32
1 Basic characteristics of uniaxial extension rheology: Comparing monodisperse and bidisperse polymer melts Yangyang Wang, Shiwang Cheng, and Shi-Qing Wang a) Department of Polymer Science and Maurice Morton Institute of Polymer Science, The University of Akron, Akron, Ohio 44325-3909 Synopsis We have carried out continuous and step uniaxial extension experiments on monodisperse and bidisperse styrene-butadiene random copolymers (SBR) to demonstrate that their nonlinear rheological behavior can be understood in terms of yielding through disintegration of the chain entanglement network and rubber-like rupture via non-Gaussian chain stretching leading to scission. In continuous extension, the sample with bidisperse molecular weight distribution showed greater resistance, due to double-networking, against the yielding-initiated failure. An introduction of 20 % high molecular weight (10 6 g/mol) SBR to a SBR matrix (2.4×10 5 g/mol) could postpone the onset of non-uniform extension by as much as two Hencky strain units. In step extension, the bidisperse blends were also found to be more resistant to elastic breakup than the monodisperse matrix SBR. Rupture in both monodisperse and bidisperse SBR samples occurs when the finite chain extensibility is reached at sufficiently high rates. It is important to point out here that these results along with the concept of yielding allow us to clarify the concept of strain hardening in extensional rheology of entangled polymers for the first time. a) Electronic mail: [email protected]

Basic characteristics of uniaxial extension rheology paper (selected PUB)/Basic... · been realized that polydispersity in the molecular weight distribution significantly affects

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

  • 1

    Basic characteristics of uniaxial extension rheology:

    Comparing monodisperse and bidisperse polymer melts

    Yangyang Wang, Shiwang Cheng, and Shi-Qing Wanga)

    Department of Polymer Science and Maurice Morton Institute of Polymer Science, The

    University of Akron, Akron, Ohio 44325-3909

    Synopsis

    We have carried out continuous and step uniaxial extension experiments on monodisperse

    and bidisperse styrene-butadiene random copolymers (SBR) to demonstrate that their

    nonlinear rheological behavior can be understood in terms of yielding through

    disintegration of the chain entanglement network and rubber-like rupture via

    non-Gaussian chain stretching leading to scission. In continuous extension, the sample

    with bidisperse molecular weight distribution showed greater resistance, due to

    double-networking, against the yielding-initiated failure. An introduction of 20 % high

    molecular weight (106 g/mol) SBR to a SBR matrix (2.4×10

    5 g/mol) could postpone the

    onset of non-uniform extension by as much as two Hencky strain units. In step

    extension, the bidisperse blends were also found to be more resistant to elastic breakup

    than the monodisperse matrix SBR. Rupture in both monodisperse and bidisperse SBR

    samples occurs when the finite chain extensibility is reached at sufficiently high rates. It

    is important to point out here that these results along with the concept of yielding allow

    us to clarify the concept of strain hardening in extensional rheology of entangled

    polymers for the first time.

    a) Electronic mail: [email protected]

  • 2

    I. INTRODUCTION

    Extensional rheological behavior of entangled polymer melts has been studied for

    several decades. A clear understanding of failure behavior and material cohesive

    strength in extensional deformation is important for material designs. In the 1970s and

    80s, Vinogradov and coworkers carried out extensive studies on the failure behavior of

    monodisperse entangled polymer melts in uniaxial extension [Vinogradov (1975);

    Vinogradov et al. (1975a, 1975b); Vinogradov (1977); Vinogradov and Malkin (1980);

    Malkin and Vinogradov (1985)]. During the same period, the failure behavior of

    various commercial polydisperse polymers were also studied by several teams [Takaki

    and Bogue (1975); Ide and White (1977, 1978); Pearson and Connelly (1982)]. It has

    been realized that polydispersity in the molecular weight distribution significantly affects

    the failure behavior in uniaxial extension [Takaki and Bogue (1975)]. Since most of the

    commercial synthetic polymers are polydisperse, a general understanding of the influence

    of polydispersity on failure behavior is of great industrial value.

    A first step towards a better understanding of the molecular weight distribution

    effect is to study bimodal blends where the behavior and dynamics of each individual

    component can be readily established. Most of the previous investigations of such

    systems [Minegishi et al. (2001); Ye et al. (2003); Nielsen et al. (2006)] focused on the

    "strain-hardening" characteristic during uniform extension, leaving the failure phenomena

    largely unexplored.

    Our recent studies on a series of monodisperse linear SBR melts [Wang et al.

    (2007b); Wang and Wang (2008)] have suggested that certain failure behavior of highly

    entangled polymers in rapid uniaxial extension is analogous to the shear inhomogeneity

    revealed by particle-tracking velocimetry [Tapadia and Wang (2006); Wang (2007)], and

    can be understood in terms of the disintegration of the chain entanglement network.

    Specifically, in both startup shear and extension, entangled polymers exhibit the same

    scaling characteristics associated with the yield point in the elastic deformation regime

    [Wang et al. (2007b)], which is the point where the shear stress and engineering stress

    peak. Beyond the yield point, structural inhomogeneity develops in the form of

  • 3

    non-uniform spatial distribution of chain entanglement. The purpose of this study is to

    demonstrate that the notion of yielding can be readily applied to explain some basic

    rheological characteristics of entangled bimodal blends in uniaxial extension, including

    their failure behaviors. Moreover, this concept of yielding allows us to clarify when

    strain hardening really happens in uniaxial extension of entangled polymers.

    II. MATERIALS AND METHODS

    The bimodal blends in the current investigation were made from three of

    near-monodisperse linear styrene-butadiene random copolymers (SBR) provided by Dr.

    Xiaorong Wang at the Bridgestone Americas Center for Research and Technology. The

    molecular characteristics of three SBR melts can be derived from small amplitude

    oscillatory shear measurements using a Physica MCR 301 rotational rheometer equipped

    with 25mm parallel plates. Two samples involve 20 wt. % of SBR1M in two different

    "matrices" of SBR240K and SBR70K respectively, and are labeled as 240K/1M (80:20)

    and 70K/1M (80:20) respectively. A third mixture has the composition of SBR240K

    and SBR1M given by 240K/1M (90:10). The small amplitude oscillatory measurements

    of the bimodal blends are shown in Fig. 1, from which some basic information is

    obtained as shown in Table I, where the equilibrium melt shear modulus Geq is

    determined as the value of G' at the frequency, at which G" shows a minimum. The

    Rouse relaxation time τR of the sample was estimated as τ/3Z, according to the tube

    model [Doi and Edwards (1986)]. Due to their slight microstructure and polydispersity

    differences, the terminal relaxation time τ of these samples does not scale with the

    molecular weight M as: τ ~ M3.4

    .

    TABLE I. The Molecular Characteristics of SBR Melts

    Sample Mn (kg/mol) Mw/Mn Geq (MPa) Z τ (s) τR (s)

    SBR70K 70 1.05 0.74 25 0.67 0.0089

    SBR240K 241 1.10 0.82 98 34 0.12

    SBR1M 1068 1.23 0.85 510 11000 7.2

  • 4

    103

    104

    105

    106

    0.01 0.1 1 10 100

    240K/1M (80:20) G'

    240K/1M (80:20) G"

    70K/1M (80:20) G'

    70K/1M (80:20) G"

    Sto

    rage/

    loss

    modulu

    s (

    Pa)

    Angular frequency (rad/s)

    0.002

    Figure 1 Small amplitude oscillatory shear measurements of the bimodal blends at room

    temperature. The storage modulus and the loss modulus are represented by the open and filled

    symbols respectively.

    The uniaxial extension experiments were carried out using a first generation SER

    fixture [Sentmanat (2004); Sentmanat et al. (2005)] mounted on the Physica MCR 301

    rotational rheometer. The failure behaviors of pure melts and blends were investigated

    in two different types of testing. One was the failure during startup continuous uniaxial

    extension at a constant Hencky strain rate . The other was the failure during step

    extension where the sample was suddenly subjected to a certain amount of strain. The

    specimen failure is video-recorded to allow post-experiment analyses.

    III. RESULTS

    A. Continuous Extension

    Two types of failure mode were observed during startup continuous extension of the

    pure SBR melts. During low-rate extension of SBR240K and SBR1M melts the sample

    breakage is found to initiate from non-uniform extension. At sufficiently high rates

    rubber-like rupture was found for SBR1M when the sample broke sharply and the

    original cross-sectional dimensions returned after rupture, indicating that no part of the

    specimen suffered much irrecoverable deformation (i.e., yielding or flow)

    Fig. 2(a) and Fig. 3 present the stress-strain curves of SBR240K and SBR1M melts

    at various rates. The experiments ending in rupture are represented by open symbols.

    In filled symbols we see that the engineering stress engr always exhibits a maximum such

    that (engr)max is linearly proportional to max. This characteristic has been reported

  • 5

    before and suggested to be a signature of yielding [Wang et al. (2007b), Wang and Wang

    (2008)], beyond which chains mutually slide past one another [Wang and Wang (2009)].

    The engineering stress is also presented as a function of time in Fig. 2(b) for

    SBR240K. It is easy to see that the previously reported scaling behavior in the elastic

    deformation regime σengr ~ t1/2

    is valid for SBR240K as well [Wang et al. (2007b)].

    The end of each stress-strain curve corresponds to the onset of non-uniform extension by

    visual inspection, i.e., by examining the recorded images of the stretched specimen. On

    the other hand, at high rates rupture of SBR1M truncates the monotonic rise in the

    engineering stress engr. In other words, engr never had a chance to decline before

    rupture. Note that the data at 1 s-1

    in Fig. 3 approaches the borderline between yielding

    and rupture.

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    0 1 2 3 4

    en

    gr (

    MP

    a)

    Hencky strain,

    15 s-110 s

    -16.0 s-1

    3.0 s-1

    1.0 s-1

    0.3 s-1

    0.1 s-1

    SBR240K

    max

    (a)

    0.01

    0.1

    1

    10

    0.01 0.1 1 10Time (s)

    (b)

    15 s-1

    10 s-1

    6.0 s-1

    3.0 s-1

    1.0 s-1

    SBR240K

    0.3 s-1 0.1 s

    -1

    -0.5

    en

    gr (

    MP

    a)

    Figure 2 Engineering stress as a function of (a) Hencky strain and (b) time for SBR240K at

    various strain rates. The straight dashed line in (a) provides an indication of how the

    engineering stress increases more weakly than linearly with the Hencky strain.

    0

    2.5

    5

    7.5

    10

    0 1 2 3 4

    6.0 s-1

    3.0 s-1

    2.0 s-1

    1.0 s-1

    0.6 s-1

    0.3 s-1

    Hencky strain

    en

    gr

    (M

    Pa)

    SBR1M

    neo-Hookean

    Figure 3 Engineering stress as a function of Hencky strain for SBR1M at various strain rates.

    The stretching, ending in yielding-initiated failure and rupture, is represented by filled and open

    symbols respectively. The dotted curve is the neo-Hookean formula of Eq. (1) with Geq = 0.85

    MPa from Table 1, showing exponential growth with the Hencky strain .

  • 6

    The engineering stress-strain curves for startup continuous stretching of bimodal

    SBR blends are shown in Fig. 4, Fig. 5 and Fig. 6 respectively. The shape of the

    engineering stress-strain curve for bimodal SBR blends is qualitatively different from that

    of the pure SBR melts. At various rates, all three blends, i.e., 240K/1M (90:10),

    240K/1M (80:20) and 70K/1M (80:20), exhibited an engineering stress-strain curve

    containing two maxima and only showed non-uniform extension beyond the second

    maximum. At sufficiently high rates, the uniaxial extension of 240K/1M (80:20) and

    70K/1M (80:20) was terminated abruptly without displaying a second maximum in the

    engineering stress when the specimens ruptured. The rate dependence of strains at

    yielding-initiated failure and rupture for SBR240K, 240K/1M (90:10), and 240K/1M

    (80:20) is shown in Fig. 7. The onset of both failures in the bimodal blends is

    significantly postponed at the highest four rates relative to those of the pure components.

    0

    0.5

    1

    1.5

    2

    0 1 2 3 4 5

    15 s-1

    10 s-1

    6.0 s-1

    3.0 s-1

    1.0 s-1

    0.3 s-1

    0.1 s-1

    240K/1M (90:10)

    Hencky strain

    en

    gr (

    MP

    a)

    Figure 4 Engineering stress as a function of Hencky strain at various strain rates for the

    240K/1M (90:10) blend.

    0

    0.5

    1

    1.5

    2

    2.5

    3

    0 1 2 3 4 5

    240K/1M (80:20) 15 s-1

    10 s-1

    6.0 s-1

    3.0 s-1

    2.0 s-1

    1.0 s-1

    0.6 s-1

    0.3 s-1

    Hencky strain

    en

    gr (

    MP

    a)

    Figure 5 Engineering stress as a function of Hencky strain at various strain rates for the

    240K/1M (80:20) blend. The stretching, yielding-initiated failure and rupture, is represented by

    filled and open symbols respectively.

  • 7

    0

    0.5

    1

    1.5

    2

    0 1 2 3 4 5

    70K/1M (80:20)

    15 s-1

    10 s-1

    6.0 s-1

    3.0 s-1

    2.0 s-1

    1.0 s-1

    0.6 s-1

    Hencky strain

    en

    gr (

    MP

    a)

    Figure 6 Engineering stress as a function of Hencky strain at various strain rates for the

    70K/1M (80:20) blend. The stretching, yielding-initiated failure and rupture, is represented by

    filled and open symbols respectively.

    1

    10

    0.1 1 10 100

    SBR240K

    240K/1M (90:10)

    240K/1M (80:20)

    Elastic rupture (80:20)

    Elastic rupture (0:100)

    Fail

    ure

    Hencky

    str

    ain

    Strain rate (1/s)

    Figure 7 Failure Hencky strain as a function of applied strain rate for SBR(240K) and the

    two bimodal blends. The solid symbols represent ductile failure through yielding. The

    dashed lines show the borderline between viscoelastic and elastic regimes for the two blends.

    The open symbols denote strains at rupture for the pure SBR1M (triangles) and the 240K/1M

    (80:20) blend (diamond).

    B. Step Extension

    The engineering stress as a function of time during and after step extension of the

    pure SBR1M melt is first presented in Fig. 8 at three different amplitudes all high enough

    to produce elastic yielding, i.e., failure during relaxation. The induction times, before

    which the specimens appear intact, are all longer than the Rouse relaxation time R of 7.2

    s. After this period, one portion of the specimen started to shrink in its dimensions,

    leading to the sample breakage. Fig. 9(a) shows the elastic breakup behavior of not only

    the pure SBR240K in solid symbols but also the bimodal SBR blends of 240K/1M

  • 8

    (90:10). Here the critical Hencky strain for the breakup is just beyond 0.6 for the pure

    SBR240K, which is slightly lower than the vinyl-rich SBR melts previously reported

    [Wang et al. (2007b)]. The critical strain for SBR 240K/1M (90:10) is apparently the

    same as the pure SBR240K melt, although the induction time to break is markedly longer.

    0.01 0.1 1 10 100

    1.20.90.75

    en

    gr

    (M

    Pa)

    Applied rate: 15 s-1

    strain

    SBR1M10

    0.1

    1

    0.01

    Time (s)

    Figure 8 Engineering stress as a function of time in step relaxation experiments for the

    SBR240K melt and the 240K/1M (90:10) blend.

    0.01

    0.1

    1

    10

    0.01 0.1 1 10 100

    0.9

    0.75

    0.9

    0.75

    240K

    240K/1M

    (90:10)

    Applied rate: 15 s-1

    Time (s)

    strain

    en

    gr

    (M

    Pa)

    (a)

    0.1

    1

    0.01 0.1 1 10 100 1000

    1.2

    0.9

    0.6

    240K/1M (80:20)

    Strain

    Time (s)

    Applied rate: 15 s-1

    No breakup

    Elastic breakup

    2

    en

    gr (

    MP

    a)

    (b)

    Figure 9 Engineering stress as a function of time in step relaxation experiments for (a) the

    SBR240K melt and the 240K/1M (90:10) blend, and (b) 240K/1M (80:20) blend.

    In contrast, the critical Hencky strain c for SBR 240K/1M (80:20) shifted to c = 0.9,

    significantly higher than c ~ 0.6 for the pure SBR240K. Thus, the bimodal blends are

  • 9

    also more resistant to breakup after a step extension than the corresponding pure

    monodisperse matrix.

    IV. DISCUSSION

    A. Continuous Extension

    A.1 Yielding and rupture of Monodisperse Melts

    Yielding and non-uniform extension

    During startup deformation at rates involving Weissenberg number Wi = , after

    the initial elastic deformation, the entanglement network is forced to disintegrate when an

    imbalance emerges between the elastic retraction force and intermolecular locking force

    [Wang et al. (2007a), Wang and Wang (2009)]. In other words, the tensile force

    represented by engr is expected to display a maximum. The present study confirms that

    pure monodisperse melts exhibit yielding in constant-rate uniaxial extension. The

    yielding is signified by the emergence of a maximum in the engineering stress as shown

    in Fig. 2(a). Consistent with the reported scaling behavior of yielding in simple shear

    [Boukany et al. (2009a)], the solid symbols of Fig. 2(a) and Fig. 3 also show a shift of the

    yielding strain (corresponding to the engineering stress maximum) εy to higher values

    upon increasing the applied rate. The broad maxima at lower rates in Fig. 3 are due to

    the large polydispersity in the molecular weight distribution of SBR1M.

    There is strong evidence that the observed sample necking is not due to an elastic

    Considère-type instability, advocated recently by Hassager and coworkers [Lyhne et al.

    (2009); Hassager et al. (2010)]: When set to a stress-free state after the engineering

    stress maximum, the specimen cannot return to its original dimensions [Wang and Wang

    (2008)]. The plastic deformation and the corresponding decline in the engineering

    stress are a result of the entanglement network disintegration in presence of the

    continuing extension. Apparently, the eventual outcome is, not surprisingly, that one

    portion of the stretched specimen reaches a point of network disintegration and

    irreversible deformation before the rest of the specimen does, resulting in necking or lack

    of uniform extension. It is interesting to note from Fig. 2(a) that the specimen can

    extend uniformly well after the yield point for the group of data involving the higher rates

  • 10

    of 6.0 s-1

    and above. This dividing line coincides with the boundary between the

    viscoelastic and elastic deformation regimes, given by the condition of the Rouse

    Weissenberg number WiRouse = τR ~ 1 [Wang and Wang (2008)]. In the viscoelastic

    regime, the entanglement network could undergo irreversible deformation even before

    reaching the engineering stress maximum, so that the specimen is more ready to fall apart

    beyond the stress maximum. Conversely, in the elastic deformation regime, the network

    would be largely intact and fully recoverable up to the yield point (engineering stress

    maximum) [Wang and Wang (2008)]. As a consequence, the onset of structural collapse

    is delayed markedly beyond the yield point.

    We close this subsection by emphasizing that nearly monodisperse linear melts such

    as SBR240K only show significant strain softening in the rate range bounded by WiRouse <

    10. Fig. 2(a) shows, consistent with previous analyses [Wang et al. (2007b), Wang and

    Wang (2008)], that engr initially grows with the Hencky strain = t linearly. This may

    be expected because even the neo-Hookean model prescribes such a linear response at

    small extensions:

    engr = G(1/2),

    which has a limiting form of engr ~ 3Gfor

  • 11

    III and IV. However, our recent study [Wang and Wang (2010a)] has revealed that for

    highly entangled polymers, such a yielding-to-rupture transition is located in the middle

    of the rubbery plateau region, and does not involve any transformation to the glassy state.

    In other words, it would be misleading to associate the rupture with the term “glass-like

    zone”, as classified in the Malkin-Petrie master curve.

    The ductile failure arises from yielding of the entanglement network, during which

    linear chains mutually slide past one another [Wang et al. (2007a); Wang and Wang

    (2009)], whereas the origin of the rupture could have something to do with chain scission.

    At the very high rates, the mode of yielding, i.e., chain mutual sliding, is no longer

    available when sufficient tension in the chains can build up to cause chain scission. In

    other words, chain scission would occur before chains have reached the point of force

    imbalance [Wang and Wang (2009) to allow mutual sliding. This appears to occur in

    SBR1M at an extensional rate = 2.0 s-1

    and above as shown in Fig. 3 where the tensile

    force (i.e., the engineering stress) rises monotonically. This rapid rise in the engineering

    stress at higher rates is plausibly due to non-Gaussian stretching [Wang and Wang

    (2010a)]. In the explored rate range, SBR240K cannot be stretched fast enough at room

    temperature to produce the non-Gaussian stretching that leads to the upward rise in

    tensile force and ends in rupture via chain scission. In passing, we note that some

    alternative theoretical explanation for specimen failure during uniaxial extension has

    been made in the literature [Joshi and Denn (2003, 2004)]. We do not elaborate on this

    theoretical study because it appears to have oversimplified the process leading to the

    yielding-initiated failure.

    A.2 Failure Behaviors of the Bimodal Blends

    Effect of the high molecular weight component on yielding

    We have seen that the incorporation of a small amount of SBR1M to SBR240K

    significantly delays the sample failure. For example, at = 6.0 s-1

    , the SBR240K melt

    would yield at ≈ 1.4 and eventually undergo ductile failure at ≈ 2.6. The presence of

    10% or 20% 1M SBR postpones the onset of the ductile failure, as shown in Fig. 4 and

    Fig. 5, to much higher strains. A closer examination shows that there exists a critical

  • 12

    rate, beyond which the strength of the blends is greatly improved. As evident from Fig.

    7, the failure behavior of the 240K/1M (90:10) blend is not so different until reaching c

    ~ 6.0 s-1

    . Similarly, c ~ 2.0 s-1

    can be found for the 240K/1M (80:20) blend.

    The relaxation times of the two individual components in the 240K/1M blends are

    widely separated. As indicated in a preceding subsection, the Rouse relaxation times of

    240K and 1M are different by a factor of 60, whereas the difference in their terminal

    relaxation times is even larger, approaching 330. The wide separation of relaxation

    times causes the high and the low molecular weight components to stay in different

    deformation regimes under a given applied strain rate. In other words, the incorporated

    SBR1M chains actually formed a second entanglement network, which is highly

    plausible given the huge separation of its relaxation time scale from that of the "matrix"

    SBR240K. Specifically, the number of entanglements per chain in such a second

    network can be estimated according to the empirical scaling law for the entanglement

    molecular weight: Me() ~ Me,0-1.3

    [Yang et al. (1999)]. At the volume fractions of 0.1

    and 0.2, the second network involves respectively 32 and 73 entanglement points per

    SBR1M chain.

    Our previous study on the scaling characteristics of yielding [Wang and Wang

    (2009)] has revealed, as also demonstrated by Fig. 2(a), that the onset of yielding moves

    to higher strains with increasing Rouse Weissenberg Number WiRouse. Thus, at a given

    rate, the second entanglement network formed by the SBR1M would yield at a much

    higher strain. The blend could retain its integrity well beyond the first engineering

    stress maximum where the (first) matrix network has collapsed. This evidently

    indicates that the second entanglement network made of SBR1M did not give in until

    much higher strains. Since the second network involves much greater entanglement

    spacing, the non-Gaussian stretching sets in at significantly higher strains around a

    Hencky strain of 3.0 as read from Fig. 5 [Wang and Wang (2010b)] than a Hencky strain

    of 2.0 read from Fig. 3 for the neat SBR1M. More discussion on non-Gaussian

    stretching is deferred to IV.D.4.

  • 13

    Rupture

    With only 10% SBR1M, the first blend did not undergo rupture in the explored rate

    range as shown in Fig. 4. It appears that at these rates the second network associated

    with 10% SBR1M never got extended sufficiently to suffer chain scission. As a

    consequence, it yields in the form of mutual chain sliding.

    It is truly remarkable that the two blends of 240K/1M (80:20) and 70K/1M (80:20)

    containing only 20 % of SBR1M would undergo rupture in the range of Hencky rates

    where the pure "matrices" only show yielding-initiated non-uniform extension at the

    same rates. Apparently at these rates there is not sufficient intermolecular gripping

    force to produce full chain extension and enough chain tension to cause chain scission in

    either pure SBR240K or SBR70K. Thus, the 80% matrix chains only yield in the form

    of mutual chain sliding. On the other hand, the second network due to SMR1M can be

    fully stretched to reach the point of chain scission. Apparently, the 80% matrix chains

    at the point of rupture have reached such a fully disengaged state that the creation of the

    new surface during rupture costs no more energy than that estimated from the surface

    tension. In other words, the incorporation of 20% SBR1M as a second network

    provided so much structural stability that the matrix, i.e., the first network made of

    SBR240K or SBR70K, was able to disintegrate fully.

    The rupture occurred at much higher strains in the blends than in the neat SBR1M

    because at the volume fraction of 0.2, a strand (made of SBR1M) between two

    neighboring entanglement points of the second network is of a much longer chain length.

    Our recent study has shown that a higher stretching ratio is required to produce full chain

    extension, which appears to be a necessary condition for chain scission [Wang and Wang

    (2010b)]. The difference between the open diamond and triangles in Fig. 7 indicates

    rupture at rather different stretching ratios due to the difference in entanglement density.

    Finally, it is more than interesting to note that the blend of 240K/1M (80:20) barely

    shows rupture at = 15 s-1

    , whereas the blend with SBR70K as the matrix would

    undergo rupture at a rate as low as = 6 s-1

    . We know that at the same rate SBR70K

    yields at a lower strain than SBR240K because the yield strain decreases with lowering

    WiRouse, which is lower for SBR70K. Consequently, SBR70K can reach a fully

  • 14

    disengaged state at a lower strain. As suggested above, the disappearance of any

    entanglement networking in the matrix appears to be a prerequisite for rupture.

    Apparently at = 6 s-1

    , when SBR1M suffers chain scission, the matrix is also already in

    a state of full disengagement, allowing rupture to take place in 70K/1M (80:20).

    Conversely, in 240K/1M (80:20) that is being extended at = 6 s-1

    , even when a

    significant fraction of SBR1M chains have become fully extended, evidenced by the

    upturn in the engr vs. curve of Fig. 5, and reached the condition for chain scission, and

    even if these chains do suffer scission, the matrix apparently has not fully collapsed at

    these strains. The engineering stress engr could further build up until most SBR1M

    chains have undergo chain scission. Beyond this point, the matrix chains further

    mutually slide past one another until the system lose uniform structural support from

    chain entanglement. This appears to be what is happening in the blend involving the

    stronger first network made of SBR240K.

    B. Step Extension

    Recent studies [Wang et al. (2007b); Wang and Wang (2008)] on the relaxation

    behavior of monodisperse entangled linear polymers after step uniaxial extension have

    revealed that the entanglement network suffers cohesive breakdown at large strains.

    Here the cohesion refers to that due to chain entanglement [Wang et al. (2007a)]. Such

    elastic breakup is analogous to those disclosed by the particle-tracking velocimetric

    observations of entangled polymer solutions [Wang et al. (2006); Ravindranath and Wang

    (2007)] and melts [Boukany et al. (2009b)] in startup simple shear. The dynamics of

    such elastic yielding appear to depend on the molecular weight of the polymer melt and

    the amplitude of the step strain. For both uniaxial extension [Wang et al. (2007b); Wang

    and Wang (2008)] and simple shear [Boukany et al. (2009b)], the elastic yielding takes

    place faster with increasing step strain amplitude. For the same amplitude of step strain,

    a sample with higher molecular weight takes a longer time to lose the cohesive integrity

    [Boukany et al. (2009b)]. In fact, the elastic yielding appears to be related to the Rouse

    chain dynamics in the sense that the induction time for the breakdown scales with the

    longest Rouse relaxation time [Wang et al. (2007a), Boukany et al. (2009b)]. Because

  • 15

    the Rouse relaxation time of SBR1M is 60 times of SBR240K (note that this ratio is far

    greater than Z2 ~ 25, due to the polydispersity of SBR1M), the incorporation of

    incorporation of SBR1M may delay the specimen breakup after step extension as shown

    in Fig. 9(a) for the 240K/1M (90:10) blend. Upon a step extension of Hencky strain of

    0.75 and 0.9, the pure SBR240K undergoes elastic yielding that leads to the sample

    failure at 3.5 and 10 s respectively as shown by the filled symbols in Fig. 9(a). The

    blend (open symbols) fails at appreciably later times of 6.6 and 13 s respectively. It

    appears that the specimen would not undergo failure until the second network associated

    with the high molecular weight SBR1M also collapse on some longer time scale.

    Something more dramatic occurs in the second blend containing 20% SBR1M.

    Here the presence of the SBR1M can prevent the sample from undergoing the cohesive

    failure. For a step extension of Hencky strain = 0.9, the blend does not suffer failure.

    Perhaps before the second network made of SBR1M undergoes any disentanglement the

    first network has already recovered its entanglement structure after experiencing elastic

    yielding. At the higher strain of 1.2, the disintegration of the second network perhaps

    occurred so quickly that the first network had no chance to return to its entangled state to

    provide any adequate structural support. Consequently, the sample failure was observed

    at this higher strain. Fig. 8 indeed shows that the breakup of the pure SBR1M

    accelerates from an induction of 28 s for the step extension of =0.9 to 11s for = 1.2.

    C. Elastic Instability or Yielding

    Over the past a few decades, there have been extensive studies of entangled polymer

    melts and solutions in extensional deformation, because of their theoretical and practical

    importance. Among all these studies, significant efforts have been made toward

    understanding the nature of specimen failure. Vincent (1960) was the first to use the

    Considère criterion [Considère (1885)] to analyze the necking instability in elongation

    and cold flow of solid plastics (PE and unplasticized PVC). Such an analysis was

    subsequently extended to polymer melts in the rubbery state [Cogswell and Moore (1974);

    Pearson and Connelly (1982)] and polymer solutions [Hassager et al. (1998); Yao et al.

    (1998)]. Doi and Edwards (1979) also suggested an elastic necking instability in their

  • 16

    original tube theory based on the Considère reasoning. McKinley and Hassager (1999)

    applied the Doi-Edwards theory and pom-pom model [McLeish and Larson (1998)] to

    predict the critical Hencky strain for failure in linear and branched polymer melts during

    fast stretching.

    0

    0.5

    1

    1.5

    2

    0 1 2 3 4 5

    15 s-1

    10 s-1

    6.0 s-1

    3.0 s-1

    15 s-1

    10 s-1

    6.0 s-1

    3.0 s-1

    Hencky strain

    240K

    240K/1M

    (90:10)

    en

    gr (

    MP

    a)

    Figure 10 Comparison of the engineering stress-strain curves for the SBR240K melt and the

    240K/1M (90:10) blend.

    In our view, the emergence of a maximum in the engineering stress implies

    weakening of the underlying structure. Beyond the maximum, the material is no longer

    the original elastic body. In rigid solids such as metals studied originally by Considère

    (1885), when the measured force passes a maximum, non-uniform extension take place

    immediately due to the very limiting amount of extensibility. The debate in the

    literature about how to apply the Considère criterion has been around the issue of whether

    necking could occur immediately after the force maximum [Barroso et al. (2010); Petrie

    (2009); Joshi and Denn (2004b)]. For rubbery materials including entangled melts, the

    force maximum occurs at a significant stretching ratio, and non-uniform extension

    usually does not occur right after the force maximum. Fig. 2(a) shows that the

    maximum occurs for = 15 s-1

    at a Hencky strain of = 1.8, and the failure strain is at

    3.4. This behavior renders the application of the Considère criterion irrelevant. The

    force peak is the yield point when the sample is weakening in its resistance against

    further external deformation. The entanglement network takes some further extension

    before disintegration [Wang and Wang (2009)]. In other words, uniform extension

    could persist for a while until the network structure of the sample becomes

    inhomogeneous. The eventual "necking" and failure are anything but a mechanical

  • 17

    (elastic) instability in our judgment. So far no theory can describe quantitatively how

    such a localization of yielding takes place. Any continuum mechanical depiction based

    on a non-monotonic relation between the tensile force and degree of extension would

    have to first explain the microscopic physics responsible for the tensile force decline with

    increasing stretching.

    More specifically, the phenomenon of the specimen failure after yielding is a result

    of localized cohesive failure, having to do with how the initial elastic deformation turns

    into plastic deformation in an inhomogeneous manner. Introduction of 10 or 20 %

    SBR1M long chains into the SBR240K produces similar engineering stress versus

    Hencky strain characteristics as the pure SBR240K up to some significant strains well

    beyond the maxima, as shown in Fig. 10. For example, at = 6.0 s-1

    , both the pure

    SBR240K and the blend show nearly identical stress-strain curve till = 2.7. A

    continuum mechanical analysis based on such curves would have predicted the onset of

    necking instability, i.e., non-uniform extension, at the same strain for both the pure

    SBR240K and its blend with SBR1M. This is far from truth: the presence of only 10 %

    SBR1M long chains significantly extended the range of uniform stretching till = 4.0,

    well beyond the strain where the necking and failure took place in the pure SBR240K.

    We assert that this suppression of non-uniform extension by the second entanglement

    network is something well beyond any existing theoretical description based on

    continuum mechanical calculations.

    More recently, Hassager and coworkers also tried to explain the delayed failure after

    rapid uniaxial extension (i.e., the elastic yielding according to Wang et al. (2007b)) on the

    basis of the Doi-Edwards model and a Considère-type analysis [Lyhne et al. (2009);

    Hassager et al. (2010)]. Since the observed elastic-yielding-initiated failures after step

    extension only involved a level of extension well below the tensile force maxima that

    occur around = 2.0, it is rather difficult to understand that this failure discussed in Fig.

    9(a)-(b) has something to do with the emergence of the non-monotonic engineering stress

    vs. strain curves shown in Fig. 2(a) and 5.

  • 18

    D. Where Is Strain Hardening

    Since strain hardening is a frequently invoked expression of extensional rheological

    behavior of polymer melts, we will discuss thoroughly this idea in the present context.

    In particular, we will review and comment on four ways in which the concept of "strain

    hardening" may get used to describe uniaxial extension of polymers.

    10-1

    100

    101

    102

    103

    104

    0 1 2 3 4 5

    106.03.02.0

    1.00.60.3

    Str

    ess

    (M

    Pa)

    Hencky strain

    Strain rate (s-1)

    240K/1M

    (80:20)

    neo-Hookean

    Figure 11 Cauchy stress as a function of Hencky strain for the 240K/1M (80:20) blend. The

    dotted line represents the stress-strain curve from the neo-Hookean model of Eq. (1) with Geq =

    0.82 MPa from Table 1.

    D.1 Strain hardening with stress-strain curve above neo-Hookean line

    One of the earlier references to strain hardening came from high extension of natural

    rubbers when the stress-strain curve appears above the prediction of the classical rubber

    elastic theory, or the neo-Hookean model for a network of Gaussian chains [Treloar

    (1944)]. This strain hardening at high stretching ratios is partially due to the finite chain

    extensibility, i.e., the system has reached the point of non-Gaussian stretching.

    Moreover, the strain-induced crystallization in natural rubber can also enhance strain

    hardening [Smith et al. (1964)]. However, entangled melts that do not undergo

    strain-induced crystallization, such as the present SBR melts and blends, only appear to

    show stress-strain curves beneath the neo-Hookean model prediction as shown in Fig. 3

    for the SBR1M and Fig. 11 for the 240K/1M (80:20) blend. Entangled polymer melts,

    in absence of chemical cross-linking, apparently always suffer so much loss of

    entanglements during extension that the overall stress-strain curve could not rise above

    the neo-Hookean line even when non-Gaussian stretching takes place. In other words, a

  • 19

    significant fraction of entangled points in the network reach the point of force imbalance

    and become ineffective to bear the load during continuous extension of entangled melts,

    which does not occur in the neo-Hookean model. More discussion on this point is given

    in the following IV.D.4

    D.2 Monotonic increase of stress in startup extension: false strain hardening

    At high Weissenberg numbers, startup deformation is known to produce stress

    overshoot in simple shear, which has recently been suggested to indicate the onset of

    yielding [Wang et al. (2007a), Wang and Wang (2009)]. Beyond the yield point, the

    shear stress or transient viscosity decreases because the elastic deformation ceases and

    disintegration of the entanglement network leads to reduced elastic resistance. In

    startup uniaxial extension, the so called true stress E, i.e., the product of the tensile force

    and stretching ratio = exp( t), would usually only monotonically grow with continuing

    extension until the point of non-uniform extension. In other words, in uniaxial

    extension, the extensional (Cauchy) stress E does not exhibit overshoot, and thus the

    transient viscosity E

    =/ at a given Hencky strain rate only monotonically rises.

    This contrast between simple shear and uniaxial extension has caused Münstedt and

    Kurzbeck (1998) to declare that "The viscosity increase as a function of time or strain,

    respectively, is called strain hardening. It is a special feature of elongational

    deformation of polymer melts."

    In our view, this difference between simple shear and uniaxial extension should be

    looked at in a different light. Upon startup deformation the transient shear viscosity also

    initially increases with time, which is just an indication that the elastic deformation is

    dominant at the beginning of startup shear. The elastic deformation cannot ensue

    indefinitely, and yielding must take place, leading to the observed decrease of the shear

    stress as well as viscosity over time. In contrast, during startup continuous extension at

    Hencky rate the dominant reason for the continuous rise in E is the simple geometric

    factor of the exponentially shrinking cross-sectional area A(t):

    = F/A(t) = engr = engr(t) exp( t), (2)

    where F is the total tensile force, and

  • 20

    A(t)=A0 exp( t) (3)

    is the time-dependent cross-sectional area with A0 being the original area. Even if

    yielding occurs, i.e., engr (t) starts to decline, and E

    would still only increase with

    time until the point of non-uniform extension and specimen failure, provided that the

    decline of engr due to yielding is not as fast as the exponential areal shrinkage, which is

    often the case. Thus, the continuous increase of E and E

    does not mean that the

    uniaxial extension has not experienced the same yielding as seen in simple shear. In

    other words, such increases do not mean that the sample’s entanglement structure is not

    deteriorating and becoming less resistant to the continuous extension. Actually, strain

    softening not hardening has taken place as long as the tensile force is declining with

    continuing extension. As far as we can tell, the point of the entanglement network

    weakening occurs at the engineering stress maximum and cannot be discerned readily

    from such quantities as the Cauchy stress and extensional viscosity.

    D.3 Upward deviation of transient viscosity from zero-rate limit

    There is yet another way in which “strain hardening” is referred to in uniaxial

    extensions of polymer melts. It refers to a specific phenomenon observed during startup

    uniaxial extension when the transient elongational viscosity shows upward deviation

    from the limiting zero-rate-viscosity vs. time curve, in contrast to the phenomenon in

    simple shear where the transient shear viscosity function is always below the zero-shear

    viscosity function. This is perhaps the most widely recognized signature for “strain

    hardening”. The phrase "strain hardening" might have first been used by Meissner

    (1975) in describing the uniaxial extensional behavior of low-density polyethylenes

    (LDPE) [Meissner (1971, 1975); Laun and Münstedt (1978)]. Linear melts with

    bidispersity in molecular weight distribution [Münstedt (1980); Koyama (1991);

    Münstedt and Kurzbeck (1998); Minegishi et al. (2001); Wagner et al. (2005); Nielsen et

    al. (2006)] and even monodisperse melts –stretched at high enough rates– [Wang and

    Wang (2008), and figures below] could also produce this upward deviation. Since

    LDPE shows the most pronounced upward trend, long chain branching has been thought

  • 21

    to play some peculiar role in producing this "strain hardening" [McLeish (2008); van

    Ruynbeke et al (2010)].

    In our view, this noted difference between simple shear and uniaxial extension is

    perhaps superficial rather than fundamental for entangled polymer melts. Entangled

    polymer melts have been found to exhibit only strain softening in simple shear as a

    consequence of yielding during startup deformation. The maximum shear stress max

    has been found to grow with the applied rate more weakly than linearly [Boukany et

    al. (2009a)] so that the peak transient shear viscosity +

    max= max/ can only decrease

    with . Moreover, the steady shear viscosity is always lower than the zero-shear

    steady-state viscosity: In the zero-rate limit, the entanglement network is intact, and it is

    the Brownian diffusion that brings the chains past one another. There is maximum

    viscous resistance to terminal flow because the equilibrium state of chain entanglement is

    preserved.

    In many cases, uniaxial extension is different in appearance only because the

    cross-sectional area A(t) of the sample keeps shrinking exponentially with time at a given

    Hencky strain rate as shown in Eq. (3). If one insists on representing the transient

    elastic response in terms of the Cauchy extensional stress of Eq. (2) and the transient

    extensional viscosity

    E

    (t) =/ = engr(t)exp( t)/ , (4)

    then the continuous increase of and E

    with time largely originates from the

    exponential factor associated with the shrinking cross-sectional area. Let consider the

    extension in the beginning in the sense that t < In the zero-rate limit, i.e., when

    Wi, we can estimate Eq. (4) by employing Eq. (1). When engr ~ Eand exp( t) =

    1, Eq. (4) turns into

    E

    (t)|0 = Et, for t 1, E

    (t)

    is also given by Eq. (5) at short times. But at longer times, the exponential factor exp(

  • 22

    t) in Eq. (4) is a rapidly rising function of time. This causes Eq. (4) to grow above Eq.

    (5) by the same exponential factor had engr(t) continued to rise linearly with the Hencky

    strain = t as shown by the dashed line in Fig. 2(a). In reality, the sample yields

    eventually. Thus, before the yield point when engr has not declined, the exponential

    factor in Eq. (4) kicks in to produce a higher E

    than the zero-rate curve. In other

    words, there is a small window of extension where the transient viscosity in Eq. (4) ticks

    upward as shown in Fig. 12. This upward deviation is largely due to the geometrical

    shrinkage of the transverse dimensions and is not true strain hardening.

    The less monodisperse sample of SBR1M can extend more before non-uniform

    extension terminates the experiments. In Fig. 3, although engr only grows by 40 % at

    = 1.0 s-1

    from = 3.0 to the point of failure, Fig. 13(a) shows a much stronger rise in

    E

    (t) because of the exponential decreasing function A(t) of Eq. (3) in the dominator of

    Eq. (2) for the Cauchy stress E. As a consequence, the data rise significantly above the

    zero-rate envelope. Actually, whenever significant extension occurs without sample

    failure, there would be a great contribution to E

    from the cross-sectional areal

    shrinkage that has little to do with strain hardening. Therefore, from now on we shall

    call this upward deviation of the transient elongational viscosity from its zero-rate curve

    "pseudo strain hardening".

    0.01

    0.1

    1

    10

    100

    0.01 0.1 1 10 100 1000

    15

    10

    6.0

    3.0

    1.0

    0.3

    0.1

    Time (s)

    3|*(1/)|

    Strain rate (s-1)

    SBR240K

    + (

    MP

    a.s)

    E

    Fig. 12 Transient extensional viscosity as a function of time for the SBR240K melt, where the

    "linear response" data given by ηE+ = 3 |η*(1/ω)| from the small amplitude oscillatory shear

    measurements are also presented as the reference [Gleissle (1980)].

  • 23

    Long chain branching (LCB) in entangled melts delays the onset of non-uniform

    extension. For example, LCB in LDPE allows such materials to display a very shallow

    maximum in the engineering stress. In other words, the engineering stress only

    decreases gradually beyond its maximum. This gives the geometrical exponential factor

    a large range of time or strain to boost the transient viscosity in Eq. (4) above the limiting

    zero-rate curve, and caused the phrase "strain hardening" to be invoked to differentiate

    the extensional rheological behavior of LDPE from that of other linear polymer melts

    such as high-density polyethylene [McLeish (2008)]. It is clear that LCB plays a critical

    role to prolong uniform extension relative to entangled melts made of linear chains.

    However, engr does typically decrease with increasing extension in LDPE. In other

    words, there is only evidence of yielding and strain softening and little sign of strain

    hardening. Actually, there is explicit evidence from birefringence measurements that

    LDPE does not suffer non-Gaussian chain stretching during extension in the typically

    explored range of strain rates and temperatures [Koyama and Ishizuka (1989); Okamoto

    et al. (1998)].

    Finally, we need to explain why most literature data on linear melts show little

    upward deviation from the limiting zero-rate curve, a fact that makes LDPE look

    somehow special. Most experiments on uniaxial extension of linear polymer melts have

    been conducted in the moderately high (rather than extremely high) rate regime, and few

    have been based on monodisperse samples. For example, up to = 3.0 s-1

    ,

    corresponding to Wi = 100, the data in Fig. 12 hardly tilted above the limiting curve, and

    some unimpressive upward deviation shows up only at the higher rates. As indicated

    above, E

    (t) would deviate exponentially fast above the zero-rate viscosity function if

    the sample would maintain linear growth of engr with the Hencky strain as shown by

    the dashed line in Fig. 2(a) In reality, the sample yields. The deviation of the actual

    data in Fig. 2(a) from this linear growth engr ~ E, i.e., the dashed straight line increases

    sharply with time. Beyond the engineering stress maximum, the deviation actually

    increases approximately exponentially, cancelling the exponential factor of exp( t) in Eq.

    (4) associated with the area shrinkage. Thus, at these intermediate rates, one can hardly

  • 24

    see any upward deviation in linear melts that readily suffers yielding. It is the yielding

    and the resulting non-uniform extension that makes it impossible to collect data points at

    longer times. As a consequence, it has been impractical to obtain the extensional

    viscosity in the fully developed steady flow state during startup continuous extension as

    noted by Petrie (2006). But Petrie did not know why it is almost impossible to reach the

    flow state [Petrie (2008)], which was due to uniform yielding as explained by Wang and

    Wang (2008).

    D.4 A case of “entanglement strain hardening”: non-Gaussian stretching

    The transition from elastic extension to flow, known as yielding in engineering

    terms, occurs during startup uniaxial extension in a wide range of rates when further

    overall elastic deformation of the entanglement network is no longer possible and the

    chains mutually slide past one another. At higher rates, chain extension could continue

    until a fraction of the chains in the sample approaches the finite chain extensibility limit.

    Chains at such a high stretching ratio appear stiffer and non-Gaussian. This is

    non-obvious from a conventional plot like Fig. 13(a). The mechanical evidence of this

    non-Gaussian stretching comes from further analysis of data such as those in Fig. 3. It

    is obvious that some level of yielding, i.e., mutual chain sliding, can and does occur

    before a fraction of the yield-surviving entanglement strands reaches the finite chain

    extensibility limit. Fig. 13(b) shows that when the shear modulus G in Eq. (1) is

    reduced from its equilibrium value of 0.85 MPa to 0.424 MPa, the neo-Hookean would

    emerge onto the data at the applied Hencky strain rate of 6.0 s-1

    at a stretching ratio of nG

    = 8.5, where the subscript nG stands for non-Gaussian. Beyond this turning point, the

    data deviate upward from the neo-Hookean curve, which can be taken as a sign of

    non-Gaussian stretching. This upward deviation is true strain hardening at the chain

    level, which we shall call "entanglement strain hardening" to differentiate from the strain

    hardening in vulcanized rubbers that produces a stress-strain curve above the

    neo-Hookean limit as discussed in IV.D.1. We see this behavior as shown in the open

    symbols in Fig. 3 for the SBR1M melt and in Fig. 5 and 6 for the binary mixtures. The

    full chain extension leads to chain scission and rupture during startup continuous

  • 25

    extension. One key characteristic for this strain hardening is the emergence of the

    upturn seen in Fig. 3 in open symbols in the pure SBR1M, and more instructively in Fig.

    13(a).

    10-2

    10-1

    100

    101

    102

    103

    10-2

    10-1

    100

    101

    102

    103

    Time (s)

    + (

    MP

    a.s)

    E

    3|*(1/)|

    SBR1M

    (a) = 1 s-1

    0

    1

    2

    3

    4

    5

    6

    7

    4 8 12 16 20

    en

    gr

    (M

    Pa)

    1

    SBR1M

    6.0 s-1

    neo-Hookean

    (b)

    nG

    Figure 13 (a) Transient extensional viscosity as a function of time for the SBR240K melt, where

    the "linear response" data given by ηE+ = 3 |η*(1/ω)| from the small amplitude oscillatory shear

    measurements are also presented as a reference [Gleissle (1980)].

    (b) Engineering stress engr versus the stretching ratio at = 6.0 s-1

    , relative to a neo-Hookean

    curve of Eq. (1) based on G = 0.424 MPa = Geq/2, i.e, half of the equilibrium shear modulus,

    implying that half of the strands in the equilibrium entanglement network are lost at the

    stretching ratio nG ~ 8.5.

    10-1

    100

    101

    102

    0.01 0.1 1 10 100 103

    106.03.02.01.00.60.3

    240K/1M

    (80:20)

    3|*(1/)|

    Strain rate (s-1)

    Time (s)

    + (

    MP

    a.s)

    E

    (a)

    0

    0.5

    1

    1.5

    2

    2.5

    10 20 30 40 50

    10

    6.03.02.0

    70K/1M (80:20)

    Strain rate (s-1

    )

    en

    gr (

    MP

    a)

    neo-Hookean

    (b)

    nG

    1

    Figure 14 (a) Transient extensional viscosity as a function of time for the 240K/1M (80:20)

    blend, where the "linear response" data given by ηE+ = 3 |η*(1/ω)| from the small amplitude

    oscillatory shear measurements are also presented as a reference [Gleissle (1980)].

    (b) Engineering stress engr versus the stretching ratio at the various rates for the 70K/1M

    (80:20) blend, relative to a neo-Hookean curve of Eq. (1) based on G = Geq2.2

    = 0.025 MPa

    where Geq = 0.85 MPa and = 0.2, where nG ~ 23, far higher than that of 8.5 for the pure

    SBR1M in Fig. 13(b).

    This upturn also shows up strongly in Fig. 5 and 6 that depict the two blends with

    the 20% long chains of SBR1M. It occurs whenever the limit of finite chain

    extensibility is approached to cause non-Gaussian stiffening of the entanglement network.

  • 26

    When expressed in terms of the transient viscosity E

    (t) as shown in Fig. 14(a),

    significant upward deviation from the zero-rate curve shows up, reminiscent of the data

    of LDPE. This strong upward deviation arises from both the exponential factor in Eq.

    (4) and the “entanglement strain hardening”, i.e., non-Gaussian stretching. For LDPE,

    there is little evidence of non-Gaussian stretching [Koyama and Ishizuka (1989);

    Okamoto et al. (1998)], yet the similar behavior is observed for the following reason:

    LDPE typically can undergo significant extension without encountering non-uniform

    extension despite the occurrence of yielding, signified by the emergence of a

    non-monotonic relation between the engineering stress and strain. The prolonged

    uniform extension allows the geometric exponential factor in Eq. (4) to produce the

    upward deviation that is well documented in the literature [Ferry (1980); Laun and

    Schuch (1989)].

    Actually, it is again more instructive to present the evidence of non-Gaussian

    stretching in the blends by referring to a hypothetical neo-Hookean behavior of the

    second network formed by SBR1M chains. The second network at a weight fraction of

    20 % has a shear modulus given by G = G(=1)2.2

    . Fig. 14 (b) shows that the data at

    high rates show significant upward deviation from the neo-Hookean curve. The

    deviation occurs at nG = 23, which significantly higher than the degree of extension

    given by nG = 8.5 for the pure SBR1M in Fig. 13(b). The separation between the blend

    and SBR1M is once again exactly a factor of 1 Hencky strain, as noted above due to the

    difference in entanglement spacing. The data at = 3.0 s-1

    and below are below the

    neo-Hookean curve, indicating that there is significant loss of entanglement due to

    yielding of the second network. The strands at these lower rates can hardly reach the

    fully extended chain limit to cause chain scission. The sample eventually fails by

    mutual chain sliding to reach a state of disengagement.

    In summary, the transient viscosity or Cauchy stress involves an exponentially

    decreasing area in the dominator of its definition so that it tends to grow in time and

    becomes greater than its value in the zero-rate limit, in contrast to the counterpart in

    simple shear where the sheared area stays constant. This geometric difference has

  • 27

    caused considerable confusion in the literature. We have examined four situations

    where the phase “strain hardening” may have emerged to characterize the extensional

    rheological behavior of entangled polymers with either linear or branched chain

    architecture. In all cases, the definition of the extensional transient viscosity permits the

    exponentially shrinking cross-sectional area to mask the origins of the physical

    phenomena. Systems that resist yielding and structural failure, such as long-chain

    branched LDPE and samples with bimodal molecular weight distribution, simply show

    greater upward deviation from its zero-rate linear response because at the same moment

    during extension the high rate test is in a more stretched state with a thinner cross-section

    than a limiting zero-rate test. This upward deviation may not imply strain hardening at

    all. The real strain hardening involving non-Gaussian stretching amounts to having an

    engineering stress that grows monotonically with increasing extension and therefore

    resist any non-uniform stretching or necking.

    V. CONCLUSION

    Ductile failure after yielding and rupture after non-Gaussian stretching have both

    been shown to occur in uniaxial extension for monodisperse and bidisperse entangled

    styrene-butadiene rubber (SBR) melts. Within the accessible range of Hencky strain

    rates, the internal chain dynamics of the SBR240K are too fast to allow full chain

    extension and rupture, whereas the SBR1M undergoes cohesive failure in the form of

    yielding and non-uniform extension at low rates, and rupture at high rates. The

    incorporation of a small fraction of SBR1M into a matrix of SBR240K greatly alters the

    characteristic responses to both startup extension and step extension.

    In particular, we have reached the following conclusions. (A) There is evidence of

    double entanglement networking. Following the disintegration of the faster network

    formed by the matrix chains, the second network made of SBR1M can retain the

    structural integrity of the specimen until it also subsequently yields. The onset of

    structural failure is considerably extended beyond that of the pure SBR240K matrix.

    The presence of the SBR1M also altered the kinetics of elastic yielding and even delayed

    the onset of elastic yielding in the blend of 240/1M (80:20). (B) More surprisingly, in

  • 28

    the same range of extensional rates, under which the pure SBR240K and SBR70K only

    fail through ductile yielding, the blend of 240/1M (80:20) and 70K/1M (80:20) suffer

    rupture. (C) The application of the Considère criterion is irrelevant because the origin

    of non-uniform extension appears to be yielding of the entanglement network, unrelated

    to any type of elastic instability. (D) We confirm our previous assertion [Wang and

    Wang (2008)] that entangled linear polymers cannot attain steady flow during startup

    uniaxial extension. In other words, such linear chain systems as the present pure SBR

    melts and their blends fail after yielding over a wide range of rates in the form of

    non-uniform extension without ever reaching a fully developed flow state. (E) At

    various extensional rates beyond the terminal regime, the monotonic rise of the Cauchy

    stress before the onset of non-uniform extension stems from its definition that involves

    the exponentially shrinking area in the denominator. This pseudo strain hardening has

    little to do with the entanglement strain hardening due to the finite chain extensibility that

    produces non-Gaussian stretching of the entanglement network.

    Acknowledgements The authors would like to express their sincere gratitude to Dr.

    Xiaorong Wang from Bridgestone-Americas Center for Research and Technology for

    providing the SBR samples in this study. This work is supported, in part, by grants

    (DMR-0821697 and CMMI-0926522) from the National Science Foundation.

  • 29

    References

    Barroso, V. C., R. J. Andrade, and J. M. Maia, “An experimental study on the criteria for

    failure of polymer melts in uniaxial extension: The test case of a polyisobutylene melt in

    different deformation regimes,” J. Rheol. 54, 605-618 (2010).

    Boukany, P. E., S. Q. Wang, and X. R. Wang, “Universal scaling behavior in startup shear

    of entangled linear polymer melts,” J. Rheol. 53, 617-629 (2009a).

    Boukany, P. E., S. Q. Wang, and X. R. Wang, “Step shear of entangled linear polymer

    melts: new experimental evidence of elastic yielding,” Macromolecules 42, 6261-6269

    (2009b).

    Cogswell, F. N., and D. R. Moore, “A comparison between simple shear, elongation and

    equal biaxial extension deformation,” Polym. Eng. Sci. 14, 573-576 (1974).

    Considère, M., “Memoire sur l'emploi du fer et de l'acier dans les constructions,” Ann.

    Ponts. Chausèes 9, 574-605 (1885).

    Doi, M., and S. F. Edwards, “Dynamics of concentrated polymer systems, Part

    4Rheological properties,” J. Chem. Soc. Faraday Trans. 2 75, 38-54 (1979).

    Doi, M., and S. F. Edwards, The Theory of Polymer Dynamics, 2nd ed. (Clarendon,

    Oxford, 1988).

    Ferry, J. D., Viscoelastic properties of polymers, (Wiley, New York, 1980). See p. 399.

    Gleissle, W., “Two simple time-shear relations combining viscosity and first normal

    stress coefficient in the linear and non-linear flow range,” in Rheology, Vol. 2, edited by

    G. Astarita, G. Marrucci and L. Nicolais (Plenum Press, New York, 1980).

    Hassager, O., M. I. Kolte, and M. Renardy, “Failure and nonfailure of fluid filaments in

    extension,” J. Non-Newtonian Fluid Mech. 76, 137-151 (1998).

    Hassager, O., J. M. R. Marin, K. Yu, and H. K. Rasmussen, “Polymeric liquids in

    extension: fluid mechanics or rheometry?” Rheol. Acta 49, 543-554 (2010).

    Ide, Y., and J. L. White, “Investigation of failure during elongational flow of polymer

    melts,” J. Non-Newtonian Fluid Mech. 2, 281-298 (1977).

    Ide, Y., and J. L. White, “Experimental study of elongational flow and failure of polymer

    melts,” J. Appl. Polym. Sci. 22, 1061-1079 (1978).

    Joshi, Y. M., and M. M. Denn, “Rupture of entangled polymeric liquids in elongational

    flow,” J. Rheol. 47, 291-298 (2003).

    Joshi, Y. M., and M. M. Denn, “Rupture of entangled polymeric liquids in elongational

    flow with dissipation,” J. Rheol. 47, 591-598 (2004a).

    Joshi, Y. M., and M. M. Denn, “Failure and recovery of entangled polymer melts in

    elongational flow,” in Rheology Reviews (2004b), edited by D. M. Binding and K.

    Walters (British Society of Rheology, Aberystwyth, 2004), pp. 1-17.

  • 30

    Koyama, K. “Studies on elongational viscosity of polymer melts,” J. Soc. Rheol. Jpn. 19,

    174-180 (1991).

    Koyama, K., and O. Ishizuka, “Birefringence of polyethylene melt in transient

    elongational flow at constant strain rate,” J. Polym. Sci. B: Polym. Phys. 27, 297-306

    (1989).

    Laun, H. M., and H. Münstedt, “Elongational behavior of a low-density polyethylene

    melt. 1. Strain rate and stress dependence of viscosity and recoverable strain in

    steady-state. Comparison with shear data. Influence of interfacial tension,” Rheol. Acta

    17, 415-425 (1978).

    Laun, H. M., and H. Schuch, “Transient elongational viscosities and drawability of

    polymer melts,” J. Rheol. 33, 119-175 (1989). See Fig. 18(a)-(b).

    Lyhne, A., H. K. Rasmussen, and O. Hassager, “Simulation of elastic rupture in extension

    of entangled monodisperse polymer melts,” Phys. Rev. Lett. 102, 138301 (2009).

    Malkin, A. Y., and G. V. Vinogradov, “Fracture of polymers in the visco-fluid state on

    stretching,” Polym. Sci. USSR 27, 245-257 (1985).

    Malkin, A. Y., and C. J. S. Petrie, “Some conditions for rupture of polymer liquids in

    extension,” J. Rheol. 41, 1-25 (1997).

    McKinley, G. H., and O. Hassager, “The Considère condition and rapid stretching of

    linear and branched polymer melts,” J. Rheol. 43, 1195-1212 (1999).

    McLeish, T. C. B., and R. G. Larson, “Molecular constitutive equations for a class of

    branched polymers: the pom-pom model,” J. Rheol. 42, 81-110 (1998).

    McLeish, T. C. B., “A tangled table of topological fluids,”, Phys. Today, p. 40, Aug. issue

    (2008).

    Meissner, J., “Dehnungsverhalten von Polyäthylen-Schmelzen,” Rheol. Acta 10, 230-242

    (1971).

    Meissner, J., "Basic parameters, melt rheology, processing and end-use properties of three

    similar low density polyethylene samples," Pure Appl. Chem. 42, 553-612 (1975).

    Minegishi, A., A. Nishioka, T. Takahashi, Y. Masubuchi, J. Takimoto, and K. Koyama,

    “Uniaxial elongational viscosity of PS/a small amount of UHMW-PS blends,” Rheol.

    Acta 40, 329-338 (2001).

    Münstedt, H., “Dependence of the elongational behavior of polystyrene melts on

    molecular weight and molecular weight distribution,” J. Rheol. 24, 847-867 (1980).

    Münstedt, H., and S. Kurzbeck, "Elongational behaviour and molecular structure of

    polymer melts," Prog. Trends Rheol. (Proc. Fifth Eur. Rheol. Conf.) 5, 41-44 (1998).

    Nielsen, J. K., H. K. Rasmussen, O. Hassager, and G. H. McKinley, “ Elongational

    viscosity of monodisperse and bidisperse polystyrene melts,” J. Rheol. 50, 453-476

    (2006).

  • 31

    Okamoto, M., A. Kojima, and T. Kotaka, “Elongational flow and birefringence of low

    density polyethylene and its blends with ultrahigh molecular weight polyethylene,”

    Polymer 39, 2149-2153 (1998).

    Pearson, G. H., and R. W. Connelly, “The use of extensional rheometry to establish

    operating parameters for stretching processes,” J. Appl. Polym. Sci. 27, 969-981 (1982).

    Petrie, C. J. S., “Extensional viscosity: A critical discussion,” J. Non-Newtonian Fluid

    Mech. 137, 15–23 (2006).

    Petrie, C. J. S., in the reviewer report for Wang and Wang (2008).

    Petrie, C. J. S., “Considère reconsidered: Necking of polymeric liquids,” Chem. Eng. Sci.

    64, 4693-4700 (2009).

    Ravindranath, S., and S. Q. Wang, “What are the origins of stress relaxation behaviors in

    step shear of entangled polymer solutions,” Macromolecules 40, 8031-8039 (2007).

    Sentmanat, M. L., “Miniature universal testing platform: from extensional melt rheology

    to solid-state deformation behavior,” Rheol. Acta 43, 657-669 (2004).

    Sentmanat, M. L., B. N. Wang, and G. H. McKinley, “Measuring the transient extensional

    rheology of polyethylene melts using the SER universal testing platform,” J. Rheol. 49,

    585-606 (2005).

    Smith, K. J., A. Green, and A. Ciferri, "Crystalllization under stress and non-Gaussian

    behavior of macromolecular networks," Colloid. Polym. Sci. 194, 49-67 (1964).

    Takaki, T., and D. C. Bogue, “The extensional and failure properties of polymer melts,” J.

    Appl. Polym. Sci. 19, 419-433 (1975).

    Tapadia, P., and S. Q. Wang, “Direct visualization of continuous simple shear in

    Non-Newtonian polymeric fluids,” Phys. Rev. Lett. 96, 016001 (2006).

    Treloar, L. R. G., “Stress-strain data for vulcanised rubber under various types of

    deformation,” Trans. Faraday Soc. 40, 59-70 (1944).

    van Ruymbeke E., E. B. Muliawan, S. G. Hatzikiriakos, T. Watanabe, A. Hirao, and D.

    Vlassopoulos, “Viscoelasticity and extensional rheology of model Cayley-tree polymers

    of different generations,” J. Rheol. 54, 643-662 (2010).

    Vincent, P. I., “The necking and cold-drawing of rigid plastics,” Polymer 1, 7-19 (1960).

    Vinogradov, G. V., “Viscoelasticity and fracture phenomenon in uniaxial extension of

    high-molecular linear polymers,” Rheol. Acta 14, 942-954 (1975).

    Vinogradov, G. V., A. Y. Malkin, and V. V. Volesevitch, “Some fundamental problems in

    viscoelastic behavior of polymers in shear and extension,” Appl. Polym. Symp. 27, 47-59

    (1975a).

    Vinogradov, G. V., A. Y. Malkin, V. V. Volesevitch, V. P. Shatalov, and V. P. Yudin, “Flow,

    high-elastic (recoverable) deformations and rupture of uncured high molecular weight

    linear polymers in uniaxial extension,” J. Polym. Sci.: Polym. Phys. 13, 1721-1735

    (1975b).

  • 32

    Vinogradov, G. V., “Ultimate regimes of deformation of linear flexible chain fluid

    polymers,” Polymer 18, 1275-1285 (1977).

    Vinogradov, G. V., and A. Y. Malkin, Rheology of Polymers (Mir, Moscow, 1980)

    Wagner, M. H., S. Kheirandish, K. Koyama, A. Nishioka, A. Minegishi, and T. Takahashi,

    “Modeling strain hardening of polydisperse polystyrene melts by molecular stress

    function theory,” Rheol. Acta 44, 235-243 (2005).

    Wang, S. F., S. Q. Wang, A. Halasa, and W. L. Hsu, “Relaxation dynamics in mixtures of

    long and short chains: tube dilation and impeded curvilinear diffusion,” Macromolecules

    36, 5355-5371 (2003).

    Wang, S. Q., “A coherent description of nonlinear flow behavior of rntangled polymers as

    related to processing and numerical simulations,” Macromol. Mater. Eng. 292, 15–22

    (2007).

    Wang, S. Q., S. Ravindranath, P. E. Boukany, M. Olechnowciz, R. P. Quirk, A. Halasa,

    and J. Mays, “Non-quiescent relaxation in entangled polymeric liquids after step shear,”

    Phys. Rev. Lett. 97, 187801 (2006).

    Wang, S. Q., S. Ravindranath, Y. Y. Wang, and P. E. Boukany, “New theoretical

    considerations in polymer rheology: elastic breakdown of entanglement network,” J.

    Chem. Phys. 127, 064903 (2007a).

    Wang, Y. Y., P. E. Boukany, S. Q. Wang, and X. R. Wang, “Elastic breakup in uniaxial

    extension of entangled polymer melts,” Phys. Rev. Lett. 99, 237801 (2007b).

    Wang, Y. Y., and S. Q. Wang, “From elastic deformation to terminal flow of a

    monodisperse entangled melt in uniaxial extension,” J. Rheol. 52, 1275-1290 (2008).

    Wang, Y. Y., and S. Q. Wang, “Exploring stress overshoot phenomenon upon startup

    deformation of entangled linear polymeric liquids,” J. Rheol. 53, 1389-1401 (2009).

    Wang, Y. Y., and S. Q. Wang, “On elastic rupture in glass-like zone of uniaxial extension

    in entangled melts,” Rheol. Acta, in press (2010a). Here we have shown for SBR melts

    with microstructure similar to the present SBR melts that the onset of non-Gaussian

    stretching corresponds to a Cauchy stress around 11 MPa.

    Wang, Y. Y., and S. Q. Wang, “Extensibility vs. entanglement spacing: rapid uniaxial

    deformation behavior of entangled melts and binary mixtures,” Manuscript under

    preparation (2010b).

    Yang, X., S. Q. Wang, and H. Ishida, “A solution approach to component dynamics of

    A/B miscible blends. 1. Tube dilation, reptation, and segmental friction of polymer A,”

    Macromolecules 32, 2638-2645 (1999).

    Yao, M., G. H. McKinley, and B. Debbaut, “Extensional deformation, stress relaxation

    and necking failure of viscoelatsic filaments,” J. Non-Newtonian Fluid Mech. 79,

    469-501 (1998).

    Ye, X., R. G. Larson, C. Pattamaprom, and T. Sridhar, “Extensional properties of

    monodisperse and bidisperse polystyrene solutions,” J. Rheol. 47, 443-468 (2003).