You are on page 1of 8

Polymer 53 (2012) 3594e3601

Contents lists available at SciVerse ScienceDirect

Polymer
journal homepage: www.elsevier.com/locate/polymer

Modelling tie chains and trapped entanglements in polyethylene


F. Nilsson a, X. Lan a, T. Gkourmpis b, M.S. Hedenqvist a, U.W. Gedde a, *
a
KTH Royal Institute of Technology, School of Chemical Science and Engineering, Fibre and Polymer Technology, SE-100 44 Stockholm, Sweden
b
Borealis AB, SE-444 86 Stenungsund, Sweden

a r t i c l e i n f o a b s t r a c t

Article history: A Monte Carlo random walk model was developed to simulate the chain structure of amorphous layers in
Received 1 April 2012 polyethylene. The chains emerging from the orthorhombic crystal lamellae were either folding back
Received in revised form tightly (adjacent re-entry) or performing a random walk (obeying phantom chain statistics) forming
20 May 2012
statistical loops or tie chains. A correct amorphous density (ca. 85% of the crystalline density) was
Accepted 21 May 2012
Available online 8 June 2012
obtained by controlling the probability of tight folding. Important properties like fracture toughness
depend on the number of chains covalently linking together the crystalline regions. The model structure
was analysed with a novel numerical topology algorithm for calculating the concentration of tie chains
Keywords:
Polyethylene
and trapped entanglements. The numerical efficiency of the algorithm allowed molecular cubic systems
Tie chains with a side length of 100 nm to be readily analysed on a modern personal computer. Simulations showed
Trapped entanglements that the concentration of trapped entanglements was larger than the concentration of tie chains and that
the thickness of the amorphous layer (La) had a greater impact than the crystal thickness (Lc) on the tie-
chain concentration. In several other commonly used models, such as the HuangeBrown model, the
influence of trapped entanglements and the effect of the La/Lc ratio are neglected. Simulations using as
input the morphology data from Patel generated results in agreement with experimental rubber
modulus data.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction amorphous polyethylene [7], crystallization of polyethylene [8] and


the inter-crystalline layers of polyethylene [9e11]. Atomistic
The structure of a semi-crystalline polymer such as polyethylene processes involving relatively small molecules can be mimicked by
can on a limited length scale be viewed as a sandwich with alter- molecular dynamics simulation. Unfortunately, due to the high
nating layers of crystalline and amorphous regions. As sketched in computational cost, molecular dynamics models are not yet
Fig. 1, the chains in the amorphous regions can be either tight folds, capable of predicting the concentrations of tie chains and trapped
statistical loops, loose chain ends (cilia), fully amorphous chains, tie entanglements. Although molecular dynamics simulation can
chains or chains connecting via permanently entangled loops to the estimate atomistic processes containing a relatively small number
adjacent crystal, the latter being referred to as trapped entangle- of molecules, the statistics required to compute the concentrations
ments [1,2]. of tie chains and other entanglements by this simulation method is
Early recognition of covalent bridges (tie chains) between computationally very expensive.
crystal lamellae was due to electron microscopy studies by Keith Among the computationally inexpensive statistical tie chain
et al. [3]. The importance of tie chains for fracture toughness and models, the HuangeBrown model [5] is the most widely used. This
slow crack growth is generally recognized. Notable studies on this model predicts the fraction of tie chains (P) according to:
theme are due to Brown et al. [4,5] and Seguela [6]. Theoretical 0 1
models to predict properties of the amorphous interlayer are ZL  
1@ 4b3
usually based on molecular dynamics simulation, statistical models P ¼ 1  pffiffiffiffi r 2 exp b2 r 2 drA (1)
or Monte Carlo random walk models.
3 p
0
Molecular dynamics models and similar models that include the
dynamics of the atoms or groups of atoms have been used to study where b2 ¼ 1.5/(6.8 nl2) for linear polyethylene [12] with n back-
bone carbonecarbon bonds each with a bond length (l) of 0.153 nm
[12]. The critical length L is usually chosen as twice the crystal
* Corresponding author. Tel.: þ46 8 7907640; fax: þ46 8 208856. thickness (Lc) plus the amorphous layer thickness (La). Later
E-mail address: gedde@kth.se (U.W. Gedde). improvements of the HuangeBrown model include attempts to use

0032-3861/$ e see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.polymer.2012.05.045
F. Nilsson et al. / Polymer 53 (2012) 3594e3601 3595

speed. However, none of the currently presented models is


sufficiently realistic. They either oversimplify the amorphous
chain structure, ignore the effect of crystal thickness, simplify or
ignore the entanglements, assume non-branched chains of
constant or even infinite length, use small computational
domains, ignore the need to control the polymer density and
crystallization temperature or do not use realistic morphology
data. The aim of the present study was to improve the model
developed by Nilsson et al. [19] and to attempt to eliminate the
main shortcomings of the afore-mentioned Monte Carlo models.
The new model includes a method to control the amorphous
density and an advanced, realistic and efficient novel numerical
topology method for calculating the fraction of trapped chain
entanglements.

2. Models and methods

2.1. Model overview

In the model, the procedure for estimating the concentrations of


tie chains and trapped entanglements was divided in two consec-
utive steps. In the first step, parallel semi-crystalline layers were
built using a Monte Carlo method [19] allowing the positions of the
carbon atoms in the inter-crystalline layers to be stored for further
Fig. 1. Schematic description of the possible shapes of polymer chains entering the
analysis. In the second step, a novel numerical algorithm for
amorphous region between two crystalline layers: (a) tight fold, (b) statistical loop, (c)
loose chain end, (d) free chain, (e) tie chain, (f) trapped entanglement involving two calculating the degree of entanglement between pairs of loops was
statistical loops. used to determine the total fraction of entanglements as well as tie
chains. The algorithm was based on the observation that the only
finite integration limits [13], allowing a distribution in molar mass information needed to validate the existence of a knot is that
and short-chain branching [14], as well as chain entanglements describing the properties of the intersection points of the curves
[6,15]. Drawbacks of these mentioned statistical methods include when mapped on a plane.
the use of an arbitrary L, the assumption that L is equally influenced
by a change in La and in 2Lc, and the assumption that trapped 2.2. The semi-crystalline layer model
entanglements are always formed if two chains are sufficiently
close to each other. In this study we assumed that the crystal thickness (Lc), the
Monte Carlo simulations for predicting tie chain formations amorphous thickness (La), the crystallization temperature (T), the
include the Guttman model [16], which uses gamblers ruin statis- amorphous density (ra), the crystal density (rc) and the weight
tics to predict the shape of random walks of polymer chains in the average molar mass ðM W Þ as input parameters either already
amorphous interlayer. This enables both the average length of known or determined by experimental methods. Molar mass
statistical loops and the average length of the tie chains for infinite distributions could be used instead of M W and branches could be
chains to be estimated. Early on-lattice simulations were per- included, but these features were not incorporated in this study.
formed by Bryant and Lacher [17] to assess entanglements of The dimensions of the simulation box were chosen, as
simplified random walk paths using knot theory [18]. These Dx ¼ Dy ¼ 100 nm and Dz ¼ k(La þ Lc) where k is a positive integer
simulations ignored the influence of crystal thickness and constant. The axis z is the normal to the crystal fold surface, y is the
approximated each random walk path with a simple triangle or direction of crystal growth (b-axis) and x the direction along the
square. Nilsson et al. [19] developed a more realistic off-lattice crystal front orthogonal to the other two directions. Once the
Monte Carlo model for predicting tie chain concentration using periodical computational domain was set, the model calculated an
a phantom chain random walk model to simulate the chains in the orthorhombic grid, which described the positions on the crystal
amorphous interlayer. (001) surfaces where the simulated polymer chains could enter or
Several entanglement models have been previously suggested. exit.
Padding and Briels [20] developed a model for handling entangle- An atomistic model was generated as a sequence of chains
ments in course-grained simulations of polymer melts. They were starting in the amorphous region at random. Each atom was
modelling uncrosslinked polymers where no trapped entangle- assigned a set of internal coordinates defining the chain confor-
ments could ever be formed; their definition of an entanglement mation in terms of bond lengths (0.153 nm; [12];), bond angles
was thus simply a temporary crossing of two polymer chains. (109.47 ) and torsion rotations. Each chain in the model was built
Lacher et al. [21] performed early on-lattice random walk simula- using a stochastic Monte Carlo method, leading to a number of
tions of trapped entanglements between crystal layers and several different initial configurations, without taking into account the
authors including Michalke et al. [22] analysed on-lattice simula- excluded volume. Ignoring the excluded volume is based on the
tions with knot theory. An excellent review on trapped entangle- Flory principle that chains in the melt are unperturbed and their
ment models applicable to polymer networks was presented by structural characteristics can be considered to be the sum of all
Lang et al. [23]. So far, no off-lattice method for simulating and possible configurations of a single chain [24]. The three torsional
analysing trapped entanglements in semi-crystalline polymers conformations used were trans (T), gauge (G) and anti-gauge (G0 )
with realistic molar weights has been presented. with the dihedral angles 180 , 60 and 60 , respectively. The
Monte Carlo random walk models can potentially provide statistical weight of T was set to 1 and the statistical weights (s) of
a good compromise between physical realism and computational the two gauche conformers (G and G0 ) are:
3596 F. Nilsson et al. / Polymer 53 (2012) 3594e3601

  otherwise it continues its progression in the amorphous region.


E
s ¼ exp  (2) Typically, a search radius of 1 nm was used as a limit, but the model
RT
was not very sensitive to the exact value once it was larger than
where E is the potential energy difference between the gauche and about 0.5 nm.
the trans states, equal to 2.1 kJ mol1 [24], and R is the gas constant. When the chain exits a crystal surface, it can either make a tight
The first conformer was simulated according to these statistical fold by immediately reentering the crystal at any adjacent empty
weight factors. The probability for the next carbonecarbon bond in position or start a new random walk in the amorphous region. The
a polyethylene chain to have a specific dihedral angular state was probability of tight folds (pi) was controlled by the model through
determined according to the following probability matrix (statis- pi þ 1 ¼ pi þ (rA/rC)i  (rA/rC)AIM, so that the desired ratio of 0.852
tical weight matrix with each row being normalized, i.e. the sum of between amorphous and crystalline densities was obtained [12].
the elements in a row is equal to unity): The procedure continued until the positions of all carbon atoms in
2 3 the entire simulated chain were determined. New chains were
1 s s initiated in the same way until the maximum filling degree x was
6 1 þ 2s 1 þ 2s 1 þ 2s 7
T
6 7 reached. The filling degree was defined as the number of occupied
6 1 s 7
P ¼ 6
6 1þs 0 7 7G
grid positions on the crystal surfaces over total number of crystal
6 1þs 7 (3) positions. Since the numerical errors of the model were largest at
4 1 s 5 0
0 G either very low or very high filling fractions, x ¼ 0.5 was found to be
1þs 1þs
the optimal compromise. The positions of all carbon atoms were
T G G0 stored in a matrix and analysed by the entanglement algorithm. To
achieve good statistics and to allow for a good representation of all
where the rows correspond to the last decided bond state and the relevant conformational characteristics each individual setting
columns correspond to the adjacent not yet decided bond. As an corresponded to an average of J simulated systems, where typi-
example, if the previous state was G then the probability for cally J ¼ 100 was used.
entering T is given by 1/(1 þ s).
The random walk routine generated results relatively closely 2.3. The entanglement algorithm
resembling a polyethylene phantom chain; the characteristic ratio
for long polymer chains amounted to 6.5 at 413 K, which is close to The concentrations of trapped entanglements and tie chains
the experimental value at this temperature, 6.8 [25]. It should be were determined by analysing the carbon atom coordinate matrix
noted that the method used included two simplifications: (i) the with our novel knot algorithm. The stored amorphous chain
dihedral angles for G and G0 were set to þ60 and 60 . More segments were distinctly separated in the matrix and the tie chain
accurate values are 67.5 [26]. (ii) the statistical weight factors of concentration was calculated as the ratio of tie chains over the
GG0 and G0 G were set to 0. In practice the chain will adapt to number of filled grid positions. All free-chains, loose ends, tight
a distorted GG0 with the dihedral angles þ100 and 67 [26]. The folds and tie chains were removed at this stage because only
two simplifications are counteractive and the final result is statistical loops can form trapped entanglements. All loops with
reasonably close to the real phantom chain in terms of character- crystal entries on opposite crystal surfaces were compared in pairs
istic ratio. It is clearly possible to use alternative random walk with each other.
routines such as course-grained models [27]. Fig. 2a and b demonstrates the principle for the first step in the
If the simulated chain ‘touched’ one of the crystal surfaces, i.e. identification of the trapped entanglements. Each loop was pro-
came closer to the surface than one bond length in the z direction, jected onto the xy-plane and mapped onto a grid with a spacing
and the remaining chain length was longer than the crystal thick- chosen so that one step in the grid corresponded to a step slightly
ness, then the chain attempted to enter the crystal. If there were larger than 0.153 nm, i.e. the distance between two adjacent
still empty crystal grid positions on the surface sufficiently close to covalently bonded carbon atoms. All grid positions containing at
the entry, then the chain would enter the crystal at the closest node, least one carbon atom were saved as the number 1 in a sparse

Fig. 2. a,b. Sketches of pairwise comparison of two loops anchored in adjacent crystals. (a) 3D graph of the two loops. (b) projection of the loops onto a grid in the xy-plane. Chain A
is coloured turquoise and chain B is violet. The chains can only collide inside the red bricks, corresponding to positions where the added sparse matrix representations of the loops
are summed to ‘2’. Inside the red bricks, all chain segments on opposite chains must be tested in pairs for collision in order to find the coordinates of all possible intersection points
of the projected chains. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
F. Nilsson et al. / Polymer 53 (2012) 3594e3601 3597

matrix of integers. In order to create continuous lines, the nearest vector had been removed, the same concept was applied on the
grid positions adjacent to diagonal grid movements were also filled. second vector. This procedure was iterated until no changes
When two loop projections were compared, their sparse matrix occurred over a whole cycle. Only if the sign vectors could be
representations were added. The only regions that could poten- reduced to empty vectors were no knots present.
tially contain intersections were those with a sparse matrix sum The concentration of entanglements was obtained by first taking
equal to 2, which reduced the computational cost of the method the lower of the calculated number of entanglements and the
remarkably. double number of loops on the side with fewest loops involved in
Inside the areas with potential crossovers, all active line entanglements. This number was divided by the product of the
segments from the first chain were tested for crossovers against all number of filled positions and the filling fraction in order to take
the line segments from the second chains. This was made with into account the fact that the number of entanglements increases in
a standard lineeline collision detection method [28]. Inside the red proportion to the square of the filling fraction.
square in the example shown in Fig. 2b, two line segments of loop A
must be compared with four segments of loop B, resulting in 8
lineeline comparisons and two crossovers. For each chain and 3. Results and discussion
intersection point, the z-value and the contour length from the loop
starting point to the intersection were computed. For both chains, 3.1. Parameter study: effect of molar mass, crystal thickness, and
the sequence of intersection points was stored as a vector with (þ) amorphous layer thickness
if the chain passed above the other and () otherwise, resulting to
a sign vector of the form: v1 ¼ [þ   þ þ þ ]. If all the signs were In this section the effect of some important parameters on the
equal, no entanglement was present. Trapped entanglements were concentrations of tie chains (xtc) and trapped entanglements (xte) is
present if the number of elements in v1 was odd, or if it was even presented. Some of the values given to these parameters were not
and the number of (þ) was odd. Otherwise the self-crossings had to realistic, e.g. the ratio of the amorphous density to the crystal
be computed and a procedure for reducing the two sign vectors was density received clearly unrealistic values in some of the demon-
applied. strations. The key parameters were given the following standard
The first reduction step (Fig. 3a) was that two adjacent (þþ) or values unless otherwise stated: Nx ¼ Ny ¼ 100 nm, Lc ¼ La ¼ 10 nm,
( ) signs on curve A could be removed if there were no self T ¼ 300 K, x ¼ 0.5, rA ¼ 1 nm, ra/rc ¼ 0.852, J ¼ 100 and
intersections between p1 and p2 on A and if all intersection points M ¼ 140 kDa. All the parameters are defined in Section 2.2.
between p1 and p2 on B had the same sign. The second step (Fig. 3b) Fig. 4a presents the results of simulations using different values
was that two adjacent (þþ) or ( ) signs on curve A could be for the molar mass and ra/rc (note that the realistic value is 0.852).
removed if all self intersections on line A between p1 and p2 had the As expected, the tie chain concentration increased with increasing
same signs as p1 and p2 while all encircled intersections on line B molar mass (using the standard values for the other parameters)
had the opposite sign. The third and also final step (Fig. 3c) was that according to the expression:
two adjacent (þþ) or ( ) signs could be removed if all the con-
xtc ¼ a  b  expðc  log MÞ (4)
tained self intersection points had the property that their signs and
the signs of their encircled intersection points contained only one where a, b and c are constants. The concentration of tie chains
sign change, e.g. [þ þ  ] while the intersections on B had the increased in a linear fashion with increasing ra/rc in the range from
same sign. Additional reduction steps might be needed for 0.8 to 0.975. Note that the other key parameters were set at their
capturing every possible knot configuration, but these three steps standard values. The strict demand that ra/rc ¼ 0.852 was met by
definitely captured the vast majority of the configurations obtained having a high concentration of tight chain folds. This is relaxed by
with random walks. allowing higher ra/rc values leading to a decrease in the tight folds
When the first sign vector was traversed and all removable sign and increase in the concentration of all statistical amorphous
doublets including their corresponding intersections in the second chains, including tie chains.

Fig. 3. aec. Simplification steps 1e3 used in the entanglement algorithm.


3598 F. Nilsson et al. / Polymer 53 (2012) 3594e3601

Fig. 4. a). Tie-chain fraction as a function of Mw for different values of rA/rC. The tie-chain fraction increases with increasing values of both variables. (b). Tie-chain fraction as
a function of Mw and crystal thickness Lc. The tie-chain fraction increases with increasing Mw but decreases approximately linearly with increasing Lc. (c). Tie-chain fraction as
function of Mw and amorphous thickness La. The tie-chain fraction increases with increasing Mw but decreases approximately as 1/La with increasing La. (d). Tie-chain fraction (filled
symbols) and entanglement fraction (open symbols) as a function of Lc/La when the critical distance L ¼ 2Lc þ La ¼ 25 nm. Note that the influence of the entanglements is not
negligible and that the functions are not constants, as is assumed in Brown type models.

Fig. 4b shows that the tie chain concentration decreased in an entanglements were more numerous than the tie chains in the
approximately linear fashion with increasing crystal thickness for systems of low crystallinity and sufficiently high molar mass,
different constant molar masses from 14 to 1400 kDa (keeping the whereas systems with higher crystallinity were more dominated by
other key parameters at the standard values, including the amor- the tie chains. These simulations again indicated that the molecular
phous layer thickness). Fig. 4c shows that the tie chain concentra- continuity (expressed in the concentrations of tie chains and trap-
tion was inversely proportional to the thickness of the amorphous ped entanglements) was sensitive to the Lc/La ratio.
layer thickness (keeping the other key parameters at the standard
values, including the crystal thickness). These data are not in
accordance with the HuangeBrown model, because this model 3.2. Application of the model to real polyethylenes
does not predict that the ratio La/Lc has any impact on the tie chain
concentration. Experimental data for 11 linear polyethylenes with different
Fig. 4d presents data for the concentrations of tie chains and molar masses (Mw from 20 kDa to 750 kDa and densities at room
trapped entanglements from simulations using a constant value temperature between 938 and 976 kg m3) reported by DesLauriers
(25 nm) for the critical length L ¼ 2Lc þ La. The molar mass was varied and Rohlfing [14] were used as input in the simulations. The input
between 14 and 1400 kDa and the temperature was constant at data used in the simulations were obtained directly from these
417 K. The crystallinity was varied between 14 and 67% by changing authors or by further calculation based on the following equations:
the Lc/La ratio (but always keeping L constant at 25 nm). The tie chain  
concentration increased in an approximately linear fashion with r kg m3 ¼ 1074:8  24:1  log Mw (5)
increasing Lc/La. The concentration of trapped entanglements
showed a more complex dependence on Lc/La. For the longest chains
Tm ðKÞ ¼ 0:006423  r2 þ 12:72  r  5878 (6)
(M ¼ 1400 kDa), the entanglement fraction decreased almost line-
arly with increasing Lc/La. The systems built with molecules of
medium molar mass (M ¼ 140 kDa) showed a parabolic function and
the system built with molecules of a molar mass of 14 kDa displayed 0
0:624  Tm
a linear increase in the concentration of trapped entanglements with
Lc ðnmÞ ¼   (7)
0 T
Tm m
increasing Lc/La. The simulations indicated that the trapped
F. Nilsson et al. / Polymer 53 (2012) 3594e3601 3599

   
rc r  ra
wc ¼  (8)
r rc  ra

rc  Lc  ð1  wc Þ
La ðnmÞ ¼ (9)
ra  wc
DesLauriers and Rohlfing [14] provided molar mass data as well
as the following input data used: Tm 0 ¼ 418.5 K, r ¼ 852 kg m3 (at
a
20  C) and rc ¼ 1006 kg m3 (at Tm 0 ). Eq. (5) was obtained by linear

fitting to data reported by DesLauriers and Rohlfing [14]. Eq. (6) was
obtained by fitting a second order polynomial function to the data
of Patel et al. [12] and Mirabella and Bafna [29]. Eq. (7) is the
GibbseThomson equation. Eq. (8) relates mass crystallinity (wc) to
density (r) for polyethylene. Eq. (9) relates the thicknesses of the
crystal and amorphous layers to the mass crystallinity.
Fig. 5 shows xtc and xte as functions of molar mass using the
input data obtained from DesLauriers and Rohlfing [14]. Very low
molar mass polymers had essentially no tie chains or trapped
entanglements. Both xtc and xte increased strongly between 100 Fig. 6. Tie-chain and entanglement fractions as functions of crystallization tempera-
and 1000 kDa, above which the increase was only moderate (Fig. 5). ture. Data from Barham et al. [26] are used as input to the model. The entanglements
are effected more than the tie chains by crystallization temperature.
Both curves have a similar shape; the only difference being that, at
any given molar mass the concentration of trapped entanglements Fig. 7 shows the result of simulations based on experimental
was approximately twice the concentration of the tie chains. structural data provided by Patel et al. [12] for 13 polyethylenes
Fig. 6 shows the effect of crystallisation temperature in the case including both homopolymers (linear polyethylenes) and copoly-
of a polymer with a relatively low molar mass (32.1 kDa). Input data mers produced by single-site technology. Patel et al. [12] provided
for Lc were obtained from experimental data provided by Barham data for Lc, La and volume crystallinity (12e69%). In the same study,
et al. [30], who showed that the initial crystal thickness (Lc*) had the data for the rubber modulus for the same set of polymers were
following dependence on the degree of supercooling (DT): presented. Our simulation results were normalized and plotted
together with the normalized rubber modulus data and corre-
k1
L*c ðnmÞ ¼ þ k2 (10) sponding HuangeBrown simulation data. At low vc-values, our
DT model data agreed very well with the rubber modulus data, but at
where k1 and k2 are constants, and DT was calculated taking high vc it gave at least a better prediction than the HuangeBrown
0 ¼ 418.5 K for the equilibrium melting point [14]. The other
Tm model. The main reason for this is that the HuangeBrown type
variables used in the simulations reported in Fig. 6 have the same model erroneously neglects the effect of changes in Lc/La. The
values as those used in Fig. 5. Fig. 6 shows that xtc showed experimental data show, as depicted in Fig. 8, that even though the
a pronounced decrease with increasing crystallisation temperature critical length L diminished rapidly with vc, this was due only to
at high crystallisation temperatures (>380 K), whereas after crys- a decrease in Lc. The more important factor La decreased instead,
tallisation at temperatures below 380 K the polymers showed and this partly counteracted the increase in L. The effect would
essentially a constant xtc. The concentration of trapped entangle- probably be even more pronounced if the simulations could be
ments showed, interestingly, a gradual (but retarding) increase performed with correct molar weight distributions instead of using
with decreasing crystallisation temperature (Fig. 6). only the weight average molar mass value.

Fig. 5. Tie-chain and entanglement fractions as functions of Mw. Experimental Lc and Fig. 7. Normalized tie chain fraction as a function of polymer density plotted together
La data from DesLauriers and Rohlfing [14] are used. The influence of entanglements is with the Brown model [4,5] (;), the present model (-), and the normalized exper-
approximately twice as large as the influence of tie chains. imental rubber modulus data (B) [12].
3600 F. Nilsson et al. / Polymer 53 (2012) 3594e3601

Fig. 9. The stretch potential for tie chains and entanglements for one realistic polymer
system with polymer in-data from Patel et al. [12]. The stretch potential for trapped
Fig. 8. The experimental in-data from Patel et al. [12] used in Fig. 7. The Lc, La and L- entanglements is clearly larger than that for tie chains, indicating that one of the latter
values are plotted as a function of crystallinity; La decreases but Lc increases with has a greater impact on the mechanical properties of the polymer than one of the
increasing wc. former.

The mechanical properties are however not only affected by the


In the simulations reported in Fig. 9, the probability for tight
fraction of tie chains and trapped entanglements, but also on the
folds obtained with our model became 0.5912  0.0095. This can be
degree to which the chains are stretched. For a tie chain with end
compared with values reported by other authors, e.g. 0.744
points pA and pB, n links, link length l and jjvABjj ¼ jjpB  pAjj, the
according to Kumar and Yoon [31], 0.73 by Marqusee [32],
“stretch potential factor” is stc ¼ nl6/jjvABjj where the constant
0.50e0.80 according to Mansfield [33], 0.65 according to Hoffman
6 ¼ cos(90  109.47/2) ¼ 0.8165 is chosen so that s ¼ 1 for all-trans
et al. [34] and 0.2 according to Flory and Yoon [35]. Note that our
chains. A corresponding expression can be derived for two loops on
value includes only the tight folds that are not generated by random
opposite crystal faces by first calculating the minimum distance
walks; a value including all tight folds would be higher, close to 0.7.
corresponding to two triangles formed by chains in all-trans. The
triangles will have one common point pE. The end points of the
loop-forming triangle 1 are pA and pB while the end points of 4. Conclusions
triangle 2 are pC and pD. The midpoints are denoted
pMAB ¼ (pA þ pB)/2 and pMCD ¼ (pC þ pD)/2 and the vector between A Monte Carlo simulation method has been developed for
the midpoints vMM ¼ pMCD  pMAB. The distances between the modelling tie chains and trapped entanglements in lamellar semi-
points were calculated as jjvAEjj ¼ jjpE  pAjj, jjvBEjj ¼ jjpE  pBjj, crystalline polymers such as polyethylene. Particularly important,
jjvCEjj ¼ jjpE  pCjj and jjvDEjj ¼ jjpE  pDjj. The numbers of links in the a novel numerical topology algorithm for finding entanglements
chains are n1 and n2. If the proportions between the chain lengths between pairs of loops was included. The simulations revealed that
are conserved, pE can be found by solving Eqs. (11) and (12) the concentration of trapped entanglements is substantial and that
numerically with respect to d. they must have a major impact on the stress transfer between crys-
tals. The model predicted that the impact of the amorphous thickness
pE ¼ pMAB þ d  vMM ð0  d  1Þ (11) is greater than that of the crystal thickness on the molecular conti-
nuity between adjacent crystals expressed by the concentrations of
kvAE k þ kvBE k n tie chains and trapped entanglements. This explains why our model
¼ 1 (12)
kvCE k þ kvDE k n2 gives better predictions than the HuangeBrown model in the case of
experimental rubber modulus data.
Once the distances between pE and the loop end points have
been determined, the ‘stretch potential factor’ for entanglements
Acknowledgements
can be determined as:

6  ðn1 þ n2 Þ  l The Swedish Research Council (grant 621-2005-6138) and Bor-


ste ¼ (13) ealis AB, Sweden, are acknowledged for their financial support.
ðkvAE k þ kvBE k þ kvCE k þ kvDE kÞ
Stretch simulation results for one of the polymers reported by References
Patel et al. [12] are shown in Fig. 9. The following polymer
parameter values were used: Lc ¼ 9.2 nm, La ¼ 10.8 nm, M ¼ 99 kDa, [1] Gedde UW, Mattozzi A. Adv Polym Sci 2004;169:29.
[2] Nilsson F, Gedde UW, Hedenqvist MS. Comp Sci Tech 2011;71:216.
ra/rc ¼ 0.852 and T ¼ 414 K [12]. The following simulation [3] Keith HD, Padden FJ, Vadimsky RG. J Polym Sci Polym Phys 1966;4:267.
parameters were used: Dx ¼ Dy ¼ 100 and J ¼ 1000. A total of over [4] Brown N, Lu X. Polymer 1995;36:543.
one million tie chains and entanglements were analysed. The [5] Huang Y-L, Brown N. J Polym Sci Polym Phys 1991;29:129.
[6] Seguela R. J Polym Sci Polym Phys 2005;43:1729.
stretch potentials for the tie chains were in general smaller than for
[7] Mattozzi A, Hedenqvist MS, Gedde UW. Polymer 2007;48:5174.
the trapped entanglements, indicating that the average tie chain [8] Meyer H, Müller-Plathe F. Macromolecules 2002;35:1241.
has a greater impact on the mechanical properties of the polymer [9] Balijepalli S, Rutledge GC. J Chem Phys 1998;109:6523.
than the average trapped entanglement. In this simulated polymer [10] Balijepalli S, Rutledge GC. Comp Theo Polym Sci 2000;10:103.
[11] Wilhelmi JL, Rutledge GC. J Phys Chem 1996;100:10689.
system, the mean stretch potential was 6.92  0.19 for tie chains [12] Patel RM, Sehanobish K, Jain P, Chum SP, Knight GW. J Appl Polym Sci 1996;
and 8.38  0.36 for trapped entanglements. 60:749.
F. Nilsson et al. / Polymer 53 (2012) 3594e3601 3601

[13] Strebel JJ, Moet A. J Polym Sci Polym Phys 1995;33:1969. [26] Boyd RH, Phillips PJ. The science of polymer molecules. Cambridge: Cam-
[14] DesLauriers PJ, Rohlfing DC. Macromol Symp 2009;282:136. bridge University Press; 1993.
[15] Yeh JT, Runt J. J Polym Sci Polym Phys 1991;29:371. [27] Binder K. In: Colbourn EA, editor. Computer simulations of polymers. Harlow,
[16] Guttman CM, DiMarzio EA, Hoffman JD. Polymer 1981;22:1466. UK: Longman Scientific and Technical; 1994. p. 91e129.
[17] Sumners D. J Math Chem 1987;1:1. [28] Schneider P, Eberly D. Geometric tools for computer graphics. San Francisco:
[18] Bryant J, Lacher R. Topology Proc 1988;13:1. Elsevier; 2003.
[19] Nilsson F, Hedenqvist MS, Gedde UW. Macromol Symp 2010;298:108. [29] Mirabella FM, Bafna A. J Polym Sci Polym Phys 2002;40:1637.
[20] Padding JT, Briels WJ. J Chem Phys 2001;115:2846. [30] Barham PJ, Chivers RA, Keller A, Martinez-Salazar J, Organ SJ. J Mater Sci 1985;20:
[21] Lacher RS, Bryant JL, Howard LN. J Chem Phys 1986;85:6147. 1625.
[22] Michalke W, Lang M, Kreitmeier S, Göritz D. Phys Rev E 2001;64:12801. [31] Kumar SK, Yoon DY. Macromolecules 1989;22:3458.
[23] Lang M, Michalke W, Kreitmeier S. J Comput Phys 2003;185:549. [32] Marqusee JA. Macromolecules 1986;19:2420.
[24] Flory PJ, Yoon DY, Dill KA. Macromolecules 1984;17:862. [33] Mansfield ML. Macromolecules 1983;16:914.
[25] Flory PJ. Statistical mechanics of chain molecules. Munich, Vienna, New York: [34] Hoffman JD, Guttman CM, DiMarzio EA. Faraday Disc Chem Soc 1979;68:177.
Hanser Publishers; 1969. [35] Flory PJ, Yoon DY. Macromolecules 1984;17:862.

You might also like