Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences...

58
MSc Chemistry Analytical Sciences Literature Thesis Recent developments in polyolefin fractionations by Jeroen Knooren December 2013 Supervisor: Dr. W. Th. Kok

Transcript of Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences...

Page 1: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

MSc Chemistry

Analytical Sciences

Literature Thesis

Recent developments in polyolefin fractionations

by

Jeroen Knooren

December 2013

Supervisor:

Dr. W. Th. Kok

Page 2: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 1 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Page 3: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 2 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

MSc Chemistry

Analytical Sciences

Literature Thesis

Recent developments in polyolefin fractionations

by

Jeroen Knooren

December 2013

Jeroen (J.A.A.) Knooren (UvA) 6413501

University of Amsterdam - FNWI

Master of Chemistry – Analytical Sciences

Supervisor UvA: Dr. W. Th. Kok

November 20th, 2013

Page 4: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 3 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Page 5: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 4 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Abstract Since the total polyolefin (PO) production exceeds 100.000 kiloton per year, while still showing an

exponential growth, the PO research is one of the most important areas for both the academic world as

the industrial companies. As a result of the increasing requirements in properties and performance of

PO and the broadening of the application field by applying structural changes, using different catalysts

within the polymerization process, the analytical requirements for complete characterization of

polyolefin structures is challenging. Within this report several techniques for the characterization of the

chemical composition of polyolefins are presented and compared.

The two main fractionation techniques available adsorption and crystallization are complementary to

each other in terms of separation characteristics. For adsorption based fractionations high temperature

high performance liquid chromatography (HT-HPLC) seems to be preferred over temperature gradient

interaction chromatography (TGIC) because of the superior efficiency and separation power; however,

HT-HPLC is limited to detection by evaporative light scatter detection (ELSD) whereas TGIC is applicable

to a wide variety of detectors. Regarding the crystallization techniques, temperature rising elution

fractionation (TREF), successive self-nucleation and annealing (SSA) and crystallization elution

fractionation (CEF), it can be clearly stated that CEF performed better than TREF. Whether CEF is also

preferred over SSA depends on the application, for a rough characterization of molecular properties, SSA

might be sufficient, whereas CEF will give a more comprehensive view for somewhat longer analysis

times. Also the choice between an interaction based or a crystallization based technique depends on the

analytical question that has to be answered.

Another technique for polyolefin fractionations is so called cross-fractionations, where two -preferably

orthogonal- separation techniques are applied in a comprehensive way. Regarding 2D separations, high

temperature size exclusion chromatography (HT-SEC) coupled to HT-HPLC, in which the added

dimension can physically separate the fractions, whereas it is actually applied as a molecular weight

sensitive detector, shows to be a very promising technique for the separation of complex molecules

especially when compared to TREFxSEC. Nevertheless, there are some inconsistencies between the

molecular weight polydispersity determined by SECxHT-HPLC and those determined solely by SE. This

shows that the newly developed technique -it was only introduced in 2010- still seems to show some

teething problems; which will most likely be resolved in the upcoming years, making this an extremely

powerful techniques for complex polyolefin characterizations.

Apart from SEC combined to another technique for either crystallization or adsorption/interaction based

fractionation, due to the complementarity TGICxCEF or even SECxCEFxTGIC might be good combinations

for a more complete characterization of polyolefin.

Page 6: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 5 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Table of Contents

Abstract ......................................................................................................................................... - 4 -

Table of Contents ........................................................................................................................... - 5 -

1. Introduction ........................................................................................................................... - 6 -

2. General information ............................................................................................................... - 9 -

2.1. What Are Polymers ....................................................................................................................... - 9 -

2.2. What Are Polyolefins .................................................................................................................. - 10 -

2.3. Properties of Polyolefins ............................................................................................................. - 11 - 2.3.1. Material Properties of Polyolefins ................................................................................... - 11 - 2.3.2. Molecular Properties of Polyolefins ................................................................................ - 12 -

2.4. Characterization of Polyolefins ................................................................................................... - 15 -

2.5. Detection & Quantification ......................................................................................................... - 16 - 2.5.1. Infrared Detector ............................................................................................................. - 16 - 2.5.2. Viscometer ...................................................................................................................... - 16 - 2.5.3. Evaporative Light Scattering Detector ............................................................................ - 16 - 2.5.4. Refractive Index Detector ................................................................................................ - 16 - 2.5.5. Light Scattering Detector ................................................................................................ - 17 -

3. Crystallization Based Fractionation Methods ......................................................................... - 18 -

3.1. Temperature Rising Elution Fractionation .................................................................................. - 18 -

3.2. Crystallization Elution Fractionation ........................................................................................... - 22 -

3.3. Successive Self-Nucleation and Annealing & Stepwise Crystallization ....................................... - 24 -

4. Interaction Based Fractionation Methods .............................................................................. - 32 -

4.1. High Temperature High Performance Liquid Chromatography .................................................. - 32 -

4.2. Temperature Gradient Interaction Chromatography ................................................................. - 38 -

5. Two dimensional fractionations ............................................................................................ - 44 -

6. Outlook ................................................................................................................................ - 48 -

7. Conclusion & Recommendation ............................................................................................ - 50 -

Literature references.................................................................................................................... - 52 -

Page 7: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 6 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

1. Introduction Some of the most important commodity polymers are polyolefins (PO). Particularly polyethylene (PE)

and polypropylene (PP) are globally the major plastic materials. The PO production exceeds 100.000

kiloton per year, while still showing an exponential growth [1]. The importance of PE and PP in the

world-market of plastics is related to the diversity of possible structural differences and their influence

on the end characteristics of the polymer. One of the main catalyst species in the polymerization of PO is

the Ziegler-Natta catalyst, which was discovered in the early 1950s. This catalyst led to a drastic

reduction of the production costs, and enabled synthesis of new types of polyolefins. Further

optimization of this catalyst throughout the years resulted in very active and stereospecific catalysts,

which, in combination with improved processing techniques, led to a broad range of possible

applications for PO. As a result, polyolefins can now fulfil the full range of polymeric demands, from hard

thermoplastics to soft elastomers, while still using solely ethylene and propylene as building blocks.

In order to tailor the properties of the polymer to the required application, it is vital to correlate end

properties of the polymer with molecular characteristics as measured by analytical techniques. Several

studies towards the effect of molecular characteristics on polymer properties have been conducted in

the past. Especially the effect of copolymerization and stereo- and region-defects, as studied by De Rosa

et al. [2-4], on the properties and crystallization of PP increased the understanding.

Just like polymers consisting of a single monomer, like PE or PP, a polymer can also consist of different

monomers; in that case it is referred to as a copolymer. A copolymer is built from different monomeric

units, resulting in a different composition. If ethene is polymerized with the addition of α-olefins (mono-

unsaturated paraffins with the double bond in the first position, e.g. 1-hexene), only two carbons of the

α-olefin become part of the backbone, while any additional carbons will be observed as a side chain.

This leads to a short chain branching distribution (SCB or SCBD) since the comonomer will be built into

the polymer at random locations.

The distribution of the α-olefin along the polymer chain is particularly important, since many

copolymers can exhibit highly heterogeneous intra- and/or inter-molecular-co-monomer distributions

[5]. The development of reliable techniques to characterize the short chain branching distribution

(SCBD) has been the goal of many researchers in the field, especially because both the physical and

mechanical properties are highly influenced by the SCBD.

Within polymer characterization, important properties are the molecular weight distribution (MWD),

the chemical composition distribution (CCD) consisting of the SCB, comonomer content, etc. and the

molecular topology, e.g. long chain branching (LCB). For the characterization of the MWD, first of all

there is size exclusion chromatography (SEC), which for decades is the main tool, used for the

characterization of MWD, LCB and SCB. In SEC, the sample is passed through a column filled with

particles with a range of different pore sizes. The separation in SEC is based on hydrodynamic volume,

which determines whether or not a molecule fits in the pores. Large molecules, not able to penetrate

Page 8: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 7 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

the pores, pass in between particles and elute first, while the smallest molecules fit in all the pores, and

thus have a longer pathway, these will elute last.

Another of these techniques is Field Flow Fractionation (FFF), which is a separation technique in which a

flow passes through a narrow channel, while, perpendicular to that flow a second field (e.g. thermal

gradient) is applied, separating the compounds over the width of the channel. As a result of the laminar

flow through the channel, high flow speed in the middle, zero flow at the wall, the elution time, and thus

separation, also depends on the position of the compound within the channel, enhancing the

separation. This technique was already developed in 1966 by Giddings [6], and can be used as a tool to

determine the molecular weight distribution, similar to SEC. Furthermore there are some additional

MWD based separation techniques like molecular topology fractionation (MTF), slalom chromatography

and hydrodynamic chromatography (HDC); these are however not further discussed within this report.

One of the main techniques for the characterization of the CCD is temperature rising elution

fractionation (TREF), introduced by Wild et al. [7], in 1991. This separation technique is based on

crystallization of the polymer, which is again linked to the SCBD. After slowly crystallizing a polymer from

solution onto an inert support, e.g. glass beads, the polymer is dissolved and eluted at successively rising

temperatures. As a result of the slow crystallization, the influence on the molecular weight is limited,

since separation is mainly determined by differences in crystallinity dependant on e.g. SCBD differences.

Owing to the long measurement times of TREF, other techniques, similar to TREF have been developed.

One of these techniques is so-called crystallization analysis fractionation (CRYSTAF) [8] (1994), which is

in principle the same as TREF; however, CRYSTAF monitors the concentration of the polymer in solution

during the crystallization stage, whereas TREF monitors the concentration of the polymer in solution

during the elution stage. Thus, CRYSTAF simply skips the elution step which is performed in TREF,

resulting in shorter analysis times. However, since the crystallization in TREF is the most time-

consuming, CRYSTAF still has a significant analysis time. Another method similar to TREF is crystallization

elution fractionation (CEF), which is more or less an improved version of TREF with shorter analyses

times and higher resolutions. The difference between TREF and CEF is the fact that during crystallization

a low flow rate is applied. Subsequently, just like with TREF the temperature is gradually increased for

elution of the compounds. The addition of “pre-fractionation” during the precipitation process, leads to

an increased resolution.

Apart from TREF, CEF and CRYSTAF which make use of a solvent during fractionation, some other

techniques do not require a solvent for the fractionation of polymers by CCD. These techniques, called

thermal fractionation techniques, are able to fractionate a polymer as it is, by melting and solidifying the

polymer itself. These techniques, e.g. successive self-nucleation and annealing (SSA) and stepwise

crystallization (SC), are performed in combination with differential scanning calorimetry (DSC) as

analytical tool. A method similar to SSA-DSC was first introduced by Gray et al. [9], in 1964, whereas SC-

DSC was first shown by Varga et al. [10] in 1976. SC-DSC, as the name implies, step-wise cools down the

polymer, after heating it above the melting temperature, until it completely solidifies. SSA-DSC also

Page 9: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 8 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

starts with heating the polymer above the melting temperature, to subsequently cool-down until

complete solidification, however, after the cool down the polymer is heated again to a temperature

slightly lower than the previous melting temperature, followed by a new cooling step. This process is

then repeated several times.

Furthermore there are two, interaction based, CCD separation techniques, being high temperature high

performance liquid chromatography (HT-HPLC) and temperature gradient interaction chromatography

(TGIC). As the name already implies, HT-HPLC is a HPLC method, ran at higher temperatures, typically

above 100°C, and as in most HPLC methods, the elution is performed isothermally using a solvent

gradient. For HT-HPLC this concept has not changed, however, TGIC, which was proposed as a method

only very recently, uses a temperature gradient, as already implied by the name, which is the opposite

of HPLC, since TGIC is performed isocratically. For both HT-HPLC and TGIC, the separations are based,

among others, on a combination of the interaction between the stationary phase and the molecules,

and the solubility of the molecules in the mobile phase.

In order to further enhance separation capabilities, several groups are publishing work regarding so-

called cross-fractionations. These cross-fractions are basically two dimensional separations, e.g. SEC-

TREF to get a two dimensional plot of the MWD on the x-axis and the SCBD of copolymers, or the

tacticity of PP, on the y-axis.

Increased complexity in polymer composition requires improved analytical techniques for correct

characterization. With a technique like SEC, HDC or FFF, a fractionation can be performed which is

mainly based on the molecular weight distribution (MWD) of the polymer, however, based on the scope

of this research, MWD separation techniques will not be further discussed. Since TREF methods are

currently up and running within DSM Resolve, a movement towards CRYSTAF, which offers similar

results within a slightly shorter time-span by compromising the resolution, does not seem viable;

therefore also CRYSTAF is not further discussed within this report.

In the upcoming chapter, some general information about polymers is presented. Chapter 3 contains the

crystallization based fractionation techniques like TREF and CEF; furthermore SSA and SC as thermal

fractionation techniques will be compared and discussed. In the subsequent chapter more details will be

given about the interaction based separation techniques TGIC and HT-HPLC. Chapter 5 will discuss

developments regarding cross-fractionations followed by an outlook for polyolefin fractionations in

chapter 6. In chapter 7 the findings will be concluded and recommendations will be made.

Page 10: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 9 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

2. General information In this chapter some general information regarding the discussed topics is.

2.1. What Are Polymers A polymer, also referred to as macromolecule, consists of a repetition of a single type, or a few different

types, of building blocks, the so-called monomers. These polymers are synthesized through the process

of polymerization, the reaction of monomer molecules in a chemical reaction forming polymer chains.

Butane and hexane could already be seen as a polymer consisting of two or three ethene monomer

building blocks respectively, although these molecules are not produced through a polymerization

process.

Regarding the origin of polymers, they can be divided in three different groups:

- natural polymers, e.g. cotton

- semi-synthetic polymers (produced from natural polymers), e.g. cellulose

- synthetic polymers (polymerization reaction of monomers), e.g. polyethylene

The chain of a polymer can be built from a single type of monomer, resulting in a so-called homo-

polymer, or from a combination of different types of monomers, resulting in a so-called co-polymer. For

copolymers there are several different categories: random, alternating, block and graft copolymers, as

shown in Table 1. Whereas for random and alternated copolymers the name already implies its

distribution, block copolymers have a linear chain in which different units at different length of the same

monomer, while graft copolymers have a branched chain with a different monomer in the branch than

in the backbone.

Table 1: Different copolymer compositions; A and B representing two different types of monomers.

Apart from the skeletal structures, polymers can also be classified in three groups based on properties:

- Elastomers

o Elastic, stretchable, rubberlike polymers with high amount of crosslinking; cannot be

reshaped after curing.

- Thermosets

o Network structure resulting in rigid material; cannot be reshaped after curing.

Page 11: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 10 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

- Thermoplastics (either crystalline [ordered structure] or amorphous [unorganized])

o Linear of branched structured polymers which are solid when cooling and melt while

heating, enabling the possibility to reshape the polymer.

Crystalline polymers show an ordered structure, having parallel chains in the solid phase, whereas

amorphous polymers are incapable of creating this organization, leading to uncontrolled folding of the

chain. Since the crystallization of thermoplastics is not a straight forward process the polymers become

semi-crystalline materials, as shown in Figure 1. Since crystalline materials are hard and rigid and

amorphous materials are hard and elastic, the semi-crystalline materials result in strong and ductile

polymers.

Figure 1: Semi-crystalline polymer with crystalline and amorphous regions.

2.2. What Are Polyolefins Examples of thermoplastics are polyolefins (PO), which are among the most important commodity

polymers. Particularly polyethylene (PE) and polypropylene (PP) are major global plastic materials.

Polyolefins are produced using olefins (unsaturated paraffins) as monomer building blocks. Some

product examples in which PO’s are used are shampoo bottles, plastic foil, films, gas pipes and garden

furniture.

For polyethylene several different polymerization methodologies are used in order to obtain different

structures and densities. Some examples of polyethylene are high density poly ethylene (HDPE), low

density polyethylene (LDPE), linear low density polyethylene (LLDPE), and very low density polyethylene

(VLDPE), as illustrated in Figure 2. Whereas HDPE is unbranched resulting in a high crystallinity, LDPE is

strongly branched which results in a more amorphous character. In LDPE the crystallinity is hindered by

long chain branching (LCD) whereas LLDPE is more hindered by short chain branching (SCB), as a result

of the branched chain not fitting into the crystals.

Page 12: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 11 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 2: Different types of PE and their density and crystallinity.

2.3. Properties of Polyolefins Properties of polyolefins can be divided into two categories. The material properties which are the

properties at the macroscopic level; the properties observed and utilized in normal day life; within the

scope of the research the material properties like the tensile strength or Young’s modulus are not

further discussed. Additionally the polyolefins have specific molecular properties which are influenced

by the used catalyst, co-monomer, process parameters, additives, etc. These molecular properties

require complex analytical techniques, however, even with all available analytical techniques; full

characterization and full understanding of the functioning of a polymer might still remain vague.

2.3.1. Material Properties of Polyolefins The material properties can be divided into mechanical properties and phase properties. Within the

material properties of polyolefins solely the phase properties melting point and glass transition

temperature are briefly described within this report, since these are the characteristics determined by

DSC. For other material properties like the tensile strength or the Young’s modulus (mechanical

properties) other literature should be consulted.

Regarding the phase behaviour, an important aspect is the melting point of the polymer (Tm). For

polyolefins the melting point, which refers to a transition from (semi-)crystalline towards amorphous, is

taken into account as a characterization parameter as a result of the broad melting point range of

polyolefins, approx. 165°C for PP and approx. 130°C – 105°C for PE depending on crystallinity; higher

crystallinity results in a higher melting temperature.

Another property of phase behaviour is the glass transition temperature (Tg), which is the temperature

at which the polymer changes from a glassy state into a liquid and/or rubbery state. At low temperature

the polymer exists in the glassy / solid state. After reaching the glass transition temperature, the

polymer chains gain movement and change conformation. By increasing the temperature above the

glass transition temperature, some polymers will melt and the polymer chains can move freely. The

glass transition temperature can be influenced by altering the degree of branching or crosslinking.

Another method of influencing the Tg is the addition of plasticizers into the polymer.

Page 13: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 12 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

2.3.2. Molecular Properties of Polyolefins Polyolefins can be generally divided into homopolymers, consisting of a single type of monomer and

copolymers, consisting of several different types of monomers. Based on this division, the molecular

properties can be divided into properties applicable to all polyolefins, or those only applicable to

copolymers.

2.3.2.1. Applicable To All Polyolefins The molar mass distribution, crystallinity and additives are molecular properties that are applicable to all

polymers.

2.3.2.1.1. Molar Mass Distribution A very important property of polyolefins or even polymers in general, is the molar mass distribution.

This value quantifies the average molecular mass of a polymer. With respect to the average molecular

mass, it can be calculated in different ways, one being the number average molecular mass (Mn), in

which the molecular mass is simply calculated by dividing the sum of the different molecular masses

(per molecule) by the number of molecules, as shown in Equation 1.

Equation 1

Another method is the calculation of the weight average molecular mass (Mw), which corresponds to the

average weight of a monomer which is present within the polymer. The average molecular mass (Mw) is

calculated as shown in Equation 2.

Equation 2

The third method for the calculation of the molar mass is the Z average molar mass (Mz), which

correlates more to the higher molecular weight parts of the molecular weight distribution. In general

the Mn corresponds to the lower molecular mass fraction, the Mw gives information about the middle of

the molecular mass distribution and the Mz represents a value of the higher molar mass range. The Z-

average molecular mass (Mz) is calculated as shown in Equation 3.

Equation 3

Page 14: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 13 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

The correlation between the weight average molecular mass (Mw) and the number average molecular

mass (Mn) is called the polydispersity. The polydispersity is a method for distinguishing between a

narrow and a broad molecular weight distribution within a polymer. The polydispersity index (PDI) can

simply be calculated by Equation 4, and represents the width of the molar mass distribution.

Equation 4

n

w

M

MPDI

2.3.2.1.2. Crystallinity Furthermore, crystallinity as discussed in chapter 2.1 is an important factor in the characterization of

polyolefins. The crystallinity refers to the structural order, which has an influence on for instance the

hardness, density, transparency.

2.3.2.1.3. Additives Another molecular property of polymers, which have an influence on the material properties are the

additives added to the polymer. Some examples of additives added to a polymer are:

- Plasticizers, which are used to influence the Tg.

- Stabilizers, used to make the polymers more resistant to temperature-driven oxidation.

- Talc is added to influence crystallinity and thus the modulus of the polymer.

2.3.2.2. Applicable To Copolymers Only The molecular properties as described below require the use of different monomer types, and are

therefore only applicable to copolymers.

A copolymer is built from different monomeric units, resulting in a different composition. If ethene is

polymerized with the addition of α-olefins (mono-unsaturated paraffins with the double bond in the first

position, e.g. 1-hexene), only two carbons of the α-olefin become part of the backbone, while any

additional carbons will be observed as a side chain. This leads to a short chain branching distribution

(SCB or SCBD) since the comonomer will be built into the polymer at random locations or in more blocky

sequences.

If during the polymerization process a low concentration of diolefine (e.g. butadiene) is dosed into the

reaction of ethylene, the butadiene can react with two ethylene molecules, one being part of the

polymer backbone and the other leading to a side chain formed on the additional in-saturation from the

butadiene, this results in the so-called long chain branching distribution (LCBD) and/or network

formation, e.g. polybutadiene rubber.

Another aspect is the chemical composition distribution (CCD) or the sequence distribution (SD) which is

the order of the different monomers within the polymer. When performing a block polymerization, this

is called the block length distribution (BLD), which corresponds to the block lengths of the different

Page 15: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 14 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

monomeric units within the polymer backbone. These distributions are of course also related to the

SCBD, since the different units will give different branches and thus influence the SCBD.

Furthermore there are some other monomeric distributions, like the grafting distribution (monomer A in

the backbone with different side chain lengths of monomer B) and the functional type distribution (end-

group distribution); a schematic overview is given in Figure 3.

Figure 3: Schematic view of different distributions in polyolefins; A and B are monomeric units, C and D are end groups; reprinted from [11].

2.3.2.3. Applicable To Homopolymers Only The tacticity distribution is an example of a molecular property which is only applicable to

homopolymers. In case of a polyethylene, the monomer, ethylene, is completely built into the polymer

backbone, as shown in Figure 4. In case of polypropylene the additional methyl-group of the monomer

can be located on both sides of the backbone, which leads to stereo chemical distributions, referred to

as tacticity. There are three different types of tacticity sequences, being syndiotactic, all methyl groups

on the same side, isotactic, the methyl groups on alternating sides, and atactic which corresponds to a

random arrangement, as shown in Figure 5.

Figure 4: Ethylene built into polymer backbone.

Figure 5: Tacticity of polypropylene, a) syndiotactic, b) isotactic and c) atatic distribution.

Page 16: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 15 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

2.4. Characterization of Polyolefins Looking at the different molecular and material properties as discussed in the previous chapter, it can be

expected that a complete characterization of polyolefins is a complex science, especially, since the list of

the properties as discussed can be further extended by for instance morphological properties and

molecular topology. For the characterization of the molecular properties as discussed above, a short

overview is presented in Figure 6 below, in which different characterization techniques are connected to

the different aspects of the polymer properties. Out of these properties, the main focus within this

report is the characterization of the chemical composition distribution or the short chain branching

distribution.

Figure 6: Overview of some of the properties of a polyolefin including characterization techniques. TGIC=Temperature Gradient Interaction Chromatography; HT-HPLC: High Temperature High Performance Liquid Chromatography; NMR: Nuclear Magnetic Resonance Spectroscopy; CRYSTAF: CRYSTAlication Fractionation; CEF: Crystallization Elution Fractionation; TREF: Temperature Rising Elution Fractionation; DSC: Differential Scanning Calorimetry; SSA: Successive Self-nucleation and Annealing; SC: Step-wise Crystallization; SEC: Size Exclusion Chromatography; FFF: Field Flow Fractionation and HDC: HydroDynamic Chromatography.

Page 17: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 16 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

2.5. Detection & Quantification In this paragraph a few details are given about some of the detectors commonly used for the

macromolecular detection and quantification. With most of the fractionation techniques discussed in

the upcoming chapters one or more of these detectors can be coupled in order to acquire specific

information. In general the detectors can be divided in three categories:

- Concentration detectors (e.g. infrared, evaporative light scattering detector and refractive index

detector), used for quantification.

- Mass sensitive detectors (e.g. light scattering detector and viscometer), used for the molar mass

determination.

- Chemical composition detectors (e.g. infrared detector).

2.5.1. Infrared Detector The infrared detector, as the name already states, is a detector sensitive to infrared radiation of the

passing effluent. Infrared radiation (700nm – 1mm) is similar to visible light (380nm – 700 nm),

although, due to the longer wavelengths in infrared they are not visible for the human eye. The

advantage of infrared is that the absorbed wavelength is dependent of the chemical neighbourhood of

the atom. Therefore, apart from solely quantitation, this detector can also be used for identification

purposes. A disadvantage is the fact that this detector is sensitive to the background, e.g. the used

solvent; furthermore the infrared region has lower intensity photons leading to lower sensitivities.

2.5.2. Viscometer The viscometer is a detector which measures the intrinsic viscosity of the passing effluent, which can be

used for the determination of the molecular size, conformation and structure. Since different solvents

have different viscosities, the viscometer can only be used in setups using a single solvent.

2.5.3. Evaporative Light Scattering Detector The idea behind the evaporative light scattering detector (ELSD) is the possibility to evaporate the

solvent, whereas the polymers do not evaporate. The not-evaporated polymers are then transported as

a smoke into a lamp or a laser beam, with a light sensitive detector measuring the scatter of the laser

beam. This detector, however, is not linear with concentration and furthermore, while using different

solvents, e.g. a gradient during a HT-HPLC run, the sensitivity during the run might vary slightly.

2.5.4. Refractive Index Detector The refractive index (RI) is a measure for the propagation of radiation through a medium. The RI

detector measures the change in RI of the column effluent passing through the flow cell. The sensitivity

of the RI is highly dependent of the RI difference between the reference and the effluent passing

through the cell. Furthermore, since different solvents have different RI’s the RI detector cannot only be

applied for single solvent chromatographic setups.

Page 18: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 17 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

2.5.5. Light Scattering Detector With the light scattering detector, the Rayleigh scatter of the polymer in the effluent is measured under

one or more angles. This is a measure for both the molecular weight and the concentration, which can

be deducted by using the Zimm equation, as shown in Equation 5.

Equation 5

𝐾𝐶

𝑅∅=

1

𝑀𝑃∅+ 2𝐴2𝐶

In which C = concentration; A2 = second virial coefficient of solution; RØ = excess of Rayleigh scatter; M =

molecular weight; PØ = particle scattering factor and K = composite of optical and fundamental

constants.

Page 19: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 18 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

3. Crystallization Based Fractionation Methods In this chapter the theoretical aspects of the crystallization based fractionation methods temperature

rising elution fractionation (TREF), crystallization elution fractionation (CEF), stepwise crystallization (SC)

and successive self-nucleation and annealing (SSA), are discussed.

3.1. Temperature Rising Elution Fractionation Temperature Rising Elution Fractionation (TREF) is an often applied method for polyolefin fractionations.

The separation within TREF is based on the crystallizability of the polymer chains. Since most polyolefins

are crystallizable, TREF is a well suited technique for the fractionation of these polymers.

TREF consists of several steps; initially the sample is dissolved in a high boiling point solvent, e.g. xylene

or trichlorobenzene (TCB). This step can be performed at normal heating rates, since the speed is not

critical for the end result. The polymer solution is mixed with an inert support, e.g. glass beads; this

procedure takes place within a column, of which the temperature is controlled. In the next step the

temperature is slowly decreased. Typical cool down rates are around 0.1°C per minute, resulting in very

long cool down times.

The effect of cooling is crystallization of polymer as function of crystallinity. At a lower temperature, the

less crystalline part remains dissolved whereas the more crystalline part will precipitate. If the cooling

process is too fast, co-crystallization of fractions which have different crystallinities can occur. During

this cool down process, layers of different crystalline fractions, almost independent of the molecular

weight [12], are deposited on the inert support. As schematically illustrated in in Figure 7 (reprinted

from [12]), the most crystalline fraction forms the inner layer and the least crystalline fraction the outer

layer.

Figure 7: Schematic view of an inert support coating with different crystalline fractions of a polymer, after the TREF cooling process, reprinted from [12].

Subsequently to the cooling process a temperature increasing program is started in combination with a

solvent flow, in order to elute the fractions, as illustrated in Figure 8. Initially the least crystalline (or

most amorphous fraction) will elute, whereas the highest crystalline fraction will only dissolve at high

temperatures and thus elute last.

Page 20: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 19 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 8: Schematic overview of elution of a TREF fractionation, reprinted from [12].

TREF can be performed on an analytical scale (a-TREF) or preparative scale (p-TREF). For a-TREF [13-22],

online detection is performed during the fractionation, this typically results in a TREF curve, as shown in

Figure 9. On-line detection of analytical TREF can be performed by e.g. an infrared detector, a light

scatter detector, a viscometer or a refractive index detector (see paragraph 0). With p-TREF [15, 23-33]

no detectors are used. The aim is to separate larger amounts of fractions as a function of temperature.

This allows to subsequently performing additional analytical techniques like NMR or DSC. In literature,

initially TREF was performed as analytical technique, however, over the years the use of analytical scale

TREF seems to stabilize, whereas the preparative scale TREF shows an upwards trend. This trend is an

indication for the increased complexity of the samples, and therefore the increased demands on the

analytical techniques, and for that purpose preparative TREF is a very powerful technique in order to

strengthen the power of other techniques like NMR.

Figure 9: Example of a TREF curve (IR intensity plotted versus elution temperature, showing a non-retained, amorphous peak on the left and a retained, crystalline peak on the left),obtained from a copolymer, reprinted from [22].

TREF is an effective method for blend and copolymer separation [7, 15, 34]. The crystallinity of ethylene

copolymers is mainly based on the short chain branching content, TREF chromatograms shows a

separation based on the short chain branching distribution (SCBD), which provides information about

the polymerization mechanism and nature of the catalyst.

For polypropylene (PP), the crystallinity is mainly based on the stereoisomerism. Analysing PP by means

of TREF will therefore give information about the tacticity distribution of the PP. However, due to a lack

Page 21: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 20 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

of reliable calibration curves isotacticity determinations are usually performed by preparative TREF in

combination with NMR [12, 26-28].

As proven by Tso et al. [16], TREF coupled to a series of detectors, offers insight into the heterogeneity

of resin blends by infrared detection, in combination with some molecular weight information from the

viscometer and light scattering detector. With these three detectors coupled in series, the so-called 3D-

TREF is a rather powerful technique for the characterization of polyolefins. However, the main

disadvantages of TREF are the inability to fractionate amorphous samples, the appearance of co-

crystallization and the time consumption.

In an article of Gabriel et al. [21], the comparison of TREF, CRYSTAF and DSC (upcoming chapter), and

showed similar results regarding the SCBD of LLDPE. The main advantage of TREF is the high resolution

and the excellent separation of differently branched components in PO, as shown in Figure 10 [35],

however, Figure 11, originating from [21], shows a superior resolution and separation power for DSC,

compared to TREF. Furthermore it should be noted that the cooling time for TREF was as low as 0.03

K/min, whereas the DSC was cooled at a rate of 0.1K/min. Thus the work of Gabriel et al. seems to show

superior performance of DSC over TREF, unlike what is shown in Figure 10, originating from Vela Estrada

et al. [35], while being less time consuming. This was demonstrated for respect to the comonomer

distribution of LLDPE; however, for the investigation of polyolefin blends the combination of the three

methods is found to be recommendable. Additionally it should also be noted that the principle of

separation in TREF and DSC are different, whereas TREF uses a solvent, DSC does not, while crystallinity

of the polymer can differ based on the presence of solvent.

Figure 10: Overlay of TREF, CRYSTAF and DSC for the SCBD of a Ziegler-Natta-LLDPE, reprinted from [35].

Page 22: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 21 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 11: left, mLLDPE/mHDPE/PP blend analysed by TREF and DSC; right, metallocene mLLDPE by means of TREF and DSC, reprinted from [21].

Overall analytical scale TREF seems to be a well-established but rather slow method for the

determination of short chain branching distribution or tacticity, based on a separation as function of the

crystallization. The separation power and resolution seems to be rather low, but the advantage is the

use of a single solvent, making use of more detectors easier, resulting in more characteristic information

from a single analysis (3D-TREF). Furthermore, the calculation of the composition of a polyolefin blend

by use of a model is more easy for TREF. As a result of the relatively low resolution in combination with

the long analysis time, a more recent technique, crystallization elution fractionation (CEF) will be

preferred over TREF, and therefore it can be expected that the use of analytical scale TREF will slowly

decrease.

Preparative scale TREF on the other hand is a very useful, but also time-consuming and labour intensive,

technique for generating relatively large fractions of the polyolefin, rendering it capable of performing

less sensitive techniques like 13C-NMR, although, with the enhanced sensitivity of cryo-probes used for

NMR, the sample requirements have decreased, which might results in CEF offering sufficient sample

loading and fractionation capabilities. Even the use of on-line CEF-NMR might be introduced; however,

as long a p-TREF will be used, a-TREF will be required for the determination of the elution profile in

order to select the correct fractions.

Overall it can be concluded that TREF is a powerful fractionation technique, but it has a few

disadvantages being:

- Inability to fractionate amorphous samples (>8-10% comonomer); this can also be seen as an

advantage in order to easily quantify the amount of amorphous structure within a polymer

- Co-crystallization can occur

- Long analysis time

Page 23: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 22 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

3.2. Crystallization Elution Fractionation One of the most recent developments in crystallization based separation techniques is crystallization

elution fractionation (CEF), introduced by Monrabal et al.[36, 37]. CEF is a method similar to TREF. While

comparing CEF to TREF the difference is the use of a small flow over the system during the crystallization

phase which results in an initial separation based on the crystallinity. Afterwards for both CEF and TREF

the elution phase is initiated, in case of TREF the sample is still not fractionation whereas with CEF a

“pre-fractionation” has taken place, which is further improved during the elution step where the system

is slowly heated, as explained in the previous chapter discussing TREF. The total process is illustrated in

Figure 12.

Figure 12: Separation principle within TREF (top) and CEF (bottom); CEF adds separation during the crystallization stage, resulting in improved separation in the end, reprinted from [36].

One of the advantages of CEF over TREF and CRYSTAF is the increased analysis speed (as low as 30

minutes per sample), furthermore CEF can be performed on a typical TREF instrument and can thus be

used in combination with viscometry, light scattering, composition and molar mass sensitive detectors.

Another advantage is that the separation can be improved by tuning the experimental conditions

according to the crystallization range of the sample. And since the increased separation of CEF over TREF

is originating from the separation during crystallization, optimizing these parameters has a significant

effect on the end result.

Monrabal et al. [37] also discussed a solution for the reduction of co-crystallization with CEF, by

performing multiple cycles of crystallization and elution, which also led to an increase in resolution apart

from the decrease of co-crystallization.

Suriya et al. [38] compared the performance of CRYSTAF and CEF, and concluded that CEF is more robust

and has less co-crystallization than CRYSTAF. Furthermore CEF requires shorter analysis times for similar

resolutions as achieved within CRYSTAF.

Page 24: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 23 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Overall CEF proves to be a good “upgrade” from the existing TREF methods, resulting in similar or

improved resolution achieved in a shorter analysis time with less co-crystallization. The incapability of

separation of more amorphous samples will still be present, although this can be seen as an advantage

since the quantification of an amorphous phase within a polymer can be performed with ease.

Page 25: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 24 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

3.3. Successive Self-Nucleation and Annealing & Stepwise Crystallization Both stepwise crystallization and successive self-nucleation and annealing (SSA) are techniques usually

performed prior to a Differential Scanning Calorimetry (DSC) measurement resulting in SC-DSC or SSA-

DSC. DSC is a thermal analysis technique in which the difference in heat flow rate to the sample and to a

reference sample is measured, while they are subjected to a controlled temperature program. DSC can

be applied for e.g. the characterization of materials, the analysis of the thermal transitions within a

sample, the determination of oxidative stability, the determination of the purity of substances and to

analyse the influence of a catalyst on a reaction.

There are basically two ways to perform DSC, heat flux DSC and power compensation DSC. Heat flux DSC

uses a single oven to heat both the reference material and the sample material by means of thermal

conduction, as illustrated in Figure 13. During this process the heat delivered towards the sample and

the reference, the heat capacity, is measured.

Figure 13: The furnace of a (disk-type) heat flux DSC; S is sample, R is reference, reprinted from [39].

Unlike heat flux DSC, power compensation DSC consists of two separate furnaces, as illustrated in Figure

14. Both are constantly kept at the same temperature, thus during a change in the sample, the heat

requirement in order to maintain a stable temperature changes. These changes are appearing as

deviations of the DSC curve.

Figure 14: The furnace of a power compensation DSC; S is sample, R is reference, reprinted from [39].

For simple samples the DSC will result in thermograms, the visualization of the analysis result of a DSC

run, with a single peak (Figure 15 a), whereas for (semi-)crystalline polymers multiple melting peaks are

frequently observed.

Page 26: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 25 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 15: a) Single peak thermogram of HDPE, reprinted from internal document; b) overlay of several more complex DSC thermograms, b) reprinted from [40].

The influence of the thermal history of polyethylene on the melting curves has been reported back in

the 60’s by Gray et al. [9]. Based on these findings, the use of a thermal fractionation (temperature

based fractionation performed within the polymer itself) prior to DSC might be a possible alternative to

temperature rising elution fractionation (TREF) [7] or crystallization analyses fractionation (CRYSTAF) [8].

In order to “introduce” thermal history into the polyethylene, several methods have been developed out

of which stepwise crystallization (SC) [14, 41-48] and successive self-nucleation/annealing (SSA) [5, 26,

47, 49-58] are the most common. SSA is widely used for characterization of the microstructure of

polymers, mainly for the VLDPE (very low density polyethylene) and LLDPE (linear low density

polyethylene), nevertheless, in some articles also the characterization of HDPE (high density

polyethylene) [59, 60] and iPP (isotactic polypropylene) [26, 27] are performed by SSA-DSC.

Both stepwise crystallization and successive self-nucleation/annealing are methods for the thermal

fractionation of polymers prior to the recording of a DSC thermogram. Stepwise crystallization is a

method in which the sample is heated above the melting point to subsequently, in steps of

Page 27: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 26 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

approximately 5°C, slowly cool down (up to a few hours per step) the sample. This results in structured

crystallization.

The effect of performing stepwise crystallization on the polymer, prior to the DSC analyses, is illustrated

in Figure 16 (DSC thermogram without SC) and Figure 17 (DSC thermograms after performing SC). It can

be concluded that the SC does significantly enhance the resolution of the m-VLDPE, whereas the effect

for the m-LLDPE is visible, but it seems to be less effective. Furthermore, the stepwise crystallization

method used by Razavi-Nouri et al. was performed with 10 steps (385-340K) for the m-LLDPE and 14

steps (385-320K) for the m-VLDPE, each step consisting of 2 hours of crystallization time. Including

sample handling, warming and cooling periods, it took approximately 24-36 hours per sample to

perform the stepwise crystallization.

The resulting thermogram is a measure for the required melting energy. This can be either a positive

(Figure 15) or a negative (Figure 16 & Figure 17) value. The energy required for melting the sample is an

indication for the quantity of sample that melts at a certain temperature. After performing thermal

fractionation like SC or SSA the sample starts showing several melting peaks. These peaks represent the

different fractions of the sample with their own melting temperature. The intensities represent the

quantities of the fractions within the sample. And the melting temperatures can be calculated back

towards characteristics of that fraction, which is discussed at a later stage in this chapter.

Figure 16: m-VLDPE and m-LLDPE DSC thermograms without step-wise crystallization, reprinted from [42].

Page 28: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 27 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 17: m-VLDPE (left) and m-LLDPE (right) DSC thermograms after performing step-wise crystallization, reprinted from [42].

Successive Self-Nucleation and Annealing is a method which is comparable to stepwise crystallization.

Instead of stepwise cooling down the sample after heating, the sample is cooled down to a low

temperature, resulting in self-nucleation of the polymer. Subsequently it is heated to a lower

temperature, resulting in annealing of the not-melted crystals at the reduced temperature. The main

advantages of applying a SSA fractionation include reduced measurement time, molecular segregation,

enhanced resolution and improved molecular structural information [5]. A comparison graph of the

temperature ramps is shown in Figure 18 (these values are fictive), from which it becomes clear that in

the time SC passed a few cycles SSA performs more than four times as many cycles, resulting in a faster

process. According to Zhang et al. [18] and Arnal et al. [60], SSA is a method designed to fractionate

semi-crystalline polymers by lamellar thicknesses of the crystals and/or by methylene sequence length.

Furthermore, SSA enhances the sensitivity of DSC, by minimizing the extent of co-crystallization

between dissimilar chain segments [59].

Figure 18: Comparison between SSA and SC temperature ramp; starting temperature 160°C, first Ts (self-seeding temperature) 140°C, step size 50°C, isothermal SC 120 minutes, isothermal SSA 5 minutes, ramp SSA heating/cooling 10°C/min; these values are solely for illustrative purposes and not based on experimental results.

Page 29: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 28 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

SSA is a method to subject a polymer to a series of annealing steps. During these steps, the polymer is

fractionated based on the crystallization behaviour, which is dependent on e.g. the type of polymer (for

blends), the short chain branching and the molecular mass. This technique involves partial melting of the

existing crystals followed by recrystallization, while using not-molten crystals as self-nuclei. Subsequent

to the annealing steps, the sample is cooled and a DSC measurement is performed. The resulting

endotherm will show peaks corresponding to the different annealing steps. Figure 19 shows a DSC

thermogram of two polymers before and after performing SSA. The SSA method as applied by Lorenzo

et al. [61], consisted of 13 steps (124°C – 59°C) for the 11U4 and 7 steps (102°C – 72°C) for the HPB, with

heating rates of 5°C/min to -20°C and self-seeding times of 5 minutes, it took on average 30 minutes per

step, resulting in approximately 8 hours for the 11U4 and 6 hours for the HPB polymer to perform the

SSA.

Figure 19: DSC heating scans (10 °C/min) for different polymers, (a) before and (b) after SSA fractionation, reprinted from [61].

In order to actually compare the SC and SSA method, the same samples should be used for the different

methods, which has been performed by Starck et al. in 2002 [46]. In Figure 20, an overlay is given of the

thermograms of two Ziegler-Natta-catalysed ethylene-1-butene copolymers. Both have been

fractionated with SC and SSA, rendering this overlay ideal for the comparison of these two methods.

Page 30: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 29 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 20: DSC thermograms of two ZN-catalysed ethylene-1-butene copolymers, obtained after performed either SC or SSA, reprinted from [46].

While interpreting the data of the EB 5-V (the short thermograms), it is clear that the SSA fractionation

shows a better separation between the peaks. Additionally the SSA-DSC shows up to 11 peaks, whereas

the SC-DSC only shows 7 peaks. Similar results are obtained from the EB1-ZN polymer, for which the

SSA-DSC shows up to 11 peaks, compared to the 7 peaks as detected for the SC-DSC thermogram. The

amount of peaks perfectly corresponds to the number of steps that were performed during the thermal

fractionation. From these results it can be concluded that the use of SSA seems to be a more efficient

thermal fractionation method, resulting in higher resolution thermograms. These results, as achieved by

Starck et al. [46], are in agreement with those of Müller et al. [62]. Additionally, the stepwise

crystallization of the polymers took on average about 18 hours, whereas the successive self-nucleation

and annealing was performed in only 4 hours on average. Thus SSA is significantly faster than SC, while

achieving better resolution [5, 18, 46, 60, 62], as shown in Figure 20.

After performing SSA-DSC, the resulting thermogram could be de-convoluted by using peak fitting

software, as shown in Figure 21. Whereas Figure 21 shows a perfect fit between the experimental and

the peak fit data for both the HPB and 11U4, for the 11U4 a different peak shape is calculated for the

second last peak (right figure; generated peaks (bottom)). Compared to the other peak shapes, it is

unlikely that the peak at 120°C will have such a wide base, especially since the step size during

fractionation was constant, it is expected that the peak width at the base are similar. Performing these

peak fittings should therefore be done with caution.

Page 31: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 30 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 21: SSA-DSC after performing peak fitting on HPB (left) and 11U4 (right) polymers, reprinted from [61].

Hosoda suggested a relationship between the melting temperature of a m-LLDPE copolymer, of ethylene

with 1-hexene or 1-butene, and the degree of short chain branching [63]. Based on the equation by

Hosoda [63], Razavi-Nouri [42] calculated the short chain branching content of a m-VLDPE and a m-

LLDPE. They found a good agreement with the calculations based on the theory of the rubber elasticity.

However, since the melting point of a polymer depends not only on the branching but also on the

composition and the molecular weight, the resulting short chain branching should be interpreted in a

semi-quantitative way. Furthermore quantitative analysis by means of DSC is difficult since the intensity

originates from the combination of the amount of material and the enthalpy of fusion at a particular

temperature [64]. Although, Torabi achieved comparable results between DSC and 13C NMR for the

short chain branching content of 6 different LLDPE’s [65].

Performing successive self-nucleation and annealing, will form lamellae within the polymer, of which a

transmission electron microscopy (TEM) image is shown in Figure 22. The lamellae thickness and

distribution depends on the composition of the polymer, and is related to e.g. the methylene sequence

lengths (MSL).

Figure 22: TEM image of 11U4; the different lamellae are visible as fibres, reprinted from [61].

The Thomson-Gibbs equation is used to calculate the relation between the lamellae (crystal) thickness

and the melting point, as shown in Equation 6, in which Tm0 = equilibrium melting temperature [K]; Tm =

melting point [K]; σ = surface energy [J/m2]; ρ = density [g/m3]; ΔHf = fusion heat for fully crystalline

Page 32: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 31 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

polymer [J/g] and LC = lamellae thickness [m]. When applying the Thomson-Gibbs equation, after

calculating the amount of material melting at a particular temperature, the temperature axis can be

converted into lamellar thickness or into MSL. This is however subject to controversy within the

literature [61]. First of all the equilibrium melting point, the melting point of a 100% crystalline polymer,

which can be calculated by either the Gibbs-Thomson plot or the Hoffman-Weeks method is an

important factor within the Thomson-Gibbs equation. The Gibbs-Thomson plot is a plot of the melting

temperature, acquired by DSC, versus the inverse of crystalline lamellae thickness, acquired by Small-

Angle X-ray Scattering. The Hoffman-Weeks plot is a plot of the relationship between the melting point

and crystallization temperature, as shown in Equation 7, in which Tm0 = equilibrium melting temperature

[K]; Tm = melting point [K]; γ = Thickening coefficient and TC = Crystallization temperature. Creating a

Gibbs-Thomson plot is very time consuming. While the Hoffman-Weeks method seems to face some

difficulties [66, 67]; within literature several values, between 141-145.5 °C [5, 60, 62], have been used as

equilibrium melting point. From which 145.5°C corresponds to infinitely thick linear polyethylene, and

that it is depressed by the introduction of branches in the polyethylene chains [61].

Equation 6

cf

mmm

LH

TTT

00 2

Equation 7

co

mm

TTT

11

Furthermore, for copolymers of ethylene and α-olefins the lamellae thickness calculated via the

Thomson-Gibbs equation do not correspond to the values as measured by TEM. This seems to be related

to the broader melting temperature range of the random copolymers [61, 68].

Although SSA-DSC is a much faster technique than TREF (previous chapter), it only gives a qualitative

description of branching structure and is unable to determine the average content of branches in

VLDPE/LLDPE samples [65], nevertheless, recent developments regarding chip based DSC led to

quantitative results [69]. For isotactic polypropylene, SSA-DSC does enable an accurate characterization

of the tacticity distribution. As proven by Virkkunen et al. [26, 27], the same results are obtained as by 13C NMR and TREF. The results showed that a correlation exists between the lamellar structure

generated in SSA and the crystallizable isotactic sequence length. This however is not applicable to all

fractions; octane soluble at 80°C fraction did not show this correlation [26]. Virkkunen also stated that

the quantitative aspect of SSA-DSC is not significantly affected by changes in the sample composition,

while measuring isotactic polypropylene.

Page 33: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 32 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

4. Interaction Based Fractionation Methods In the upcoming chapters more details will be given about the interaction based separation techniques

temperature gradient interaction chromatography (TGIC) and high temperature high performance liquid

chromatography (HT-HPLC).

4.1. High Temperature High Performance Liquid Chromatography High-performance liquid chromatography (HPLC) is a solvent separation technique which applies high

pressures in packed columns in order to optimize the resolution while reducing the analysis time. HPLC

is used to separate compounds in a mixture on basis of polarity (e.g. the electron affinity between atoms

within a molecule). The chromatographic setup consists of a mobile phase (a solvent that is constantly

pumped through the system), a stationary phase (present within in column) and a detector. The

principle behind HPLC is the interaction of the compound of interest with the stationary phase present

inside the column and the solubility of the compound in the mobile phase which is passed through the

column. In the past, HPLC was solely performed at ambient temperature (uncontrolled) or slightly

elevated temperatures (typically around 40°C). Especially in the last decade, strong progress was made

in the development, both instrumentation and methodology, of this method for polyolefin

characterization.

Performing polyolefin separations around ambient temperatures does not seem to be a viable principle

as a result of the crystalline nature of the polyolefins. Most polyolefins can only be dissolved at high

temperatures (>100°C). High boiling point solvents are thus required. The choice of solvent is therefore

limited.

Since the separation is based on the interaction between the analytes with the stationary phase, within

the column, in combination with the mobility through (diffusion), and solubility of, the analyte in the

mobile phase, many variables can be changed in order to alter the selectivity of the system or optimize

the separation of the different analytes. Furthermore, within HPLC several separation mechanisms can

be applied, e.g. steric exclusion (size based separation), adsorption (chemical interaction based

separation), partitioning (phase separation), solubility (precipitation chromatography) and ion-effects

(adjust interaction with stationary phase by ionizing analytes), in order to achieve optimal separation.

One of the main advantages of using HT-HPLC over the other techniques is the great flexibility of the

technique.

As one of the pioneers, Macko et al. [70] performed a study towards the elution behaviour of

polyethylene in isocratic HT-HPLC while using different polar mobile phases on a non-polar stationary

phase. Based on the used mobile phases, different elution behaviours were observed, from SEC

separation (elution without any (enthalpathic) interaction or adsorption) to fully retained peaks. Using

ethylene glycol monobutylether as mobile phase enabled the separation of PE (retained) from PP

(eluted in SEC mode). However, the resulting chromatogram, as shown in Figure 23, did not look

satisfying due to a lack of resolution. Furthermore it was not always possible (dependent on the

polymer/stationary phase combination) to desorb the adsorbed PE and PP from the stationary phase.

Page 34: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 33 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 23: Separation of PE and PP blends by isocratic HT-HPLC, Macko et al., reprinted from [70]; blends of PP and PE with 36 and 34 (solid line), 57 and 66 (dotted line) and 438 and 500 kg/mol (dashed line), respectively.

Up to the year 2005 solvent gradients had not been used for the separation of polyolefin by means of

HT-HPLC, simply because of a lack of knowledge and equipment. Heinz et al. [71] were the first to

publish experimental results based on gradient HT-HPLC of a polyethylene-polypropylene blend. Their

experiments led to a successful separation of PE (normal retention) and PP (eluted in SEC mode) by

using a mobile phase gradient of ethylene glycol monobutylether (EGMBE) and 1,2,4-trichlorobenzene

(TCB) in combination with a silica gel (nucleosil 500) stationary phase. Compared to the chromatograms

in Figure 23, as achieved by Macko et al. [70], the results of Heinz et al., using a gradient, showed a

significantly improved resolution for the separation of PE from PP, as shown in Figure 24. For PE a small

influence of the molar mass on the retention has been observed. It should be noted however that very

low molar mass PE (e.g. 1.1 kg/mol) is soluble in EGMBE and thus elutes in the SEC mode together with

the PP, showing no retention.

Figure 24: Separation of a PE, 126kg/mol, (right peak) and PP, 305 kg/mol, (left peak) blend by gradient HT-HPLC, Heinz et al., reprinted from [71].

Macko et al. [72] achieved a separation of linear-PE and PP from PE-PP-blends by using selective

adsorption of PE and PP on zeolites in isocratic HT-HPLC, nevertheless, the retained compound was

difficult, or even impossible to desorb. A few years later Macko et al. [73] were the first to publish a

Page 35: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 34 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

polyolefin separation based on tacticity, by means of HT-HPLC, using porous carbon-based material as

stationary phase. With 1-decanol as a mobile phase (start of gradient), linear-PE, syndiotactic-PP and

atactic-PP are adsorbed, while isotactic-PP is fully eluted without adsorption. After the isocratic elution

of isotactic-PP the retained polymer fractions are eluted by a gradient from 1-decanol to TCB, which is

illustrated in Figure 25.

Figure 25: Separation according to tacticity, by Macko et al., reprinted from [73]; four peaks detected, isotactic-PP Mw=200kg/mol, atactic-PP Mw=315kg/mol, syndiotactic-PP mw=196kg/mol, linear-PE 5 samples from 14-260kg/mol (seen as one peak), respectively.

In 2009, Macko et al. [74] were also the first to publish the separation of polyolefin chains based on the

short chain branches (SCBD), they succeeded in the separation of propene/α-olefin and ethylene/α-

olefin copolymers, by means of solvent-gradient HT-HPLC. Whereas isotactic-pp/α-olefin (≤C10 in side

chain) eluted in SEC mode, isotactic-pp/α-olefin (≥C11 in side chain) showed (slightly) increased retention

with (significantly) increasing concentration of co-monomer, rendering the system (slightly) suitable for

separation of these copolymers according to the chemical composition, as illustrated in Figure 26.

However, the opposite was observed for ethylene/1-hexene copolymer, in that case the retention

reduced with increased co-monomer concentration, as shown in Figure 27, most-likely caused by a

reduction in length of the continuous methylene sequences which is expected to adsorb to the graphite.

Furthermore isotactic-PP was separated from syndiotactic-PP, using this setup of a Hypercarb® packed

column with porous graphite, while using a linear gradient from 1-decanol to TCB, as already presented

in [73].

Page 36: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 35 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 26: Short Chain branching separation of propene/1-octadecene copolymer according to copolymer content, showing increased retention with increased co-monomer content, by Macko et al., reprinted from [74].

Figure 27: Short chain branching separation of ethylene/1-hexene copolymer according to copolymer content, showing reduced retention with increased co-monomer content, by Macko et al., reprinted from [74].

Chitta et al. [75] studied the elution behaviour of polyethylene and polypropylene standards on different

carbon sorbents using several solvents. These combinations led to new sorbent/solvent systems which

have different selectivity compared to the setups of Macko et al. [70, 73] and Heinz et al. [71]. Chitta et

al. managed to separate PP from PE, by eluting PP (isotactic-PP, atactic-PP and syndiotactic-PP) in SEC

mode and adsorb PE (including the <20kg/mol fraction) to subsequently elute the PE during a gradient.

The used chromatographic system contained ZirChrom-CARB sorbent, with a gradient from 2-ethyl-1-

hexanol to TCB at an isothermal temperature of 160°C; the resulting chromatogram is shown in Figure

28.

Page 37: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 36 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 28: Separation of PP and PE as achieved by Chitta et al., reprinted from [75].

Whereas the setup of Macko et al. [74], Figure 25, is capable of separating PP based on tacticity, the

setup of Chitta et al. [75], Figure 28, showed a slight influence of the molecular mass on the separation

of PE, up to 60kD. A combination of these two systems would be a good setup for complete

characterization of PE, PP and PE-PP blends, although this is not viable since both the mobile phase and

the stationary phase was different between the two experiments.

With the same setup as used for the separation of PP and PE in Figure 28, Chitta et al. [75] also managed

to separate ethylene/1-hexene copolymers according to their chemical composition, as previously

performed by Macko et al. [74] (Figure 27). The resulting chromatogram is shown in Figure 29. Whereas

Macko et al. [74] (Figure 27) suffered from split peaks, Chitta et al. [75] (Figure 29) do not seem to have

to cope with that problem. Furthermore, the chromatogram of Chitta et al. seems to show full

adsorption of the high co-monomer content copolymers.

Figure 29: Short chain branching separation of ethylene/1-hexene copolymer according to co-monomer content by Chitta et al., reprinted from [75].

At this moment HT-HPLC starts to become a good alternative for conventional methods like TREF and

CRYSTAF, since HT-HPLC is performed in a shorter time span. In the chromatograms as shown above, all

fractions eluted within 13 minutes, which is, compared to the time required for TREF, a significant gain

Page 38: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 37 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

in efficiency. Furthermore, it can separate amorphous fractions >8% co-monomer, unlike TREF. On the

other hand the separation within CEF and TREF is mainly based on crystallinity rending the techniques

complementary to HT-HPLC. The main disadvantage of HT-HPLC is the use of a solvent gradient, which

limits the selection of possible detectors; only the ELSD is a viable detector in combination with HT-

HPLC. Furthermore the reliability of the hardware is currently below industry standards. Apart from that

it is difficult to perform method development, because the solvent combinations are limited. And since

HT-HPLC works well for polymers that show adsorption at a stationary phase in a good solvent and elute

in another good solvent, finding the right solvents with completely different behaviour towards

adsorption of the polymers on stationary phases also requires some more research in order to simplify

the method development.

Page 39: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 38 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

4.2. Temperature Gradient Interaction Chromatography Temperature Gradient Interaction Chromatography (TGIC) is a liquid chromatography method which is

performed under isocratic mobile phase conditions whereas separation is controlled by the use of a

temperature gradient. The basic principle of TGIC is the compound-dependent temperature influence on

the interaction between the analyte and the stationary phase. Partition coefficients, the affinity

between the compound of interest and the stationary phase versus mobile phase, of the analyte are

controlled by the variation of the temperature during the elution. For most compounds, the retention

factor, which quantifies the degree of interaction between the compound and the stationary phase,

decreases at an increased temperature, resulting in adsorbed analytes to be desorbed and eluted,

although some compounds show the opposite effect [76]; illustrated in Figure 30.

Figure 30: influence of temperature (K-1) on the retention factor (ln k), left the retention is reduced at increased temperatures, reprinted from [77], right the retention increases with the temperature, reprinted from [78].

The main advantage of TGIC over (HT-)HPLC is the use of a single solvent, which not only enlarges the

selection of possible detectors, e.g. the use of a refractive index detector, viscometer and IR detector,

but also improves the performance of for instance the ELSD detector. Furthermore, TGIC offers an

increased loading capacity compared to HT-HPLC, rendering it more capable of performing polymer

fractionations on a preparative scale for off-line analysis by means of other methods, e.g. infrared or

NMR. A downside of the use of TGIC is the narrow optimum separation conditions for a specific

polymer, requiring more optimization of these settings [79-81].

Initial TGIC experiments, as performed by Lee, et al. [80], showed molecular weight separation of

polystyrene which outperformed the resolution as achieved by SEC. These TGIC experiments were

afterwards performed again on polystyrene, Poly(methyl methacrylate), polyisoprene and poly(vinyl

chloride) by the same group at the Pohang University of Science and Technology, between 1998 and

2006 [79, 81-88]. However, these experiments were all conducted at temperatures up to 60°C [79],

which is not sufficient for the characterization of polyolefins.

Further developments of TGIC led to methods which achieved separation based on chemical

composition. For these methods, just like with HT-HPLC as discussed previously, one fraction elutes in

SEC mode, whereas the other fraction is initially adsorbed, to subsequently elute during the

temperature gradient. An example is given in Figure 31, in which the work of Lee, et al. [89], the SEC

Page 40: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 39 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

separation of PI (polyisoprene) and the TGIC separation of PS (polystyrene) in a PI-PS mixture (a), and

vice versa (b), is presented.

Figure 31: TGIC chromatograms of PI-PS mixtures; (a) NP-TGIC, PI separated in SEC mode, PS separated in TGIC mode; (b) RP-TGIC, elution reveres since PI is adsorbed in reversed phase, reprinted from [89].

As shown by Teramachi et al. [90], TGIC can be used for analytes that have decreased retention at

increased temperature and analytes that have increased retention at increased temperatures by

adjusting the direction of the gradient, increasing versus decreasing temperature in time. A positive

(increasing) temperature gradient was effective in systems consisting of NH2 modified silica with

tetrahydrofuran (THF) and n-hexane as eluent-mixture, whereas a negative temperature gradient was

effective for a system in which the THF is replaced by chloroform. However, poly(ethylene oxide) (PEO)

is one of the only polymers which show increased retention at increased temperatures.

Whereas the development of TGIC capabilities was being further studied for years, it took until 2011 for

the first publication on High Temperature Thermal Gradient Interaction Chromatography (HT-TGIC) on

the separation of polyolefins to appear in Macromolecules. In this paper of Cong et al. [91] described

HT-TGIC as a new characterization technique to quantify co-monomer distributions of which the

performance was proven by the characterization of a series of ethylene octene copolymers with octene

co-monomer contents between 0% and 50.7%.

As a result of the lack of solubility of polyolefins at ambient temperature, elevated temperatures are

required for dissolving these polymers. The HT-TGIC experimental setup, using HYPERCARB stationary

phase in combination with 600PPM butylated hydroxytoluene (BHT) in ortho-dichlororbenzene (ODCB)

as mobile phase, used is shown in Figure 32, reprinted from [91], in which, TS = stabilization

temperature; FC = flow rate of pump during cooling; FE = flow rate of pump during elution; FF = flow rate

of pump during flushing; tloop = stabilization time; tcolumn = precooling time; RC = cooling rate; TC = final

temperature of cooling process; tC = post-cooling time; tE = soluble fraction elution time; RE = elution

rate and TE = final elution temperature.

Page 41: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 40 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 32: Temperature profile of HT-TGIC, reprinted from [91]. TS = stabilization temperature; FC = flow rate of pump during cooling; FE = flow rate of pump during elution; FF = flow rate of pump during flushing; tloop = stabilization time; tcolumn = precooling time; RC = cooling rate; TC = final temperature of cooling process; tC = post-cooling time; tE = soluble fraction elution time; RE = elution rate and TE = final elution temperature.

In Figure 33 an overlay of a series of chromatograms for the different copolymers (ethene + 0% - 50.7%

octene) is presented. Whereas TREF or CRYSTAF are normally incapable of fractionating copolymers with

>10% co-monomer, as a result of the increased amorphousness, the TGIC method clearly shows

interaction up to 32.5% mol% of octene in a ethene-octene copolymer.

The peak at a temperature of 0°C, 50.7mol% octene, eluted isothermally as a sharp peak, whereas the

polyethylene and the 1.33mol% octene eluted at high temperatures as sharp peaks. The signals in

between seem to show more broadening, although, it looks like the broadening does decrease with

increasing elution temperatures. Observing the shape, a clear tail behind the apex, might point towards

a molecular weight distribution of the different fractions, or secondary interactions between the analyte

and the stationary phase; nevertheless, the actual cause remains unknown.

Page 42: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 41 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 33: Chromatographic overlay of different ethene/octene copolymers, varying from 0% to 50.7% octene, showing a clear co-monomer based separation; reprinted from [91].

Furthermore, the TGIC method seems to show a linear relationship between the elution temperature

and the copolymer concentration, as shown in Figure 34. Apart from some random error in the

measurement points, the 32.5 mol% and 50.7 mol% octene seem to deviate from the line. For the 32.5

mol% that might be caused by the asymmetric peak shape, resulting in the top of the peak eluting at a

lower temperature, whereas the 50.7 mol% shows no retention, which can already be expected starting

from approximately 48.8 mol%, based on the calibration line.

Figure 34: Linear relationship between octene content in ethene/octene copolymers and elution temperature for HT-TGIC, reprinted from [91].

This dependence of the elution temperature on the copolymer content has been shown by performing

preparative HT-TGIC in combination with 13C-NMR, which is presented in Figure 35. These NMR results

are, however, a bit low regarding the octene mol%. The 96°C-102°C fraction contains between 8.5 and

10 mol% of octene, whereas in the chromatogram, as presented in Figure 33, from 96°C-102°C the

tailing of 19 mol% and 21.7 mol% octene copolymer elute, together with the fronting of the 13.88 mol%.

Page 43: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 42 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

This would mean that either the mol% octene should be between 13.88 mol% and 19 mol% instead of

8.5%-10%, or that the NMR results are not properly calculated. This does not count for the 126°C-132°C

fraction, whereas it contains between de 5.5 and 6.5 mol% according to the NMR results, the

chromatogram shows the tailing of the 8.52 mol% and the fronting of the 3.99 mol%. Therefore it might

also point towards broad copolymer distributions for the different copolymers analysed, which would

also explain the wide peaks as observed in the chromatogram.

Figure 35: Left, fractions as collection with HT-TGIC; right, octene mol% in ethene/octene coplymers for the different fractions as measured by 13C-NMR; reprinted from [91].

Another aspect shown by Cong et al. is the interaction between the graphite of the HYPERCARB column

and the polyolefin. Whereas TREF or CRYSTAF show elution temperatures for the polyethylene around

100°C and DSC show a melting temperature around 135°C, the elution temperature on the HYPERCARB

column using the TGIC method, is around 150°C. With some simple experiments in which the elution

and stabilization (starting) temperatures have been varied, Cong et al. proved a clear interaction

between the HYPERCARB and the polyethylene for temperatures below 140°C [91], which is illustrated

in Figure 36. With a starting temperature of 140°C no interaction occurred, and the polymer eluted

directly (red sharp peak). When decreasing the temperature to 135°C, partial interaction is observed, a

small fraction still eluted instantly (green line, first peak) whereas the main part shows interaction and

eluted around 145°C. Decreasing the temperature further, to 130°C or below, full interaction is observed

and the whole sample elutes after the 140°C which was previously enough to elute the sample. This

points towards the presence of interaction.

Page 44: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 43 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 36: Influence of stabilization temperature on the final elution temperature and thus the interaction with the stationary phase, reprinted from [91]. At a temperature of 140°C no interaction occurs between the polyethylene and column is observed, whereas temperature below 130°C show a clear interaction, that even requires increased temperatures for elution. The additional peak in the green chromatogram is related from a low molecular weight fraction of the polyethylene.

An important advantage of TGIC is the use of a single solvent, compared to a solvent gradient in HT-

HPLC, enlarging the selection of possible detectors and enabling multiple detection modes in series, e.g.

a coupled infrared - viscometer – light scattering, for the comprehensive determination of

microstructures [91]. Furthermore, TGIC is applicable towards a larger range of co-monomer content,

compared to e.g. TREF, and due to the absence of crystallization, co-crystallization is no longer an issue.

Overall it can be concluded that (HT-)TGIC might become a very promising technique for the

characterization of polyolefins based on the chemical composition distribution. However, since only a

few publications are available about TGIC and only a single publication about HT-TGIC, which is required

for the analysis of polyolefins, developments are still in an early stage. The results, as presented by Cong

et al., are promising, although optimizations have to be performed on for instance peak shapes and

separation efficiencies. Additionally, the technique should be applied to other copolymers and blends,

and different solvent/column setups should be tested, in order to verify the finds.

Page 45: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 44 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

5. Two dimensional fractionations Molecular characterization of synthetic polymers used to be dominated by SEC. However, the downside

of SEC is that several molecular characteristics determine the behaviour of the analyte, since the sizes of

polymers in solution highly depends on e.g. molar mass, chemical structure and physical architecture.

An alternative would therefore be the use of liquid chromatography (LC) in which the separation is

based on a single molecular characteristic, which is strongly enhanced, while suppressing the other

characteristics in order to neglect the influences [92]. Such suppression or enhancement can be

achieved within LC separation by applying different separation mechanisms such as exclusion,

adsorption, partitioning, solubility or ion-effects. As Giddings [93] stated, for each dimension present

within a sample, an analytical separation dimension is required within the separation system employed,

in order to achieve optimal separation efficiency. Combining these statements, quickly leads to the

choice for multiple dimension fractionations, as first applied by Balke [94].

In literature, many articles about 2D (polymer) separations can be found [11, 92, 95-102], including

some reviews [11, 92, 100, 102]; nevertheless, the amount of articles discussing 2D separations of

polyolefins is rather limited. Whereas TREFxSEC was already performing during the late 20th and early

21st century [103], and off-line TREFxSSA is discussed in the review of Müller et al. [5], 2D LC separations

of polyolefin were not reported until 2010, by Roy et al. [104, 105].

An example of the 2D separation, is the work by Ginzburg et al. [105]. They used HT-HPLCxSEC as

analytical setup, to separate polyolefins independent of their crystallizability. As an example, the

separation of a blend of PE, ethylene-vinyl acetate (EVA) and poly(vinyl acetate) (PVAc), by HT-HPLCxSEC

and TREFxSEC is shown in Figure 37.

Figure 37: HT-HPLC x SEC contour plot (left) of a blend of PE, EVA (20 mol% and 57 mol% VA) and PVAc, compared to TREFxSEC (right) on the same sample; it can be concluded that the separation with HPLC is significantly better than TREF although TREF might give different information since it is complementary; reprinted from [105].

Unlike 2D separations in gas chromatography, in which the additional separation power is solely used to

physically separate peaks, 2D separations in polymer analyses can be performed for different purposes.

As shown in Figure 37, the added dimension is not used for the actual separation of the compounds, but

Page 46: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 45 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

solely for the determination of the MWD by means of a SEC separation. Therefore this additional

separation dimension seems to be used as a method of identification, rather than separation.

The method as applied by Ginzburg et al. separates the polymers regarding polarity in the first

dimension, whereas in the second dimension a molecular mass separation was performed, in

approximately 5 hours. Apart from the improved separation compared to TREFxSEC, as shown in Figure

37, another advantage of HT-HPLCxSEC versus TREFxSEC is the absence of the co-crystallization effects,

which might be present in TREF. Furthermore HT-HPLCxSEC is capable of separating semi-crystalline and

amorphous polyolefins, whereas TREFxSEC is not capable of separating amorphous (>8mol% co-

monomer) polyolefins and thus ideal for determining the amorphous content.

Roy et al. [104] analysed polyolefins by means of HT-HPLCxSEC resulting in a two dimensional contour

plot of the composition versus the molecular weight. In Figure 38 the influence of the octene content in

ethylene octene copolymers is represented in 2D contour plots. From these plots it is clearly visible that

the elution volume of the LC decreases with an increasing octene content. Even the quantification was

proven to be quite accurate, with a 50/50 weight% blend sample showing two peaks of 1975 and 2067

area respectively, an error of <5%, which might be related to integration errors, as a result of the poor

peak shape, shown in Figure 39.

Figure 38: 2D contour plot of a series of ethylene octene copolymers with varying octene concentrations, reprinted from [104]. In the plots it is clearly visible how the peak shifts towards the left, at increased octene content.

Page 47: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 46 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Figure 39: Reconstructed 1D chromatogram for a 50/50 weight % blend sample, reprinted from [104].

In 2010, Ginzburg et al. [106] presented a method for the characterization of polyolefins by means of so-

called comprehensive high-temperature two-dimensional liquid chromatogram (HT 2D-LC), consisting of

HT-HPLCxSEC. With the system operating at 160°C, using a gradient from 1-decanol to 1,2,4-TCB on a

Hypercarb® column enabled the separation of polyolefins according to their chemical composition,

followed by a molecular mass separation performed on the SEC column. Analysing a blend of iPP (1.1

and 60 kg/mol), aPP (211 kg/mol), sPP (196 kg/mol) and PE (1.18 and 22 kg/mol) showed an influence of

the molecular mass of the iPP 60kg/mol fraction on the solubility in decanol, since a higher molecular

mass fraction of the iPP 60kg/mol only eluted after the gradient towards TCB was initiated. This partial

adsorption of the higher molecular mass fraction of the iPP 60kg/mol might be related to the quality

difference of the 1-decanol, since Macko et al. [73, 74], did not see this adsorption or possible

differences in tacticity, which might be confirmed by TREFxSEC. The resulting chromatogram is

presented in Figure 40.

Figure 40: Contour plot (HT 2D-LC) of a blend of iPP (1.1 and 60 kg/mol), aPP (211 kg/mol), sPP (196 kg/mol) and PE (1.18 and 22 kg/mol), reprinted from [106].

Furthermore, Figure 41, shown below demonstrates the real power of HT 2D-LC (HT-HPLCxSEC), since a

model mixture of 7 components, being PE (2x), iPP (2x), sPP, aPP and an E/P copolymer show full

Page 48: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 47 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

separation. This means that a copolymer of interest can be physically separated from a blend of

different polymers. Comparison of these results to TREFxSEC [106], showed that the HT-HPLCxSEC setup

is independent of the sample crystallinity.

Figure 41: Contour plot (HT 2D-LC) of a blend of iPP (1.1 and 60 kg/mol), aPP (211 kg/mol), sPP (196 kg/mol), PE (1.18 and 126 kg/mol) and E/P-copolymer (81.3 weight% ethylene), reprinted from [106].

The separation of iPP, PE and E/P copolymer, was repeated by Lee et al. in 2011 [107]. Nevertheless, the

results as presented by Ginzburg et al. [106] in 2010, seemed to be more impressive.

Cheruthazhekatt et al. [108] published a multidimensional analysis of the composition of an impact PP

ethylene-propylene copolymer (EPC), by using a combination of TREF, SEC-FTIR-HPer DSC, HT-HPLC and

HT 2D-LC. After performing a preparative TREF fractionation, HT-SEC-FTIR proves information regarding

the chemical composition and crystallinity, as a function of the molar mass and HT-SEC-DSC yields the

melting and crystallization behaviour related to the chemical heterogeneity of the copolymer, from the

different fractions. HT-HPLC shows a separation between the homopolymers (e.g. PE) and the

copolymer. Furthermore, HT 2D-LC (HT-HPLCxSEC) was used to analyse the TREF fractions according to

chemical composition versus the molar mass. Analysis of the bulk sample (non-stabilized impact

polypropylene copolymer with 10.5 mol% ethylene) by these techniques does not provide in-depth

information regarding the chemical composition and thermal behaviour; therefore a preparative TREF

fractionation was performed prior to the analyses. The results pointed out that the techniques used,

show some overlap, but overall they are complementary to each other, however, performing such an

extended research is of course highly time and resource consuming.

Overall it can be concluded that 2D methods show to be very promising for the separation of complex

molecules. Although, as stated by Lee et al. [107], there are some inconsistencies between the

molecular weight poly-dispersities determined by SECxHT-HPLC and those determined solely by SEC.

This shows that the newly developed technique, it was only introduced in 2010 [104, 105], still seems to

show some teething problems.

Page 49: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 48 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

6. Outlook SSA-DSC seems to be superior to SC-DSC regarding both the analysis time and the achieved efficiency, it

can therefore be expected that SSA might become the main thermal fractionation prior to DSC analyses.

This growth might be even strengthened by the development of autonomous equipment being capable

of performing these techniques without requiring human interaction. Especially for the characterization

of polypropylene regarding its tacticity distribution, SSA-DSC seems to be a reliable technique, for most

fractions, as shown by Virkkunen et al. [26, 27]. However, for LDPE SSA-DSC only seems to give

qualitative results for the branching structure, and is thus not able to determine the average branching

content [65].

Furthermore on the side of SSA, the DSC analysis is currently being improved, regarding sample

requirements and ramping speeds. A recent publication of Mathot et al. [69], shows scan speeds up to

20000°C/s for heating and cooling empty cells, whereas speeds of 1000°C/s have been applied on actual

samples, of which the results were even reported as quantitative.

Currently analytical scale TREF seems to be a well-established but rather slow method for the

determination of short chain branching distribution or tacticity, based on a separation as function of the

crystallization. The separation power and resolution seems to be a rather low, in combination with the

long analysis time, therefore other techniques seem to be preferred. CEF, which is an improved version

of TREF (faster, higher resolution and less co-crystallization), seems to be a viable replacement for TREF

since it gives improved separation in a reduced timespan. Furthermore, techniques like TGIC or HT-HPLC

might be used for some of the applications on which TREF is currently performed, although TGIC and HT-

HPLC can also be complementary to each other [108]. Therefore it can be expected that the use TREF for

analytical scale characterization will slowly decrease.

Preparative scale TREF on the other hand is a very useful, but also time-consuming and labour intensive,

technique for generating relatively large fractions of the polyolefin, rendering it capable of performing

less sensitive techniques like 13C-NMR, however, with the enhanced sensitivity of cryo-probes used for

NMR, the sample requirements have decreased, and therefore CEF, which will most likely become

capable of fractionating a sample at analytical scale, seems to become a serious threat for preparative

TREF. Nevertheless, preparative TREF will still play an important role in the characterization of complex

polyolefins for the coming years.

On the other hand the interaction based techniques like HT-HPLC and TGIC are expected to become

commonly implemented techniques. The main advantage of TGIC is the use of a single solvent,

compared to a solvent gradient in HT-HPLC, enlarging the selection of possible detectors and enabling

multiple detection modes in series, e.g. a coupled infrared - viscometer – light scattering, for the

comprehensive determination of microstructures [91]. Furthermore, TGIC is, just like HT-HPLC,

applicable towards a larger range of co-monomer content, compared to e.g. CEF, and as a result of the

high temperature, co-crystallization is no longer an issue. The main disadvantage of HT-HPLC is the use

of a solvent gradient, which limits the selection of possible detectors. Eventhough HT-HPLC might be

Page 50: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 49 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

preferred at the stage, for both techniques, developments are still in a rather early stage. Regarding

TGIC some optimizations have to be performed on for instance peak shapes and separation efficiencies;

furthermore it should be applied to a broader range of polyolefins.

Whether a choice will be made between HT-HPLC (multiple solvents, single temperature) or TGIC (single

solvent, multiple temperatures) mainly depends on the selection of detectors and the application for

which it will be used, in combination with future developments in the analytical world.

Regarding 2D separations, HT-SECxHT-HPLC shows to be a very promising technique for the separation

of complex molecules, as demonstrated by Ginzburg et al. [105, 106], especially when compared to

TREFxSEC. However, as stated by Lee et al. [107], there are some inconsistence between the molecular

weight polydispersity determined by SECxHT-HPLC and those determined solely by SEC. This shows that

the newly developed technique, it was only introduced in 2010 [104, 105], still seems to show some

teething problems; which will most likely be resolved in the upcoming years, making this an extremely

powerful techniques for the complex polyolefin characterizations.

Another option would be the combination of HT-TGIC and HT-HPLC by simply combining a solvent and

temperature gradient in a single run, this might even further enhance the separation capability, and it

does not require additional hardware.

Based on the complementarity of CEF and TGIC or HT-HPLC, 2D setups will arise. Especially CEFxTGIC

might become a very strong 2D technique in the future, due to the possibility of simultaneously utilizing

a series of detectors.

Overall the developments of the techniques will go into miniaturizing, e.g. chip based separation [69,

109, 110], enabling the use of parallel characterizations, in a way comparable to the loop storage

method of LC-NMR. In this technique, the fast LC separation is automatically fractionated into different

sample loops, which can subsequently be analysed by the relatively slow NMR technique. However, the

downside to miniaturisation is peak broadening.

Towards the polymer fractionation it might be an option to connect a solvent based technique, e.g. HT-

HPLC or HT-TGIC, to a series of chips, on which fractions are collected. These chips can subsequently run

a SSA-DSC in parallel, resulting in highly valuable information in a rather short time span.

Page 51: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 50 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

7. Conclusion & Recommendation Based on the current developments, SSA-DSC, HT-HPLC and CEF are the most promising techniques.

Eventhough HT-HPLC, as interaction based technique, seems to be limited regarding detection; it is still

preferred over the TGIC method, which requires further development.

SSA-DSC is can be seen as a fast technique in order to gain a first impression on the molecular structure

of the polymer. It can therefore be a method which is used for deciding whether HT-HPLC is used, or

even complementary, performing SSA-DSC on fractions collected from the HT-HPLC, as a result of the

low sample requirements of the HT-HPLC.

Analytical TREF and CRYSTAF can be seen as techniques sliding towards extinction, as a result of the

introduction of CEF. Based on the different separation techniques, crystallization versus interaction, HT-

HPLC and CEF are complementary to each other [108], which can be utilized by 2D setups. Whereas

preparative TREF will still be performed in the near future order to fractionate co-polymers prior to e.g. 13C-NMR characterizations.

Whether CEF, HT-HPLC, SSA-DSC or (on-line coupled) combinations of these techniques will be used,

depends mainly on the analytical questions that have to be answered.

Table 2, shown on page - 51 -, gives a summary of the techniques as discussed within this report.

Page 52: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 51 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Table 2: Overview of different techniques discussed within this report.

SSA TREF CEF HT-HPLC TGIC

Crystallization X X X

Chemical Composition X X

Branching (Copol) X X X X X

Tacticity (PP) X X X X X

Chemical Composition X X X X X

Speed Average Very Slow Fast Very Fast Fast

Solvent None Single Single Multiple Single

Detector None Multiple Multiple Single Multiple

Resolution High Average Average High Low-Average

Scale Analytical Analytical +

Preparative

Analytical

(+ Preparative?)

Analytical Analytical

Selectivity Crystallinity Crystallinity Crystallinity Full selectivity Crystallinity +

some Amorphous

Required sample Very low Average Average Low Low

Known issues Quantification Co-crystallization Minimal Co-

crystallization

Under

development

Beginning of

development

Complex

calibration

Selectivity

Amorphous

fraction

Selectivity

Amorphous

fraction

Detector

compatibility

Peak shapes

Solubility /

Adsorption

Separation

efficiencies

Future expectations Faster Replaced by

other technique

Preparative scale Temperature

gradient

Improved

understanding

Improved

quantification

Replacement of

TREF

One of the main

techniques

One of the main

techniques

Preparative scale

Separation principle

Outcome

Characteristics

Page 53: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 52 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

Literature references 1. Kaminsky, W., Macromolecular Chemistry and Physics, 2008. 209: p. 459-466. 2. De Rosa, C., et al., Structure-property correlations in polypropylene from metallocene catalysts:

stereodefective, regioregular isotactic polypropylene. Journal of American Chemical Society, 2004. 51: p. 17040-17049.

3. De Rosa, C., F. Auriemma, and L. Resconi, Influence of chain microstructure on the crystalization kinectics of metallocene-made isotactic polypropylene. Macromolecules, 2005. 24: p. 10080-10088.

4. De Rosa, C., et al., Crystallization behavior of isotactic propylene-ethylene and propylene-butene copolymers: effect of comonomers versus stereodefects on crystallization properties of siotactic polypropylene. Macromolecules, 2007. 18: p. 6600-6616.

5. Müller, A.J. and M.L. Arnal, Thermal fractionation of polymers. Progress in Polymer Science, 2005. 30(5): p. 559-603.

6. Giddings, J.C., New separation concept based on a coupling of concentration and flow non-uniformities. Journal of Separation Sciences, 1966. 1(1): p. 123-125.

7. Wild, L., Temperature Rising Elution Fractionation. Advances in Polymer Science, 1991. 98. 8. Monrabal, B., Crystallization analysis fractionation: A new technique for the analysis of

branching distribution in polyolefins. Journal of Applied Polymer Science, 1994. 52(4): p. 491-499.

9. Gray, A.P. and K. Casey, Thermal analysis and the influence of thermal history on polymer fusion curves. Polymer Letters, 1964. 2(4): p. 381-388.

10. Varga, J., J. Menczel, and A. Solti, Memory effect of low-density polyethylene crystallized in a stepwise manner. Journal of thermal analysis, 1976. 10: p. 433-440.

11. Philipsen, H.J.A., Determination of chemical composition distributions in synthetic polymers. Journal of Chromatography A, 2004. 1037(1–2): p. 329-350.

12. Xu, J. and L. Feng, Application of temperature rising elution fractionation in polyolefins. European Polymer Journal, 2000. 36(5): p. 867-878.

13. Watkins, J.J., et al., Fractionation of high density polyethylene in propane by isothermal pressure profiling and isobaric temperature profiling. The Journal of Supercritical Fluids, 1991. 4(1): p. 24-31.

14. Karoglanian, S.A. and I.R. Harrison, A comparative study of ULDPE utilizing DSC and TREF. Thermochimica Acta, 1992. 212(0): p. 143-149.

15. Karoglanian, S.A. and I.R. Harrison, The similarity of compositional distribution information generated by DSC and TREF. Thermochimica Acta, 1996. 288(1–2): p. 239-245.

16. Tso, C.C. and P.J. DesLauriers, Comparison of methods for characterizing comonomer composition in ethylene 1-olefin copolymers: 3D-TREF vs. SEC-FTIR. Polymer, 2004. 45(8): p. 2657-2663.

17. Wang, C., et al., Microstructures of a highly short-chain branched polyethylene. Polymer, 2001. 42(4): p. 1733-1741.

18. Zhang, M., D.T. Lynch, and S.E. Wanke, Effect of molecular structure distribution on melting and crystallization behavior of 1-butene/ethylene copolymers. Polymer, 2001. 42(7): p. 3067-3075.

19. Hasan, A.T.M.K., et al., Surface analytical approach to TiCl3-based Ziegler–Natta catalysts combined with microstructure analysis of polymer. Polymer, 2010. 51(16): p. 3627-3635.

20. Kang, J., et al., Polymerization control and fast characterization of the stereo-defect distribution of heterogeneous Ziegler–Natta isotactic polypropylene. European Polymer Journal, 2012. 48(2): p. 425-434.

Page 54: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 53 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

21. Gabriel, C. and D. Lilge, Comparison of different methods for the investigation of the short-chain branching distribution of LLDPE. Polymer, 2001. 42(1): p. 297-303.

22. Aust, N., et al., Optimization of run parameters of temperature-rising elution fractionation with the aid of a factorial design experiment. Polymer Testing, 2006. 25(7): p. 896-903.

23. Xue, Y., et al., Characterization of the microstructure of impact polypropylene alloys by preparative temperature rising elution fractionation. European Polymer Journal, 2011. 47(8): p. 1646-1653.

24. Ramachandran, R., et al., Branch length distribution in TREF fractionated polyethylene. Polymer, 2011. 52(12): p. 2661-2666.

25. Harding, G.W. and A.J. van Reenen, Polymerisation and structure–property relationships of Ziegler–Natta catalysed isotactic polypropylenes. European Polymer Journal, 2011. 47(1): p. 70-77.

26. Virkkunen, V., et al., Tacticity distribution of isotactic polypropylene prepared with heterogeneous Ziegler–Natta catalyst. 2. Application and analysis of SSA data for polypropylene. Polymer, 2004. 45(14): p. 4623-4631.

27. Virkkunen, V., et al., Tacticity distribution of isotactic polypropylene prepared with heterogeneous Ziegler–Natta catalyst. 1. Fractionation of polypropylene. Polymer, 2004. 45(9): p. 3091-3098.

28. Viville, P., et al., Characterization of the molecular structure of two highly isotactic polypropylenes. Polymer, 2001. 42(5): p. 1953-1967.

29. Xu, J., et al., Effect of composition distribution on miscibility and co-crystallization phenomena in the blends of low density polyethylene with conventional and metallocene-based ethylene–butene copolymers. Polymer, 2001. 42(8): p. 3867-3874.

30. Xu, J., X. Xu, and L. Feng, Short chain branching distributions of metallocene-based ethylene copolymers. European Polymer Journal, 2000. 36(4): p. 685-693.

31. Simanke, A.G., C. de Lemos, and M. Pires, Linear Low Density Polyethylene: microstructure and sealing properties correlation. Polymer Testing, (0).

32. Lugao, A.B., et al., Temperature rising elution fractionation, infra red and rheology study on gamma irradiated HMSPP. Radiation Physics and Chemistry, 2002. 63(3–6): p. 509-512.

33. Xu, J., et al., Temperature rising elution fractionation of PP/PE alloy and thermal behavior of the fractions. European Polymer Journal, 2002. 38(9): p. 1739-1743.

34. Soares, J.B.P. and A.E. Hamielec, Temperature rising elution fractionation of linear polyolefins. Polymer, 1995. 36(8): p. 1639-1654.

35. Vela Estrada, J.M. and A.E. Hamielec, Measurement of the bivariate distribtuion of composition and molecular weight for binary copolymers. Polymer Reaction Engineering, 1992-1993. 1(1): p. 171-201.

36. Monrabal, B., et al., Crystallization Elution Fractionation. A New Separation Process for Polyolefin Resins. Macromolecular Symposia, 2007. 257(1): p. 71-79.

37. Monrabal, B., et al., Advances in Crystallization Elution Fractionation. Macromolecular Symposia, 2009. 282(1): p. 14-24.

38. Suriya, K., S. Anantawaraskul, and J.B.P. Soares, Cocrystallization of ethylene/1-octene copolymer blends during crystallization analysis fractionation and crystallization elution fractionation. Journal of Polymer Science Part B: Polymer Physics, 2011. 49(9): p. 678-684.

39. Brown, M.E., Introduction to Thermal Analysis: Techniques and Applications (2nd Edition). 2001: Kluwer Academic Publishers.

40. Pearson, M., Capabilities of the Polymer and Advanced Composites Testing and Characterization Laboratory, 2011.

Page 55: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 54 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

41. Keating, M., I.H. Lee, and C.S. Wong, Thermal fractionation of ethylene polymers in packaging applications. Thermochimica Acta, 1996. 284(1): p. 47-56.

42. Razavi-Nouri, M., Studies of comonomer distributions and molecular segregations in metallocene-prepared polyethylenes by DSC. Polymer Testing, 2006. 25(8): p. 1052-1058.

43. Chen, F., R.A. Shanks, and G. Amarasinghe, Crystallisation of single-site polyethylene blends investigated by thermal fractionation techniques. Polymer, 2001. 42(10): p. 4579-4587.

44. Shanks, R.A. and G. Amarasinghe, Crystallisation of blends of LLDPE with branched VLDPE. Polymer, 2000. 41(12): p. 4579-4587.

45. Keating, M.Y. and E.F. McCord, Evaluation of the comonomer distribution in ethylene copolymers using DSC fractionation. Thermochimica Acta, 1994. 243(2): p. 129-145.

46. Starck, P., K. Rajanen, and B. Löfgren, Comparative studies of ethylene–α-olefin copolymers by thermal fractionations and temperature-dependent crystallinity measurements. Thermochimica Acta, 2002. 395(1–2): p. 169-181.

47. Michell, R.M., et al., Effect of sequence distribution on the isothermal crystallization kinetics and successive self-nucleation and annealing (SSA) behavior of poly(ε-caprolactone-co-ε-caprolactam) copolymers. European Polymer Journal, 2010. 46(6): p. 1334-1344.

48. Zhang, F., et al., Improved thermal fractionation technique for chain structure analysis of ethylene/α-olefin copolymers. Polymer, 2002. 43(3): p. 1031-1034.

49. Pérez, C.J., M.D. Failla, and J.M. Carella, SSA study of early polyethylenes degradation stages. Effects of attack rate, of average branch length, and of backbone polymethylene sequences length distributions. Polymer Degradation and Stability, (0).

50. Arnal, M.L., J.J. Sánchez, and A.J. Müller, Miscibility of linear and branched polyethylene blends by thermal fractionation: use of the successive self-nucleation and annealing (SSA) technique. Polymer, 2001. 42(16): p. 6877-6890.

51. Białek, M., K. Czaja, and B. Sacher-Majewska, Studies of structural composition distribution heterogeneity in ethylene/1-hexene copolymers using thermal fractionation technique (SSA): Effect of catalyst structure. Thermochimica Acta, 2005. 429(2): p. 149-154.

52. Lorenzo, A.T., et al., Confinement effects on the crystallization and SSA thermal fractionation of the PE block within PE-b-PS diblock copolymers. European Polymer Journal, 2006. 42(3): p. 516-533.

53. Pérez, C.J., et al., The use of SSA fractionation to detect changes in the molecular structure of model ethylene–butene copolymers modified by peroxide crosslinking. Polymer Degradation and Stability, 2009. 94(10): p. 1639-1645.

54. Pérez, C.J., M.D. Failla, and J.M. Carella, Advantageous use of SSA technique to observe effects of thickness, antioxidant and oxygen in gamma irradiated low density polyethylene. Thermochimica Acta, 2012. 538(0): p. 67-74.

55. Sabino, M.A., Oxidation of polycaprolactone to induce compatibility with other degradable polyesters. Polymer Degradation and Stability, 2007. 92(6): p. 986-996.

56. Bialek, M. and C. Krystyna, Studies of structural composition distribution heterogeneity in ethylene/1-hexene copolymers using thermal fractionation technique (SSA) Effect of catalyst structure. Thermochimica Acta, 2005. 429: p. 149-154.

57. Gauthier, E., et al., Correlation between chain microstructural changes and embrittlement of LLDPE-based films during photo- and thermo-oxidative degradation. Polymer Degradation and Stability, (0).

58. Li, R., et al., New polypropylene blends toughened by polypropylene/poly(ethylene-co-propylene) in-reactor alloy: Compositional and morphological influence on mechanical properties. Polymer, 2009. 50(21): p. 5124-5133.

Page 56: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 55 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

59. Spencer, M., J.S. Parent, and R.A. Whitney, Composition distribution in poly(ethylene-graft-vinyltrimethoxysilane). Polymer, 2003. 44(7): p. 2015-2023.

60. Arnal, M.L., et al., Applications of Successive Self-Nucleation and Annealing (SSA) to Polymer Characterization. Journal of Thermal Analysis and Calorimetry, 2000. 59(1): p. 451-470.

61. Lorenzo, A.T., M.L. Arnal, and A.J. Müller, High Speed SSAThermal Fractionation and Limitations to the Determination of Lamellar Sizes and Their Distributions. Journal of Macromolecular Chemistry and Physics, 2006. 207: p. 39-49.

62. Müller, A.J. and Z.H. Hernandez, Successive self-nucleation/annealing (SSA): A novel technique to study molecular segregation during crystallization. Polymer Bulletin, 1997. 39(4): p. 465-472.

63. Hosoda, S., Structural distribution of LLDPEs. Polymer Journal, 1988. 20. 64. Karbashewski, E., L. Kale, and A. Rudin, Characterization of linear low density polyethylene by

temperature rising elution fractionation and by differential scanning calorimetry. Journal of Applied Polymer Science, 1992. 44(3): p. 425-434.

65. Torabi, S.R. and N. Fazeli, A rapid quantitative method for determination of short chain branching content and branching distribution index in LLDPEs by DSC. Polymer Testing, 2009. 28(8): p. 866-870.

66. Alamo, R.G. and E.K.M. Chan, Macromolecules, 1992. 25: p. 6381. 67. Mandelkern, L. and R.G. Alamo, Thermodynamic Quantities Governing Melting, in Physical

Properties of Polymers Handbook, J.E. Mark, Editor. 1996, American Institure of Physics: Woodbury.

68. Mandelkern, L., Crystallization of Polymers. Vol. 1, Equilibrium Concepts. 2002, Cambridge: Cambridge University.

69. Mathot, V., et al., The Flash DSC 1, a power compensation twin-type, chip-based fast scanning calorimeter (FSC): First findings on polymers. Thermochimica Acta, 2011. 522(1–2): p. 36-45.

70. Macko, T., et al., Elution behavior of polyethylene in polar mobile phases on a non-polar sorbent. Journal of Chromatography A, 2003. 988(1): p. 69-76.

71. Heinz, L.-C. and H. Pasch, High-temperature gradient HPLC for the separation of polyethylene–polypropylene blends. Polymer, 2005. 46(26): p. 12040-12045.

72. Macko, T., H. Pasch, and R. Brüll, Selective removal of polyethylene or polypropylene from their blends based on difference in their adsorption behaviour. Journal of Chromatography A, 2006. 1115(1–2): p. 81-87.

73. Macko, T. and H. Pasch, Separation of Linear Polyethylene from Isotactic, Atactic and Syndiotactic Polypropylene hy High-Temperature Adsorption Liquid Chromatography. Macromolecules, 2009. 42: p. 6063-6067.

74. Macko, T., et al., Separation of propene/1-alkene and ethylene/1-alkene copolymers by high-temperature adsorption liquid chromatography. Polymer, 2009. 50(23): p. 5443-5448.

75. Chitta, R., et al., Elution behavior of polyethylene and polypropylene standards on carbon sorbents. Journal of Chromatography A, 2010. 1217(49): p. 7717-7722.

76. Greibrokk, T. and T. Andersen, High-temperature liquid chromatography. Journal of Chromatography A, 2003. 1000(1–2): p. 743-755.

77. Molander, P., et al., Determination of retinoids by packed-capyllary liquid chromatography with large-volume on-column focusing and temperature optimization. Journal of Chromatography A, 1999. 847(1): p. 59-68.

78. Szabelski, P., et al., Experimental studies of pressure/temperature dependence of protein adsorption equilibrium in reversed-phase high-performance liquid chromatography. Journal of Chromatography A, 2002. 950(1): p. 41-53.

Page 57: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 56 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

79. Lee, W., et al., Characterization of Poly(methyl methacrylate) by Temperature Gradient Interaction Chromatography with On-Line Light Scattering Detection. Macromolecules, 1998. 31: p. 344-348.

80. Lee, H.C. and T. Chang, Polymer molecular weight characterization by temperature gradient high performance liquid chromatography. Polymer, 1996. 37(25): p. 5747-5749.

81. Chang, T., et al., Polymer characterization by temperature gradient interaction chromatography. Macromolecular Chemistry and Physics, 1999. 200: p. 2188-2204.

82. Lee, H.C., et al., Characterization of Linear and Star Polystyrene by Temperature-Gradient Interaction Chromatography with a Light-Scattering Detector. Macromolecules, 1998. 31: p. 690-694.

83. Lee, W., et al., Temperature gradient interaction chromatography of low molecular weight polystyrene. Polymer, 1999. 40(26): p. 7227-7231.

84. Lee, W., et al., Characterization of polyisoprene by temperature gradient interaction chromatography. Macromolecular Chemistry and Physics, 2000. 201: p. 320-325.

85. Park, S., et al., Characterization of polydisperse poly(vinyl chloride) by temperature gradient interaction chromatography. Journal of Chromatography A, 2006. 1123(1): p. 22-25.

86. Ryu, J., S. Park, and T. Chang, Rapid molecular weight analysis of polymers by temperature gradient interaction chromatography. Journal of Chromatography A, 2005. 1075(1–2): p. 145-150.

87. Cho, D., et al., Characterization of a 4-miktoarm star copolymer of the (PS-b-PI)3 PS type by temperature gradient interaction chromatography. European Polymer Journal, 2003. 39(11): p. 2155-2160.

88. Park, S., et al., Temperature gradient interaction chromatography and matrix-assisted laser desorption/ionization time-of-flight mass spectrometry analysis of air terminated polystyryllithium. Journal of Chromatography A, 2002. 958(1–2): p. 183-189.

89. Lee, W., et al., Characterization of polystyrene and polyisoprene by normal-phase temperature gradient interaction chromatography. Journal of Chromatography A, 2001. 910(1): p. 51-60.

90. Teramachi, S., H. Matsumoto, and T. Kawai, Direction of temperature gradient for normal-phase temperature gradient interaction chromatography in polystyrene fractionation. Journal of Chromatography A, 2005. 1100(1): p. 40-44.

91. Cong, R., et al., A New Technique for Characterizing Comonomer Distribution in Polyolefins: High-Temperature Thermal Gradient Interaction Chromatography (HT-TGIC). Macromolecules, 2011. 44: p. 3062-3072.

92. Berek, D., Coupled liquid chromatographic techniques for the separation of complex polymers. Progress in Polymer Science, 2000. 25(7): p. 873-908.

93. Giddings, J.C., Sample dimensionality: A predictor of order-disorder in component peak distribution in multidimensional separation. Journal of Chromatography A, 1995. 703(1): p. 3-15.

94. Balke, S.T., Separation and Purification Methods, 1982. 11: p. 1-28. 95. Chitta, R., et al., Separating ethylene-propylene-diene terpolymers according to the content of

diene by HT-HPLC and HT 2D-LC. Polymer, 2011. 52(26): p. 5953-5960. 96. Gerber, J. and W. Radke, Topological separation of linear and star-shaped polystyrenes by off-

line 2D chromatography. Stars having high molar mass arms and quantification of the star fraction. Polymer, 2005. 46(22): p. 9224-9229.

97. Im, K., et al., Two-dimensional liquid chromatography analysis of synthetic polymers using fast size exclusion chromatography at high column temperature. Journal of Chromatography A, 2009. 1216(21): p. 4606-4610.

Page 58: Recent developments in polyolefin fractionations · Master of Chemistry – Analytical Sciences Supervisor UvA: Dr. W. Th. Kok November 20th, 2013 - 3 - Recent developments in polyolefin

- 57 - Recent developments in polyolefin fractionations

J.A.A. Knooren, 6413501

98. Nunes, C., et al., Simple and solvent-free methodology for simultaneous quantification of methanol and acetic acid content of plant polysaccharides based on headspace solid phase microextraction-gas chromatography (HS-SPME-GC-FID). Carbohydrate Polymers, 2006. 64(2): p. 306-311.

99. Edam, R., et al., Branched-polymer separations using comprehensive two-dimensional molecular-topology fractionation&#xa0;×&#xa0;size-exclusion chromatography. Journal of Chromatography A, 2008. 1201(2): p. 208-214.

100. Stoll, D.R., et al., Fast, comprehensive two-dimensional liquid chromatography. Journal of Chromatography A, 2007. 1168(1–2): p. 3-43.

101. Vilaplana, F. and R.G. Gilbert, Analytical methodology for multidimensional size/branch-length distributions for branched glucose polymers using off-line 2-dimensional size-exclusion chromatography and enzymatic treatment. Journal of Chromatography A, 2011. 1218(28): p. 4434-4444.

102. Baumgaertel, A., E. Altuntaş, and U.S. Schubert, Recent developments in the detailed characterization of polymers by multidimensional chromatography. Journal of Chromatography A, 2012. 1240(0): p. 1-20.

103. Faldi, A. and J.B.P. Soares, Characterization of the combined molecular weight and composition distribution of industrial ethylene/α-olefin copolymers. Polymer, 2001. 42(7): p. 3057-3066.

104. Roy, A., et al., DEvelopment of Comprehensive Two-Dimensional High Temperature Liquid Chromatography x Gel Permeation Chromatography for Characterization of Polyolefins. Macromolecules, 2010. 43: p. 3710-3720.

105. Ginzburg, A., et al., High-temperature two-dimensional liquid chromatography of ethylene-vinylacetate copolymers. Journal of Chromatography A, 2010. 1217(44): p. 6867-6874.

106. Ginzburg, A., et al., Characterization of polyolefins by comprehensive high-temperature two-dimensional liquid chromatography (HT 2D-LC). European Polymer Journal, 2011. 47(3): p. 319-329.

107. Lee, D., et al., Development of high temperature comprehensive two-dimensional liquid chromatography hyphenated with infrared and light scattering detectors for characterization of chemical composition and molecular weight heterogeneities in polyolefin copolymers. Journal of Chromatography A, 2011. 1218(40): p. 7173-7179.

108. Cheruthazhekatt, S., et al., Multidimensional analysis of the complex composition of impact polypropylene copolymers: combination of TREF, SEC-FTIR-HPer DSC, and High Temperature 2D-LC. Macromolecules, 2012. 45: p. 2025-2034.

109. Shih, C.-Y., et al., An integrated system for on-chip temperature gradient interaction chromatography. Sensors and Actuators A: Physical, 2006. 127(2): p. 207-215.

110. Shih, C., et al., On-chip temperature gradient interaction chromatography. Journal of Chromatography A, 2006. 1111(2): p. 272-278.