986, 26, 1339.
[10] Mead, D.W. J. Rheol. 1994, 38, 1797.
[11] Wasserman, S.H. J. Rheol. 1995, 39, 601.
[12] Carrot, C.; Guillet, J. J. Rheol. 1997, 41, 1203.
[13] Maier, D.; Eckstein, A.; Friedrich, Cr.; Honerkamp. J. J. Rheol. 1998, 42, 1153.
[14] Léonardi, F.; Majesté, Jc.; Allal, A.; Marin, G. J. Rheol. 2000, 44, 675.
[15] Léonardi, F.; Allal, A.; Marin, G. J. Rheol. 2002, 46, 209.

11
[16] Llorens, J.; Rudé, E.; Marcos, R.M. J. Polym Sci. Part B. Polymer Phys. 2000,
38, 1539.
[17] Thimm, W.; Friedrich, Cr.; Marth, M. J. Rheol. 2000, 44, 429.
[18] Nobile, MR.; Cocchini, F. Rheol Acta. 2001, 40, 111.
[19] Tumonello, W. H.; Cudre’-Mauroux, N. Polym. Eng. Sci. 1991, 31.
[20] Shaw, M. T.; Tuminello, W. H. Polym. Eng. Sci. 1994, 34.
[21] Bersted, B. H.; Slee, J. D. J. Appl. Polym. Sci. 1977, 21.
[22] Malkin, A. Y.; Teishev, A. E. Polym.Eng.Sci. 1991, 31.
[23] Kasehagen, L. J.; Macosko, C. W. J. Rheology. 1998, 42.
[24] Maier, D.; Eckstein, A.; Friedrich, Cr.; Honerkamp, J. J. Rheology. 1998, 42.
[25] Wu, S. Polym. Mater. Sci. 1984, 50.
[26] Wu, S. Polym. Eng. Sci. 1985, 25.
[27] Tuminello, W. H. Polym. Eng. Sci. 1986, 26.
[28] Mead, D. W. J. Rheology. 1994, 38.
[29] Doi, M.; Edwards, S. F. J. Chem. Soc. Faraday. Trans, 1978, 74, 1789.
[30] Tsenoglou, C. In Viscoelasticity of Binary Polymer Blends, ACS Polym.Prepr,
1987, 28, 185-186.
[31] des Cloizeaux, J. Europhys Lett. 1988, 5, 437.
[32] des Cloizeaux, J. Macromolecules. 1990, 23, 3992.
[33] des Cloizeaux, J. Macromolecules. 1992, 25, 835.
[34] Ferry, J.D. “Viscoelastic Properties of Polymers”, 3rd Ed, New York, Wiley, 1980,
Chapter 3.
[35] Llorens, J.; Rude, E.; Marcos, R.M. Elsevier Science. Polymer. 2003, 44, 1741.
[36] Hans, G.E. ”An introduction to polymer science”. New York. ISBN 3. 1997, 527.

12

You might also like