the chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf ·...

10
The chemical structure of the amorphous phase of propylenee ethylene random copolymers in relation to their stressestrain properties Vipin Agarwal a, b, 1 , Tim B. van Erp c , Luigi Balzano b, c, f , Markus Gahleitner d , Matthew Parkinson d , Leon E. Govaert c , Victor Litvinov b, e, ** , Arno P.M. Kentgens a, * a Radboud University, Institute for Molecules and Materials (IMM), Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlands b Dutch Polymer Institute (DPI), P.O. Box 902, 5600 AX Eindhoven, The Netherlands c Department of Mechanical Engineering, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands d Borealis Polyolene GmbH, Sankt-Peter-Strasse 25, 4021 Linz, Austria e DSM Resolve, P.O. Box 18, 6160 MD Geleen, The Netherlands f DSM Ahead, P.O. Box 18, 6160 MD Geleen, The Netherlands article info Article history: Received 30 October 2013 Received in revised form 21 December 2013 Accepted 25 December 2013 Available online 4 January 2014 Keywords: Propyleneeethylene random copolymers Solid-state NMR Mechanical properties abstract A better understanding of structure-property relations is necessary to design novel materials. In this study, we investigate the morphology and chemical structure of ve commercial grades of propylene- based polymers in relation to the change in yield- stress as a function of strain-rate. Substantial emphasis has been laid on understanding the chain microstructure in the relation to chain dynamics in the amorphous phase. Heterogeneous ZieglereNatta catalysis was used to prepare the samples with differing ratios of propylene and ethylene units. Various analytical techniques such as WAXS, SAXS, solution- and solid-state NMR were employed to characterize their structure. The results indicate a reduction in crystallinity, melting temperature, long-period and crystal thickness with increasing ethylene content. Solid-state NMR data reveal the presence of four components in these samples, which is an extension of the traditional three phase model found in most semi-crystalline polymers. The additional fourth phase is attributed to a rubber-like component that is primarily composed of chain segments rich in ethylene units and shows an increase in chain dynamics with increasing ethylene content in the samples. Mechanical experiments show that yield stress decreases with increase in the amount ethylene which can be correlated to the observed increase in chain dynamics in the amorphous phase. Ó 2014 Elsevier Ltd. All rights reserved. 1. Introduction The physical and mechanical properties of a polymer are a complex interplay of morphology and the chemical structure of the polymer. While the chemical structures are dened during syn- thesis, molecular conformation and crystal morphology distribu- tions are shaped during post-synthesis processing. The dependence of the conformation and morphology on processing conditions arises because the physical structure formation in polymers such as polyethylene and polypropylene, is often determined by crystalli- zation kinetics [1]. Depending on the processing conditions one can inuence the relative amount of various phases, the type of crystals formed, orientation, etc. [2e12]. These variations in the structure dene the resulting properties of the material. Polypropylene (PP) is an important polymer which is often encountered in our daily life because it is an economical material with a unique combination of desirable physical, mechanical, thermal, chemical and electrical properties [13e19]. Commercially, PP is available in the form of a homopolymer, random copolymers, block copolymers and rubber modied impact copolymer. Random copolymers with small fraction of a-olen (e.g., ethylene, butene, hexane, etc) units (0.2e7 mol%) were developed to improve prop- erties of PP-homopolymers such as impact strength at low- temperature, clarity, stress-crack resistance and sealing tempera- tures as well as other properties [20e25]. * Corresponding author. ** Corresponding author. DSM Resolve, P.O. Box 18, 6160 MD Geleen, The Netherlands. E-mail addresses: [email protected] (V. Litvinov), [email protected] (A.P.M. Kentgens). 1 Current address: Physical Chemistry, ETH Zurich, 8093 Zurich, Switzerland. Contents lists available at ScienceDirect Polymer journal homepage: www.elsevier.com/locate/polymer 0032-3861/$ e see front matter Ó 2014 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.polymer.2013.12.051 Polymer 55 (2014) 896e905

Upload: others

Post on 11-Jul-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

lable at ScienceDirect

Polymer 55 (2014) 896e905

Contents lists avai

Polymer

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

The chemical structure of the amorphous phase of propyleneeethylene random copolymers in relation to their stressestrainproperties

Vipin Agarwal a,b,1, Tim B. van Erp c, Luigi Balzano b,c, f, Markus Gahleitner d,Matthew Parkinson d, Leon E. Govaert c, Victor Litvinov b,e,**, Arno P.M. Kentgens a,*

aRadboud University, Institute for Molecules and Materials (IMM), Heyendaalseweg 135, 6525 AJ Nijmegen, The NetherlandsbDutch Polymer Institute (DPI), P.O. Box 902, 5600 AX Eindhoven, The NetherlandscDepartment of Mechanical Engineering, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The NetherlandsdBorealis Polyolefine GmbH, Sankt-Peter-Strasse 25, 4021 Linz, AustriaeDSM Resolve, P.O. Box 18, 6160 MD Geleen, The NetherlandsfDSM Ahead, P.O. Box 18, 6160 MD Geleen, The Netherlands

a r t i c l e i n f o

Article history:Received 30 October 2013Received in revised form21 December 2013Accepted 25 December 2013Available online 4 January 2014

Keywords:Propyleneeethylene random copolymersSolid-state NMRMechanical properties

* Corresponding author.** Corresponding author. DSM Resolve, P.O. BoxNetherlands.

E-mail addresses: [email protected] (V. Lit(A.P.M. Kentgens).

1 Current address: Physical Chemistry, ETH Zurich,

0032-3861/$ e see front matter � 2014 Elsevier Ltd.http://dx.doi.org/10.1016/j.polymer.2013.12.051

a b s t r a c t

A better understanding of structure-property relations is necessary to design novel materials. In thisstudy, we investigate the morphology and chemical structure of five commercial grades of propylene-based polymers in relation to the change in yield- stress as a function of strain-rate. Substantialemphasis has been laid on understanding the chain microstructure in the relation to chain dynamics inthe amorphous phase. Heterogeneous ZieglereNatta catalysis was used to prepare the samples withdiffering ratios of propylene and ethylene units. Various analytical techniques such as WAXS, SAXS,solution- and solid-state NMR were employed to characterize their structure. The results indicate areduction in crystallinity, melting temperature, long-period and crystal thickness with increasingethylene content. Solid-state NMR data reveal the presence of four components in these samples, whichis an extension of the traditional three phase model found in most semi-crystalline polymers. Theadditional fourth phase is attributed to a rubber-like component that is primarily composed of chainsegments rich in ethylene units and shows an increase in chain dynamics with increasing ethylenecontent in the samples. Mechanical experiments show that yield stress decreases with increase in theamount ethylene which can be correlated to the observed increase in chain dynamics in the amorphousphase.

� 2014 Elsevier Ltd. All rights reserved.

1. Introduction

The physical and mechanical properties of a polymer are acomplex interplay of morphology and the chemical structure of thepolymer. While the chemical structures are defined during syn-thesis, molecular conformation and crystal morphology distribu-tions are shaped during post-synthesis processing. The dependenceof the conformation and morphology on processing conditionsarises because the physical structure formation in polymers such as

18, 6160 MD Geleen, The

vinov), [email protected]

8093 Zurich, Switzerland.

All rights reserved.

polyethylene and polypropylene, is often determined by crystalli-zation kinetics [1]. Depending on the processing conditions one caninfluence the relative amount of various phases, the type of crystalsformed, orientation, etc. [2e12]. These variations in the structuredefine the resulting properties of the material.

Polypropylene (PP) is an important polymer which is oftenencountered in our daily life because it is an economical materialwith a unique combination of desirable physical, mechanical,thermal, chemical and electrical properties [13e19]. Commercially,PP is available in the form of a homopolymer, random copolymers,block copolymers and rubber modified impact copolymer. Randomcopolymers with small fraction of a-olefin (e.g., ethylene, butene,hexane, etc) units (0.2e7 mol%) were developed to improve prop-erties of PP-homopolymers such as impact strength at low-temperature, clarity, stress-crack resistance and sealing tempera-tures as well as other properties [20e25].

Page 2: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

Table 1Sample characteristics: Ethylene content (E), molecular weight characteristics (Mw

and Mw/Mn), the amount of rubber content expressed as the xylene cold solublefraction (XS) e (25 �C ISO 16152), melt flow rate (MFR) e 230 �C/2.16 kg ISO 1133,and flexural modulus (flex).

Sample E, mole % Mw, kg/mol Mw/Mn XS, wt% MFR, g/10 min Flex.,MPa

iPP 0.0 360 5.0 1.2 8 1600PPR4 4.1 w310 w3.4 3.3 8 1000PPR6 5.9 w310 w3.4 4.7 8 800PPR7 7.1 w310 w3.4 6.8 8 605IEPC23 23.4 365 22.0 6 840

V. Agarwal et al. / Polymer 55 (2014) 896e905 897

A number of methods have been employed to understand thestructure of semi-crystalline polymers. On a molecular level, mostexperimental techniques differ with respect to their sensitivity indetecting various phases in the polymer. In order to understandstructureeproperty relationships it is necessary to elucidate thechemical and physical structures of each phase in terms of phasecomposition, domain sizes, microstructure, conformation andchain dynamics. NMR spectroscopy has become an indispensabletool for the characterization of polymers. Typically high-resolutionsolution-state NMR spectroscopy allows identification and quan-tification of the various microstructures formed during polymeri-zation. For example, the comonomer content, degree of branching,comonomer sequence distribution, tacticity, content of regio-defects and amount of end groups in PP copolymers can all bedetermined from a single quantitative 13C solution-state NMRspectrum [26e29]. In contrast, solid-state NMR probes polymers intheir native state and under conditions that are identical to theirapplication environment. It provides a snapshot of the ensemble ofall chains in different phases. Proton NMR relaxometry combinedwith high-resolution magic-angle spinning (MAS) NMR can beexploited to probe physical phase structures, domain sizes, chainconformation and chain dynamics of the final processed material indifferent phases.

Due to the commercial importance of propyleneeethylenerandom copolymers (PPR) numerous studies of these polymershave been undertaken in order to characterize their chemicalstructure, morphology, mechanical, physical and thermal proper-ties [24,25,30,31]. Most of these studies have focused on PPR’ssynthesized with metallocene-based catalysts [21,25,30]. This ismainly caused by the practical difficulties of characterizing poly-mers synthesized via different routes leading to increasingcompositional and physical structure heterogeneity, requiringmore complex analysis.

The aim of this work is to gain in-depth insight into the chemicaland physical structures of polypropylene-based polymers producedusing heterogeneous ZieglereNatta (ZN) catalysis. To be morespecific, we investigate the chemical and physical composition ofan isotactic polypropylene homopolymer (iPP), three propyleneeethylene random copolymers (PPR) and a heterophasic propyleneeethylene copolymer (IEPC). Polymer microstructure was quantifiedusing solution-state 13C NMR spectroscopy. The effect of theamount of comonomer on phase composition and molecularmobility was studied over a wide temperature range. The phasecomposition was determined using low-field 1H NMR relaxometryand wide-angle X-ray scattering (WAXS), while small angle X-rayscattering (SAXS) was used to obtain domain sizes. High-resolution1H and 13C solid-state NMR spectroscopy were used to identifydifferent chain microstructures present in the amorphous phase.Finally, the strain-rate dependent yield stress of different sampleswas measured and is shown to be associated with variations inmolecular dynamics in the amorphous phase.

2. Experimental

2.1. Sample characteristics

All polymers were commercial grades produced by Borealis AGvia multi-stage liquid bulk and gas phase polymerization processusing a fourth-generation ZieglereNatta type catalyst with a con-ventional silane external donor. No processing step was performedto select polymer chains with a particular constitution or micro-structure. The polypropylene homopolymer, hereafter iPP, was areactor grade with a pentad isotacticity of 99.3%. The three pro-pyleneeethylene random copolymers, hereafter referred to asPPR4, PPR6 and PPR7 were all produced via a visbreaking process

and had ethylene contents of 4.1, 5.9 and 7.1 mol%, respectively, asdetermined by liquid-state 13C NMR spectroscopy. The hetero-phasic propyleneeethylene copolymer, hereafter called IEPC23,had a total ethylene content of 23.4 mol% as determined by NMR.The xylene cold soluble fraction in IEPC23 had a Mw of 492 kg/mol.Sample characteristics are provided in Table 1.

2.2. Experimental details

A detailed description of the experimental conditions for solu-tion and solid-state NMR experiments, SAXS and WAXS experi-ments and the mechanical testing are provided in the Supportinginformation.

3. Results and discussions

3.1. Analysis of polymer microstructure by solution-state 13C NMR

For the polypropylene homopolymer no indication of ethylenechain units is observed. Although this statement might appearredundant, industrially, it is common to use ethylene to reactivateZN catalysts resulting in so-called mini-random copolymers withvery low ethylene contents of between 0.1 and 0.3 mol%. Similarly,no indications of regio-irregularity are observed, consistent for ZNcatalysis. With no comonomer or regio-defects present, the mainmicrostructure originates from tacticity and end groups. Thepentad tacticity distribution of 99.3% is determined from the Pbbsignals in the methyl region [32]. This isotacticity is typical for iPPproduced by ZN catalysis. From inspection of the methyl region ofthe NMR spectra, it appears that the distribution of stereo se-quences remains approximately the same for all samples takinginto account the multiple relatively intense signals arising fromethylene incorporation. As far as the end-groups is concerned, onlyn-propyl and iso-butyl chain-ends are observed in significant andapproximately equal quantities suggesting standard chain initia-tion and termination mechanisms common for propylene homo-polymerization. With respect to crystallizable sequences theaverage meso sequence length (MSL) was found to be 126 propeneunits using the relationship, MSL ¼ 5 þ 2 [mmmm]/[mmmr].

For the propyleneeethylene random copolymers PPR4, PPR6,PPR7 and IEPC23 characteristic signals corresponding to theincorporation of ethylene are observed [33]. The ethylene content isquantified based on themethod ofWang et al. [34] by integration ofmultiple signals across the whole range of the 13C [1H] decoupledNMR spectra (Table 2). The triad comonomer sequence distribu-tions are determined using the method of Kakugo et al. [35]. Fromthe normalized triad concentrations, the ethylene content can becalculated as the sum of the ethylene centered triads (XEX) [27].Due to the different integrals used and difference sources ofsystematic error, comonomer contents determined by the methodof Wang and Kakugo are commonly not comparable atlow comonomer contents. The content of isolated ethylene units

Page 3: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

Table 2Chainmicrostructure in polypropylene samples: themole andweight fractions of ethylene units (E), triad distribution, content of isolated ethylene units and the average lengthof propylene sequences (nPP).

Sample [E] [PPP] mol % [PPE] mol % [EPE] mol % [PEP] mol % [PEE] mol % [EEE] mol % [XEX] mol % I % nP

mol % wt %

PPR4 4.1 2.8 90.6 6.1 0.0 2.4 0.5 0.4 3.3 73 32PPR6 5.9 4.0 86.0 7.9 0.2 3.9 1.4 0.6 5.8 67 22PPR7 7.1 4.8 83.3 9.1 0.5 4.5 1.8 0.7 7.0 65 18IEPC23 23.4 17.0 67.9 5.5 3.6 2.9 7.7 12.4 23.0 13 12

V. Agarwal et al. / Polymer 55 (2014) 896e905898

(I e “randomness”) is quantified as the relative amount of isolatedethylene sequences (EPE) compared to all ethylene centered triads(EEE, EEP, PEP). The average propene sequence length, nP, iscalculated from the triads according to the relationship, nP ¼ 1 þ 2[EE]/[EP], where the dyads [EE] and [EP] are calculated from thetriads using the known relationships. From the triad sequencedistribution and the degree of isolated ethylene (I) minor butimportant differences can be seen between the PPR copolymersand clear differences between the PPR copolymers and the heter-ophasic IEPC23 sample (Table 2). The relative content of the PEPsequence, which represents isolated ethylene incorporation, showsclear differences in the series of PPR copolymers. With increasingethylene content a higher amount of isolated ethylene units areobserved. However, comparing to the other ethylene centered tri-ads the relative ratio of PEP is similar for PPR6 and PPR7 suggestingthe same polymerization statistics. As expected, IEPC23 shows ahigher relative amount of PEE and EEE sequences, which is anindication of a higher probability of ethylene block formation. Inaccordance with this observation, a much lower amount of isolated(I) ethylene incorporation (PEP) is observed.

3.2. Analysis of polymer morphology by in-situ X-ray experiments

WAXS measurements for iPP, PPR4 and PPR7 reveal that, duringheating at 5 �C/min, crystallinity decreases gradually with tem-perature until the onset of melting, where the decrease becomesrather sharp (Fig. 1a). The presence of ethylene as a comonomer

Fig. 1. a) WAXD crystallinity as a function of temperature during sample heating at 5 �C/minrate of crystallinity change. b) Melting temperature versus ethylene content as depicted bfunction of temperature during sample heating at 5 �C/min.

clearly affects both the initial crystallinity and the melting behaviorof the materials. As expected, samples containing ethylene units(PPR4 and PPR7) exhibit lower crystallinity and melting tempera-ture (Tm) than the homopolymer (iPP). These effects can beessentially ascribed to shorter length of crystallizable propylenesequences (Table 2) [36]. Under the hypothesis that the ethyleneco-units are rejected from the crystalline lattice, the melting pointdepression (with respect to the equilibrium melting temperature,T0m) can be predicted by the theory of Flory asð1=TmÞ � ð1=T0

mÞ ¼ ð�R=DHuÞln p; where DHu is the enthalpy offusion per repeating unit, and p the probability that a randomlyselected crystallizable unit is followed by another crystallizable unit(for a truly statistical copolymer, p is simply the mole fraction ofcrystallizable units) [37,38]. However, Flory’s theory is not expectedto quantitatively describe the melting point depression observed inour experiments because (1) the experimental conditions are farfrom equilibrium, (2) the co-monomer distribution is not trulyrandom, and (3) the co-monomer may partially be included in thecrystalline lattice. When Tm is defined as the temperature at whichthe melting rate is the highest, the measured melting pointdepression can be estimated to be about 2 �C per mole percent ofethylene (Fig. 1b).

The thermal stability of crystals depends on their thickness andthe amount of lattice defects, including ethylene incorporation. Thehomopolymer (iPP), which has the longest crystallizable propylenesequences, exhibits the highest crystal thickness of w7 nm(Lc ¼ Lp$x). As shown in Fig. 1c, the average long period Lp increases

. The continuous lines are visual guides. The stars represent the points with the fastesty stars in Fig. 1a. c) Long-period determined from SAXS, and d) crystal thickness as a

Page 4: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

V. Agarwal et al. / Polymer 55 (2014) 896e905 899

upon heating, due to melting and re-crystallization processes thatespecially affect the smallest and most defective crystals and/orparts of them [39]. In this process, some of the density fluctuationswithin the crystals at the shorter length scales disappear. At thesame time, the average crystal thickness remains essentially unal-tered up to about 100 �C, as observed in Fig. 1d. At higher tem-perature, before the onset of melting, crystals tend to thicken. Thiseffect is especially pronounced in the homopolymer just prior tomelting. The thickness of remaining crystals increases up to 8 nm(Fig. 1d). Crystal thickening is clearly less pronounced in the co-polymers, whichmight be due to the higher ethylene content in theinterface regions.

3.3. Analysis of polymer morphology by solid-state NMR

3.3.1. 1H NMR spin-lattice relaxationThe proton spin-lattice relaxation time constants (T1) for the five

samples are measured as a function of temperature (Fig. 2). The T1of all the samples decreases as a function of increasing ethylenecontent, except for IEPC23. A mono-exponential function is suffi-cient to fit the experimental T1 relaxation data for the homopoly-mer and the PPR samples at all temperatures, whereas a bi-exponential function is necessary to fit the T1 data of the IEPC23sample.

For iPP, two main relaxation processes, namely methyl grouprotation and main-chain motions dominate the 1H T1 relaxationbehavior at different temperatures. Two T1 minima, one at �110 �Cand the other one at 90 �C, were observed previously in the tem-perature range from �200 �C to 160 �C [40]. The 1H T1 minimumat�110 �Cwas attributed to C3-rotation of themethyl group both inthe crystalline and amorphous phases, while the minimum at 90 �Cwas assigned to a-relaxation process in the amorphous phase,namely dynamic glass transition at a frequency of tens of MHz. InFig. 2 we observe that around 80 �C, when the a-process becomesthe dominant source of the T1 relaxation, the variation in 1H T1 isthe smallest between different PPR samples. At w0 �C, when noneof the relaxation process dominate, maximum differences are seenbetween the 1H T1’s of the PPR copolymers. Upon approaching thea-process minimum, the short and long 1H T1 components ofIEPC23 can no longer be distinguished. The bi-exponential natureof the T1 data for the impact polymer results from phase separationof propylene-rich and ethylene-rich chains and/or chain fragments.

Fig. 2. Plot of 1H T1 for different samples as a function of temperature. The legend boxin the Figure indicates the sample designation. The IEPC23 sample has a bi-exponentialT1 relaxation in the temperature range between the two minima.

The major fraction of IEPC23 has exactly the same T1 as that of theiPP while the minor fraction shows a very short T1, indicative of fastsegmental dynamics.

3.3.2. Phase composition as determined by 1H NMR T2 relaxationProton NMR is a sensitive and robust method for determining

phase composition and molecular mobility in complex multiphasepolymers [40e42]. In the case of multiphase systems such assemicrystalline polymers, chain motion in the crystalline- andrigid/glassy amorphous phases is restricted. Therefore, the T2 ofthese phases is dominated by strong homonuclear protoneprotondipolar couplings and is on the order of 8e20 ms. 1H T2 can varyfrom hundreds of microseconds to a few milliseconds for rubberymaterials and to tens of milliseconds for highly viscous liquids.Besides predicting the number of phases/components, the methodis inherently quantitative in nature since the amplitude of therelaxation components of each phase is directly proportional to thenumber of protons present in that phase. The difference in T2relaxation time of various components increases as a function oftemperature and is largest just before the melting of the sample.

The exact theoretical description of the T2 decay rate and itsshape is difficult to predict due to the complex nature of therelaxation and morphological heterogeneities of multi-phasepolymers. However, a phenomenological model comprising alinear combination of functions appears to reliably define theexperimental data. In this study a linear combination of an Abra-gamian, Gaussian and two exponential functions (Eq. (1)), is foundto uniquely fit the experimental data for all samples over thetemperature range from 50 �C to 90 �C. These four relaxationcomponents are identified based on a distinct difference in mo-lecular mobility and are designated in the order of increasing T2value as rigid, semi-rigid, soft-amorphous and mobile-amorphousfractions:

AðtÞ ¼ARigid0 exp

24�

�t

Trigid2

!235,�Sin ðatÞ

at

þ ASemi�Rigid0 exp

24�

�t

TSemi�rigid2

!235

þ ASoft�amorphous0 exp

�t

TSoft�Am2

!

þ AMobile�amorphous0 exp

�t

Tmob�Am2

!

[1]

where the parameter TX2 (where X identifies the phase being rigid,

semi-rigid, soft-amorphous and mobile-amorphous) is themobility factor for different components, AX

0 is the amplitude of the Xcomponent, and parameter ‘a’ is related to second and fourthvan Vleck moments for the rigid fraction. The relativefraction of each component of this relaxation model e

AX0*100%=fARigid

0 þ ASemi�Rigid0 þ ASoft�Amorphous

0 þ AMobile�Amorphous0 g

eprovides thephase composition (inwt%)of the studiedpolymers. Itshould be noted that the terms soft-amorphous and mobile-amorphous are chosen somewhat arbitrarily to underline the dif-ference in mobility in the mobile fractions of the material one ofwhich is highly mobile. Fig. 3 shows an illustrative example of theleast square fit (using Eq. (1)), of the 1H T2 experimental data (B) andcorresponding residuals for the PPR7 sample at 70 �C. Although thegeneral quality of the fit is good, as judged by the residuals, somedeviations are observed at longer decay time (>1 ms). This behavioris caused by large amplitude motions in the mobile amorphousphase as is typical for rubbers. It is shown below that the

Page 5: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

Rigid Component

Semi-rigidComponent

Soft-AmorphousComponent

Mobile-AmorphousComponent

a) b)

c) d)

0 500 1000 1500 2000 0 500 1000 1500 2000

0 100 200 300 400 0 100 200 300 4001.0

0.1

0.01

1.0

0.1

0.01

0.010

0.006

0.002

-0.002

-0.010

-0.006

0

0.010

0.006

0.002

-0.002

-0.010

-0.006

0

Fig. 3. Proton T2 decay of PPR7 measured with one pulse and Hahn-echo pulse sequences (B) at 70 �C. Two time windows are shown (a) 400 ms and (c) 2000 ms with correspondingresiduals. The intensity axis is plotted on a log10 scale. The solid red curve represents the result of least-square fit of the decay using Eq. (1). The dotted curves represent the decay ofdifferent components that are assigned to rigid, semi-rigid, soft-amorphous and mobile amorphous fractions, components 1e4, respectively. Residuals (b, d) for both time windowsshow low (0.2%) deviation for the initial part of the decay but increased deviation at longer times. (For interpretation of the references to colour in this figure legend, the reader isreferred to the web version of this article.)

V. Agarwal et al. / Polymer 55 (2014) 896e905900

mobile-amorphousphasedisplays rubber-like behavior. The origin ofthe T2 relaxation in rubbers is complex and, as reported in the liter-ature [43,44], cannot be described by a single exponential function.

Several 1H T2 studies of iPP and PPR have been reported inliterature [45e52]. All these studies used different relaxationmodels, comprising two to five components represented by func-tions such as Gaussian’s, Weibull, exponential and polynomial. Themorphology of semicrystalline polymers is very complex and it istherefore difficult to have a unique model to define relaxationprofiles of samples synthesized and processed under differentconditions. All relaxation models for deconvolution and fitting of1H T2 data are empirical in nature. So, the best approach is to adopta relaxation model with a minimum number of components. Theaddition of extra components in the data fitting must be experi-mentally justified.

3.3.3. Phase composition and mobility as a function of temperatureA minimum of four components is necessary to fit the T2 data

above ambient temperatures. On lowering the temperature from90 �C to �20 �C, the minimum number of components required forfitting is sequentially reduced from four to three components, andbelow room temperature to two components. This is related to thetimescale of the motions in each phase as a function of temperaturein relation to the timescale of the NMR experiments. For furtherinformation concerning the fitting procedure and the adaptation ofthe relaxation model as a function of temperature the reader isdirected to the Supporting Information.

Fig. 4AeD shows the variation in relative concentration of therigid, semi-rigid, soft and mobile amorphous components whileFig. 4EeF show T2 constants for the soft and mobile amorphouscomponents of the five samples as a function of temperature. TheTRigid2 and TSemirigid

2 are not explicitly reported because these tworelaxation components were fitted with Abragamian and Gaussianfunctions and, therefore, the time constant in both these functionsis not representative for the transverse relaxation constant

associated with exponential functions. In addition, the variations in“effective” TRigid

2 (17e19 ms) and TSemirigid2 (23e26 ms) components

are minor over a temperature interval of �60 �C to 90 �C.From Fig. 4, several features are observed. For all materials, the

amount of rigid fraction decreases with increasing temperature.The rate of decrease is more prominent at temperatures slightlyabove the glass temperature (Tg) (between 0 �C and 30 �C) thanbetween 40 �C and 90 �C. Below Tg of the amorphous phase, themobility of chains in the amorphous phases is frozen on the NMRtimescale, and therefore adds to the amount of the rigid fractionwhich appears to be very similar at low temperature for all samplesexcept the phase segregated IEPC23 (Fig. 4A). A similar behavior isobserved for the amount of semi-rigid component in all samplesexcept in the temperature range from 0 to 20 �C, at which theamount of the semi-rigid fraction passes through a maximum(Fig. 4B). The initial increase of the semi-rigid fraction on loweringthe temperature is due to decreased motion in the amorphouscomponents, with the two components of the amorphous phasebecoming indistinguishable from the semi-rigid fraction at themaxima. This is supported by the fact that TSoft�Am

2 and TMob�Am2

decrease on lowering the temperature and become approximatelyequal to TSemirigid

2 around 0e20 �C (Fig. 4EeF). With a furtherdecrease in temperature the amount of the semi-rigid fractiondecreases. Such a decrease is most likely due to a further reductionof mobility making parts of the semi-rigid fraction indistinguish-able from the rigid component. The aberrant behavior of the semi-rigid fraction for the IEPC23 results from fitting issues and thereader is referred to the Supporting Information.

The soft-amorphous fraction is initially detectedbetween �10 �C and 10 �C for different samples and steadily growsas a function of temperature (Fig. 4C). The temperature range, inwhich the soft amorphous fraction appears, is consistent with theTg value of the amorphous phase obtained fromDMTA analysis [53].The concentration of the soft-amorphous fraction increases onheating and becomes almost constant above 60 �C. The TSoft�Am

2

Page 6: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

Fig. 4. Change in phase composition and mobility factor (T2) of different components for the five samples as a function of temperature. A), B), C) and D) represent the dependence ofthe amount of the rigid, semi-rigid, soft and mobile amorphous components, while E) and F) represent change in the T2 of the soft and mobile amorphous components in thetemperature range from �20 �C and 90 �C.

V. Agarwal et al. / Polymer 55 (2014) 896e905 901

steadily increases over the entire temperature range but in the lowtemperature range TSoft�Am

2 shows deviations from expectedbehavior (Fig. 4E) because the relative concentration of the soft-amorphous fraction is very small, less than 2% and this introducesproblems in accurately fittingTSoft�Am

2 relaxation component. Themobile-amorphous fraction is initially observed between 20 �C and50 �C for iPP and PPR samples and at �20 �C for the IEPC23 sample(Fig. 4D,F). It is interesting to note that the higher the ethylenecontent, the lower the temperature where the mobile-amorphousphase appears. At the onset temperature for detecting themobile-amorphous component, TMob�Am

2 is in the range of 120e150 ms and this value steadily increases to several hundreds of msecfor iPP and the PPR samples and up to 1.5 ms for IEPC23. (Fig. 4F). T2values on the order of 0.5e1.5 ms correspond to NMR line width ofless than 1 kHz, indicating fast close to isotropic motions.

3.3.4. Evidence for the presence of a rubber-like componentStatic proton spectra acquired with a T2 filter of 16 ms selects

only the mobile-amorphous component. The T2 filtered 1H spectrayields a static proton line width (full width at the half height) ofapproximately 525 Hz for PPR7 (Fig. 5a). In contrast to the wide-line 1H NMR spectrum, the line width of the T2 filtered static pro-ton spectrum does not change upon increasing the temperaturefrom 20 �C to 80 �C. This implies that due to fast molecular motionsdipolar couplings are averaged out and the observed proton linewidth results largely from dispersion of chemical shifts due tochemical structure heterogeneity, and possibly from residual

chemical shift anisotropy or dipolar coupling. Furthermore, theintensity of this component increases with increase in temperature.An array of T2 filtered 1H NMR spectra (Fig. 5a) recorded at differentMAS spinning frequency show that a modest spinning frequency of400 Hz is sufficient to resolve three different isotropic proton res-onances of PPR7. The same spectrum recorded at 21.1T staticmagnetic field reveals at least five unique proton resonances(Fig. 5b). Fig. 5c shows a 2D 1He13C J-resolved spectrum for thePPR7 sample in the absence of any homonuclear decoupling. The1JHeC values observed in the indirect dimension are the unscaledscalar couplings for the CHn groups. The doublet from the eCHmoiety shows the expected 1:1 intensity in the F1 dimension. Theintensities of the eCH2 and eCH3 multiplets slightly differ fromtheir expected values of 1:2:1 and 1:3:3:1 in an isotropic medium.

J (scalar) couplings are weak interactions observed in NMR dueto the interaction of two nuclei mediated by bonding electrons. Thisinteraction is commonly observed in NMR spectra in solutions,where molecules undergo an isotropic tumbling motion. In NMRspectra of polymers J-couplings can be observed when the mole-cules undergo rapid nearly isotropic motions as is the case inrubbers, e.g., natural rubber [54]. In rigid solids this interaction canonly be observed under multi-pulse decoupling conditions. Evenfor adamantane (a globular molecule with 17 ps correlation time ofmolecular tumbling) it is not possible to observe 1JHeC couplings atmodest spinning frequencies [55]. High-resolution proton spectraand resolved scalar couplings at modest MAS frequency pointto large amplitude chain motions only observed in polymers

Page 7: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

20253035404550-300

-200

-100

0

100

200

300

13C Chemical Shifts (ppm)13

C-1

H J

-cou

plin

g (H

z)

-1.5 -0.9 -0.3 0.3 0.9 1.5

b)a)

c)

1H Chemical Shifts(kHz)

STATIC

MAS=400Hz

MAS=860Hz

MAS=1500Hz

MAS=2500Hz

MAS=2000Hz

-2 -1.5 -1 -0.5 0 0.5 1 1.5 2

1H Chemical Shifts(kHz)

Fig. 5. a) T2 filtered 1H NMR spectra of PPR7 as a function of MAS frequency. The spectra are acquired on a 400 MHz spectrometer at a temperature of 55 �C. The dotted line indicatesthe presence of a shoulder. b) 1H MAS NMR spectrum of PPR7 acquired at 850 MHz at a temperature of 55 �C employing a spinning speed of 5 kHz. c) 2D heteronuclear J-resolvedNMR spectrum of the mobile-amorphous component of PPR7 in the absence of any special decoupling conditions and at MAS frequency of 3.4 kHz.

V. Agarwal et al. / Polymer 55 (2014) 896e905902

exhibiting fast, near isotropic chain mobility such as, e.g., butadienechains in styrene-butadiene rubbers [56]. Thus, these independentexperiments confirm the presence of a rubber-like fraction in theamorphous phase of the PPR samples. To the best of our knowledgethe presence of a rubber-like fraction in PPR samples with very lowethylene content has never been directly observed before, althoughit has been suggested previously [52]. This raises an interestingquestion about the origin and chemical structure of this mobileamorphous fraction in PPR. For the impact copolymers, such asIEPC23, the existence of a separate ethylene-rich rubbery phase iswell-known and has been confirmed by various methods (see e.g.,TEM pictures of IEPC23 in the Supporting Information). We showfurther (Section 3.3.6) that ethylene-rich domains populate themobile-amorphous component in PPR.

3.3.5. Effect of ethylene concentration on the morphology anddynamics of the amorphous phase

The concentration and mobility (T2) of both amorphous frac-tions in PPR increase with increasing ethylene content except forIEPC23 (Fig. 6). Despite the high concentration of ethylene units inIEPC23, the fraction of the soft amorphous phase in IEPC23 is thelowest of all samples due to the good phase separation of chainsegments rich in iPP from the chains populated by ethylene units.As a function of temperature, the growth in concentration of theamorphous phase is uniform over different samples while thechange inT2, especiallyTMob�Am

2 , is more pronounced when moreethylene units are present in the sample.

These results are significant in the context of addressing theissue whether polymer chains with a high-density of ethylene/stereo defects form separate domains or are dispersed randomlyover the polymer matrix. The dependence of TMob�Am

2 on ethylenecontent (including heterophasic propyleneeethylene copolymer)can only be explained if chains with high defect concentration formseparate domains in PPR. For the impact copolymer, TEM and T1data show the existence of separate rubbery domains and thesedomains show the largest variation in TMob�Am

2 as a function of

temperature. The same behavior of TMob�Am2 is observed for PPR

samples. From this trend in mobility, it is most likely that chainsegments with an amount of ethylene units above a certain con-centration aggregate and form small (possibly nanoscopic) rubberydomains. Previously, it has been shown that phase separation canalready occur in the molten state for metallocene-derived statisti-cal ethylenee a-olefins copolymers [57,58]. The tendency for phaseseparation should be even larger for the samples studied here, asthey were synthesized using ZN catalysis. This type of catalysiscauses a more heterogeneous distribution of ethylene units along-and between chains. In addition, crystallization should further in-crease the tendency for phase separation due to higher concen-tration of ethylene-rich chain fragments in the amorphous phase[58].

Unlike IEPC23, the size of these domains in PPR should be sosmall that they cannot be detected by other techniques. In binaryblends of polypropylenes with small quantities of ethylene content,the presence of rubbery domains has been previously reported[26,32,58,59]. The situation is probably very similar in the case ofPPR. However, the domain sizes will be smaller than those observedin the binary blends due to a verywide distribution of the defects inPPR chains, which would explain why these domains have neverbeen observed by other methods.

3.3.6. Chemical composition of the amorphous phasesThe solution-state NMR data discussed above shows that

random ethylene insertion gives rise to number of microstructuresin PPR’s. The knowledge of the partition of these microstructures indifferent phases is necessary in order to understand structure-property relation. The chemical composition of the crystallinephase revealed that only a small amount of ethylene units areincorporated into the crystalline phase while majority of theethylene units are partitioned into the amorphous phase. A recentstudy of ZN catalyzed i-PP shows that at a temperature of 150 �C alldefects are excluded from the crystalline phase while previousstudies of the PPR polymers prepared with metallocene catalyst by

Page 8: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

Fig. 7. Two-dimensional 1He13C carbon detected heteronuclear correlation NMRspectrum of PPR7 at 50 �C.

Fig. 6. Temperature dependent change of the relative concentration and mobility factor (T2 relaxation constant) of the two amorphous components in the five samples. a) and c)represent the change in the relative concentration of the soft- and mobile-amorphous components, while b) and d) represent the change in the TSoft�Am

2 and TMob�Am2 . The points in

the green box represent the composition of the IEPC23 sample. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of thisarticle.)

V. Agarwal et al. / Polymer 55 (2014) 896e905 903

Vanderhart and co-worker [60e62] indicate that at room temper-ature ethylene-rich chain fragments are preferentially reside in theamorphous phase. A detailed analysis of the crystalline componentin our samples will be published elsewhere. In the discussion belowwe will focus on the chemical composition of soft- and mobileamorphous fractions, since the studies by Vanderhart could notprovide such information due to the lack of spectral resolution forthe amorphous phase.

At a temperature of 50 �C, a 13C detected 2D 1He13C correlationspectrum (Fig. 7) of PPR7 shows unique correlations that were usedto identify different microstructures in the amorphous phase. One-bond proton-carbon scalar couplings were used to record the 2Dspectrum in the absence of any high-power homonuclear decou-pling. The free evolution periods in this experiment ensures thatonly the amorphous phase is visible in the spectrum. The differentresonances in Fig. 7 were assigned based on solution-state NMRassignment [34,35] and use the nomenclature suggested by Car-man and Wilkes [29]. The peaks at 46.2, 38.0, 37.7, 33.3, 30.7 28.4and 24.7 ppm can be assigned to Saa, Sag, Sad, Tdd, Tbd, Tbb and Sbbrespectively. The two peaks at 13C chemical shifts of 30.3 and27.7 ppm most likely originate from Sgd/Sdd and Sbd units, respec-tively, while the methyl region shows three peaks at 22.2, 21.3 and20.5 ppm. Besides the fine structure of the resonances of themethylgroups, a total of nine additional chemical moieties could beidentified in the amorphous phase of PPR7. From this analysis, weconclude that the amorphous phase is rich in chains with triadsequence such as PPP, PEP, PPE, EPE, PEE and EEE, i.e. has a relativelyhigh ethylene concentration.

3.3.7. Temperature dependence of the chemical composition of theamorphous phase

From the T2 and chemical shift analysis described above, it ap-pears that chain segments rich in ethylene units have an enhanced

mobility compared to propylene-rich chains. This means thatrandom ethylene insertions do not only act as structural defects butalso as dynamic defects in PPR chains. In order to substantiate thishypothesis, direct carbon excited T2-filtered 13C NMR spectra forthe amorphous phase of PPR7 were acquired at different temper-atures (Fig. 8). The spectrum at 23 �C shows the highest peak in-tensity for resonances (red dotted lines) associated with chemicalmoieties due to ethylene insertion. Considering the overallethylene content, the intensity of peaks due to PP units in theamorphous phase (black lines) is very small compared to peaks of

Page 9: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

Fig. 8. 13C T2 filtered 13C NMR spectra of PPR7 acquired at different temperatures. Theresonances indicated by the red dotted lines are due to peaks originating fromethylene units in the sample, while the black dotted lines indicate PP peaks in theamorphous phase. (For interpretation of the references to colour in this figure legend,the reader is referred to the web version of this article.)

V. Agarwal et al. / Polymer 55 (2014) 896e905904

the ethylene units. The concentration of this ethylene-rich mobileamorphous fraction increases with increase in temperature(Fig. 4D). At 50 �C, the intensity of the peaks from PP and ethyleneunits is reversed compared to 23 �C spectrum. The spectra atdifferent temperatures show that while the chemical shifts of thelines assigned to ethylene units do not change over the entiretemperature range, whereas the chemical shifts observed for the PPunits constantly shift to higher ppm values. The chemical shift of eCH carbon of PP unit is 26.7 ppm in the crystalline phase and28.95 ppm in the solution-state. The chemical shifts in the crys-talline phase indicates a specific conformation with very little de-grees of freedomwhile a solution-state chemical shift represent themean conformational state of polymer chains resulting from fastchain dynamics. Therefore, the temperature dependent chemicalshift changes indicate an increased degree of freedom for PP-rich

Fig. 9. Rate dependence of the yield stress of (a) iPP homopolymer

chains at higher temperatures. Since the peaks from ethylene de-fects do not show a shift variation one can conclude that theethylene-rich units already have a high degree of motional freedomat room temperature. This simple experiment demonstrates thatfor the amorphous phase the insertion of ethylene units in PPchains should not only be interpreted as structural defects but alsoas dynamic defects.

3.4. Yield stress as a function of strain-rate for different PPR and therole of chain dynamics

The ultimate aim of the analysis of different phases is to un-derstand the mechanical properties in relation to the molecularstructure and dynamics of polymers. The effect of temperature andthe concentration of ethylene units on chain mobility in theamorphous phase, as determined by NMR, help in better under-standing the mechanical properties of PPR.

The rate and temperature dependence of the yield stress of iPP isillustrated in Fig. 9a. With increasing temperature, a smaller valueof the yield stress is observed. A clear change in the rate depen-dence is displayed within the studied temperature and strain raterange. The change in slope indicates that there are two moleculardeformation processes contributing to the yield stress [63]. Theyield kinetics of polypropylene can be accurately captured using theRee-Eyring modification of the original Eyring theory [57,64].Essentially, this theory assumes that two processes act indepen-dently and that their contributions to the stress are additive. Theseprocesses can be linked to intralamellar (“crystalline”) slip andinterlamellar (“amorphous”) slip [63]. Important is the observationthat the relative contribution of the amorphous phase to the yieldstress becomes increasingly important at lower temperatureswhenchain mobility in the amorphous phase decreases. The 1H NMR T2data (Fig. 3) indicates that below 30 �C chain mobility in theamorphous phase freezes out. Chain dynamics, which is involved instructural reorganizations in the amorphous phase during inter-lamellar slips, matches the strain rate upon approaching Tg. This isconfirmed by the dependence of yield stress on strain rate andtemperature. At the same temperature, a higher yield stress isobserved at faster deformation rate. At lower temperatures, chainmobility is slower and, consequently, yield stress is larger. Attemperatures above 50 �C and studied strain rates, the rate of chaindynamics in the crystalline phase (a-crystalline relaxation process)facilitates intralamellar slips and the crystalline phase becomes therate determining factor of the dependence of yield stress at a givenstrain rate.

Fig. 9b shows the rate dependence of the yield stress at roomtemperature for the different polypropylenes. Here we observe thatwith increasing ethylene concentration the relative contribution of

and (b) polypropylenes with increasing ethylene concentration.

Page 10: The chemical structure of the amorphous phase of propylene ...purl.tue.nl/920941759954952.pdf · The chemical structure of the amorphous phase of propylenee ethylene random copolymers

V. Agarwal et al. / Polymer 55 (2014) 896e905 905

theamorphousphase to theyield stressdecreasesdue to the increasein molecular mobility in the amorphous phase upon insertion ofethylene comonomer units (Fig. 6b). The larger the concentration ofthe ethylene in the amorphous phase the more mobile chains seg-ments in the amorphous phase are present (Fig. 4). The amorphousphase of PPR7 ismoremobile than in PPR4. This observation explainsthe lowering of yield stress at room temperature with increasingethylene content due to lower barriers for interlamellar slips.

4. Conclusions

In-depth investigations of the phase composition and thechemical microstructure of the phases have been conducted for fivecommercial grades of polypropylene-based polymers preparedusing heterogeneous ZN catalysis. Solution-state NMR was used toidentify and quantify the microstructures present in the polymerchains that are generally consistent with ZN based polymerization.WAXD and SAXS data show a reduction in crystallinity, meltingtemperature, long-period and crystal thickness with increase inethylene content. In these samples, the temperature dependenttime-domain 1H NMR experiments unambiguously establish thepresence of a fourth fraction in addition to the crystalline phase, thecrystaleamorphous interface and the soft amorphous fraction. Thisfourth fraction is observed in all samples at temperatures 20e30 �Cabove Tg. This fraction has rubber-like dynamic properties andconsists of segments/chains rich in ethylene units.

The decrease in yield stress with increasing ethylene contentand temperature can be related to the greater mobility of chains inthe amorphous phase. More studies placing a greater emphasis onthe detailed knowledge of the chain microstructures and theirdistribution over various phases need to be performed to establishdetailed structure-properties relations. Particularly interesting willbe such studies performed in-situ under mechanical load or duringdeformation. Equipment to allow such studies using solid-stateNMR is under development in our laboratory.

Acknowledgments

Thework described in this paper is part of the Research Programof the Dutch Polymer Institute (DPI), Eindhoven, the Netherlands,project number #710. The authors thank Gerrit Janssen, HansJanssen, and Jan van Os (Radboud University) for their technicalsupport. We furthermore acknowledge the Netherlands ScienceOrganization NWO for their support of the solid-state NMR facilityfor advanced materials research.

Appendix A. Supplementary data

Supplementary data related to this article can be found at http://dx.doi.org/10.1016/j.polymer.2013.12.051.

References

[1] Cavallo D, Azzurri F, Floris R, Alfonso GC, Balzano L, Peters GW. Macromole-cules 2010;43(6):2890e6.

[2] Padden Jr FJ, Keith HD. J Appl Phys 1959;30(10):1479e84.[3] Norton DR, Keller A. Polymer 1985;26(5):704e16.[4] Hsu CC, Geil PH, Miyaji H, Asai K. J Polym Sci Part B Polym Phys 1986;24(10):

2379e401.[5] Caldas V, Brown GR, Nohr RS, MacDonald JG, Raboin LE. Polymer 1994;35(5):

899e907.[6] Farrow G. Polymer 1963;4(C):191e7.[7] Natale R, Russo R, Vittoria V. J Mater Sci 1992;27(16):4350e4.[8] Martorana A, Piccarolo S, Sapoundjieva D. Macromol Chem Phys 1999;200(3):

531e40.[9] Zia Q, Radusch HJ, Androsch R. Polymer 2007;48(12):3504e11.

[10] van Erp TB, Balzano L, Peters GWM. Acs Macro Lett 2012;1(5):618e22.

[11] Liedauer S, Eder G, Janeschitzkriegl H, Jerschow P, Geymayer W, Ingolic E. IntPolym Process 1993;8(3):236e44.

[12] Keller A, Kolnaar HWH. Flow-induced orientation and structure formation.Materials Science and technology. Wiley-VCH Verlag GmbH & Co. KGaA; 2006.

[13] Sherman LM. Plast Technol 2003;49(11):37e9.[14] Glenz W. Polypropylen (PP) 2004;94(10):69e71.[15] Van Lierde S. Med device Technol 2004;15(5):33e4.[16] Carraher Jr CE. Polym News 2005;30(6):186e8.[17] Stadlbauer M, Gahleitner M, Wachholder M, Wolfer R. Kunstst Plast Eur

2005;95(10):60e7.[18] Toensmeier PA. Plast Eng 2007;63(1):44e7.[19] Guo P, Lu M, Lu Y, Zhang S. Shiyou Huagong/Petrochemical Technol

2011;40(6):679e84.[20] Mileva D, Zia Q, Androsch R. Polym Bull 2010;65(6):623e34.[21] Hosier IL, Alamo RG, Lin JS. Polymer 2004;45(10):3441e55.[22] Poon B, Rogunova M, Chum SP, Hiltner A, Baer E. J Polym Sci Part B Polym Phys

2004;42(23):4357e70.[23] Poon B, Rogunova M, Hiltner A, Baer E, Chum SP, Galeski A, et al. Macro-

molecules 2005;38(4):1232e43.[24] Jeon K, Palza H, Quijada R, Alamo RG. Polymer 2009;50(3):832e44.[25] De Rosa C, Auriemma F, De Ballesteros OR, Resconi L, Camurati I. Chem Mater

2007;19(21):5122e30.[26] Ibbett RN. Nmr spectroscopy of polymers. Blackie Academic & Professional;

1993.[27] Randall JC. J Macromol Sci Part C Polym Rev 1989;29(2e3):201e317.[28] Busico V, Cipullo R. Prog Polym Sci Oxf 2001;26(3):443e533.[29] Carman CJ, Harrington RA, Wilkes CE. Macromolecules 1977;10(3):536e44.[30] De Rosa C, Auriemma F, De Ballesteros OR, Resconi L, Camurati I. Macro-

molecules 2007;40(18):6600e16.[31] Chen J, Cao Y, Kang J, Li H. J Macromol Sci Part B Phys 2011;50(2):248e65.[32] Gerhard H, Isa F, Matthew P. Application of melt-state NMR spectroscopy for

polyolefin characterization in IndustryIn NMR spectroscopy of polymers:innovative strategies for complex macromolecules, vol. 1077. AmericanChemical Society; 2011. pp. 401e15.

[33] Cheng HN. Macromolecules 1984;17(10):1950e5.[34] Wang WJ, Zhu S. Macromolecules 2000;33(4):1157e62.[35] Kakugo M, Naito Y, Mizunuma K, Miyatake T. Macromolecules 1982;15(4):

1150e2.[36] Alamo R, Domszy R, Mandelkern L. J Phys Chem 1984;88(26):6587e95.[37] Flory PJ. J Chem Phys 1949;17(3):223e40.[38] Flory PJ. Trans Farad Soc 1955;51:848e57.[39] Wunderlich B. Macromolecular physics, vol. 3. Academic Press; 1980.[40] McBrierty V, Douglass D, Falcone D. J Chem Soc Farad Trans 2 1972;68:1051e

9.[41] Litvinov VM. Macromolecules 2006;39(25):8727e41.[42] Fedotov VD, Schneider H. Structure and dynamics bulk polymer by NMR

methods. In: Diehl P, Fluck E, Gunter H, Kosfeld R, Seeling I, editors. NMR basicprinciples and progress. Berlin: Spinger; 1989.

[43] Cohen-Addad JP, Dupeyre R. Polymer 1983;24(4):400e8.[44] Brereton MG. Macromolecules 1990;23(4):1119e31.[45] Dadayli D, Harris R, Kenwright A, Say B, Sünnetçioǧlu MM. Polymer

1994;35(19):4083e7.[46] Hedesiu C, Demco DE, Kleppinger R, Vanden Poel G, Gijsbers W, Blümich B,

et al. Macromolecules 2007;40(11):3977e89.[47] Hedesiu C, Demco DE, Remerie K, Blümich B, Litvinov VM. Macromol Chem

Phys 2008;209(7):734e45.[48] Schreurs S, Francois J, Adriaensens P, Gelan J. J Phys Chem B 1999;103(9):

1393e401.[49] Tanaka H, Inoue Y. Polym Int 1993;31(1):9e13.[50] Weglarz WP, Peemoeller H, Rudin A. J Polym Sci Part B Polym Phys

2000;38(19):2487e506.[51] Dujourdy L, Bazile JP, Cohen-Addad JP. Polym Int 1999;48(6e7):558e64.[52] Litvinov VM, Soliman M. Polymer 2005;46(9):3077e89.[53] Gahleitner M, Jääskeläinen P, Ratajski E, Paulik C, Reussner J, Wolfschwenger J,

et al. J Appl Polym Sci 2005;95(5):1073e81.[54] Kentgens APM, Veeman WS, Van Bree J. Macromolecules 1987;20(6):1234e7.[55] Resing HA. Mol Cryst 1969;9(1):101e32.[56] Dutta N, Choudhury N, Haidar B, Vidal A, Donnet J, Delmotte L, et al. Polymer

1994;35:4293e9.[57] Ree T, Eyring H. J Appl Phys 1955;26(7):8.[58] Kamdar AR, Hu YS, Ansems P, Chum SP, Hiltner A, Baer E. Macromolecules

2006;39(4):1496e506.[59] Lohse DJ. Polym Eng Sci 1986;26(21):1500e9.[60] VanderHart D, Alamo R, Nyden M, Kim M, Mandelkern L. Macromolecules

2000;33(16):6078e93.[61] Alamo R, VanderHart D, Nyden M, Mandelkern L. Macromolecules

2000;33(16):6106e19.[62] Li Z, Miyoshi T, Sen MK, Koga T, Otsubo A, Kamimura A. Macromolecules

2013;46(16):6507e19.[63] van Erp TB, Cavallo D, Peters GWM, Govaert LE. J Polym Sci Part B Polym Phys

2012;50(20):1438e51.[64] Roetling JA. Polymer 1966;7(7):303e6.