Solutions For Polymer Analysis Development
Solutions For Polymer Analysis Development
5 ARTICLE
Investigative Failure Analysis of
Plastic Using FTIR Spectrometry
14 ARTICLE
FTIR Reflectance Spectroscopy for the
Analysis of Polymer Samples
31 ARTICLE
How to Measure Polymer Materials
with X-Ray Diffraction (XRD)
38 ARTICLE
Optimizing Polymeric Materials with
Rheological Analysis
56 ARTICLE
Producing Master Curves for
Polymeric Materials
74 ARTICLE
Preparing and Analyzing PET with
Additives
Introduction to
Polymer Research
and Development
Polymers are ubiquitous in the modern world for
their fantastic tunable properties. Changing either
the chemical composition, average molecular
weights, or the curing process of polymers makes
it possible to completely modify the material
behavior of the final polymer. With advanced
machining techniques making it possible to easily
make highly complex structures from polymers, it
is hard to find many modern objects that do not
have at least one polymer component.
This eBook covers a number of different lab-scale techniques for both the chemical,
elemental and structural analysis of polymers, rheological analysis and approaches for
polymer compound extrusion and testing. Together, a comprehensive understanding of
these techniques provides a way of fully understanding polymer materials for polymer
design or manufacturing applications.
A manufacturer of precision optical equipment developed a plastic cover for a device with
specifications for color and optical transmission, surface texture, and chemical composition.
In short, the cover was to be created from a polycarbonate – acrylonitrile butadiene styrene
(PC-ABS) blend with enough titanium dioxide to give a marginally off-white color and an
optical transmissivity of less than 0.01% T over a broad spectral range, from the UV into the
near-infrared.
The opacity was needed to block ambient (room) light from reaching the optical device and
hindering low light level measurements. Originally, each of the supplied parts fulfilled the
specifications and the product delivered a satisfactory performance
An alternative supplier underbid the earlier cover supplier, and the test parts satisfied all the
requirements for opacity. As a result of this, the production was changed to include this new
supplier.
After a short period of time, the product was failing serious performance tests. The failures
were quickly traced to ambient light creating elevated backgrounds which strongly affected
the low level optical measurements.
A visual inspection of the covers did not show obvious variations from the original, but
different control experiments led to the failure being traced back to the new cover. A root
cause analysis utilizing several techniques was carried out to efficiently discover and limit
the issue.
Experimental Results
UV-Visible Spectroscopy
Diffuse transmission measurements of the earlier cover and failed cover were taken utilizing
a Thermo Scientific™ Evolution™ 220 UV-Visible Spectrophotometer* and integrating
sphere, as shown in Figure 1.
The cover by design included high quantities of particulates, which would quickly scatter
any transmitted light. Due to this, transmittance was analyzed with an integrating sphere.
Parts of covers from both good and failing devices were positioned at the transmittance port
of the sphere and spectra were gathered from 220 to 800 nm, resulting in the spectra
demonstrated in Figure 2.
Figure 2. Diffuse Transmittance UV-Visible spectra of the failed cover (blue) and good
cover (red), gathered with an Evolution 220 UV-Visible Spectrophotometer and integrating
sphere accessory.
Virtually no transmittance was quantifiable through the good cover. In contrast to the failing
cover, where a notable amount of transmittance through the visible part of the spectrum,
This evidently provided an explanation for the poor performance, - the light leak of the
device under ambient conditions - but did not explain the root cause.
The samples were heated from ambient to 650 °C at 20 °C per minute under N₂ purge, then
cooled to 550 °C, and heated again to 1000 °C at 20 °C per minute with air purge.
The first heating ramp under nitrogen pyrolyzes the organic component of the covers, and
the last temperature ramp in air burns the rest of the organic components with only oxides
of the inorganic content remaining.
The organic decomposition profiles of both covers were almost the same, showing that
each one had the same plastic composition.
Although, the good cover contained leftover inorganic components representing 5.4% by
weight, whereas the failed cover had an inorganic component of just 2.2% by weight. This
showed an extensive variation in the amounts of inorganic filler between the covers and
gave a strong hint as to the origin of the light leak.
Figure 3. Thermogravimetric analysis weight loss curves for the good cover (red) and failed
cover (blue), demonstrating that the good cover has extensively higher inorganic content
Infrared Analysis
Infrared spectra of small pieces of each cover were collected with the integrated diamond
iS50 ATR on a Thermo Scientific™ Nicolet™ iS50 FTIR Spectrometer, displayed in Figure
4.
The built-in iS50 ATR on the Nicolet iS50 has an exclusive detector which allows the
collection of combined mid- and far-IR ATR spectra down to 100 cm⁻¹. The capability of the
iS50 ATR to gather spectra in the far-IR creates the simple measurement and discovery of
inorganic fillers in plastic parts.
Figure 4. Nicolet iS50 FT-IR spectrometer with built-in diamond iS50 ATR, iS50 ABX
Automated Beamsplitter exchanger, and sample compartment iS50 Raman accessory.
When integrated with the iS50 ABX Automated Beamsplitter exchanger on the Nicolet iS50
Spectrometer, mid and far-IR spectra can be immediately collected and stitched together
utilizing a Thermo Scientific™ OMNIC™ Macros\Pro™ Visual Basic Program to give a
single spectrum of a sample from 4000 to 100 cm⁻¹.2
The ATR spectra of the plastic parts, illustrated in Figure 5, were corrected with the
advanced ATR correction algorithm3 in Thermo Scientific OMNIC Software.
The advanced ATR correction algorithm explains both relative intensity changes generated
by sample penetration depth as a function of wavelength and also for peak shifts in the
infrared spectra because of index of refraction variations between the ATR crystal and
sample.
An examination of the infrared spectra of each plastic piece reveals the polymer
composition to be similar, however the original plastic part has an elevated baseline below
800 cm⁻¹, and a sharp peak at 360 cm⁻¹, demonstrated in Figure 6, that are missing or very
weak in the spectrum of the replacement part.
The peak at 360 cm⁻¹ is under the range of a common mid-IR spectrometer combined with
a KBr beamsplitter. The iS50 ABX with a solid substrate far-IR beamsplitter allows the far-IR
range to be accessible in this analysis, while high performance across the whole range is
uncompromised.
There are further differences between the spectra which are highlighted through a spectral
subtraction. The difference spectrum (Figure 5, bottom) indicates slight peak shifts in the
polymer bands, showing a small polymer composition difference between each part,
common when contrasting plastic parts made by alternate suppliers, but a critical spectral
difference is also observed below 800 cm⁻¹.
A library search of the difference spectrum against a forensic library of automobile paint
pigments and fillers,4 displayed in Figure 7, matches rutile, one of the crystalline forms of
titanium dioxide, suggesting a formulation difference between both of the covers.
Figure 5. Advance ATR corrected infrared ATR spectra of the good plastic cover (top),
failed plastic cover (middle), and difference spectra between the two (bottom).
Figure 6. Overlay of the advanced ATR corrected spectra of the good cover (blue) and
failed cover (red), over the spectral region from 940 to 100 cm⁻¹. Note the elevated baseline
and the absorbance band at 360 cm⁻¹ in the spectrum of the good cover that are missing or
significantly reduced in the spectrum of the failed cover.
Figure 7. FT-Raman difference spectrum between the good and failed covers (blue), and
top match from a library search against a forensic automobile paint pigment and fillers
library (red), identifying a higher concentration of rutile (titanium dioxide) in the good cover.
FT-Raman Analysis
To verify the judgments taken from the infrared analysis, the two samples were also
investigated utilizing the iS50 Raman sample compartment FT-Raman accessory on the
Nicolet iS50 Spectrometer (displayed in Figure 4).
The iS50 Raman accessory fits into the sample compartment of the Nicolet iS50 FTIR
Spectrometer and does not require an external module frequently seen in alternative FTIR
spectrometer systems.
The iS50 Raman accessory allows for the simple gathering of Raman spectra with a near-
infrared beamsplitter and an InGaAs detector mounted inside of the spectrometer.
FT-Raman spectra of the good and failed covers, in addition to the spectral difference
spectrum between them, are shown in Figure 8. FT-Raman spectroscopy supports the
collection of spectra into the far-IR region, accompanying the power of the Nicolet iS50
FTIR Spectrometer with the built-in iS50 ATR and ABX in gaining access to this region.
Again, the two spectra are highly alike, showing a comparable polymer composition, with
slight differences in the spectra observable below 800 cm⁻¹, plainly seen in the difference
spectrum.
A library search of the difference spectrum using a minerals Raman library5 is presented in
Figure 9, establishing the difference between both plastic parts as rutile, confirming the
identification from infrared analysis.
Figure 8. FT-Raman spectra of the good cover (top), failed cover (middle), and subtraction
result between the two (bottom).
Figure 9. FT-Raman difference spectrum between the good and failed covers (top), and top
library search result against a minerals Raman library (bottom), showing a higher
concentration of rutile (titanium dioxide) in the good cover.
Thermogravimetric analysis showed that the composition of the original cover had around
3% more, by weight, of an inorganic filler in contrast to the replacement cover.
Infrared ATR analysis over the mid and far-IR spectral regions demonstrated that the first
cover had an extensively higher rutile (titanium dioxide) content than the second cover. The
infrared results were verified by FT-Raman spectroscopy.
This study strongly shows that it is critical to have several tools available for root cause
analysis. Most of the tools utilized can be found on the Nicolet iS50 FTIR Spectrometer.
The Nicolet iS50 can collect multi-range spectra without compromise using the built-in iS50
ATR and iS50 Raman accessories.
The analyses given by the Thermo Scientific UV-Vis and FTIR instruments, in combination
with thermogravimetric analysis, were conclusive in discovering the root cause failure of the
plastic cover.
*This is now discontinued and replaced with Evolution One Plus UV-Vis Spectrophotometer.
The ISA 220 accessory has not changed and is compatible with Evolution One Plus.
This information has been sourced, reviewed and adapted from materials provided by
Thermo Fisher Scientific – Materials & Structural Analysis.
For more information on this source, please visit Thermo Fisher Scientific – Materials &
Structural Analysis.
Figure 1. Thermo Scientific Nicolet Summit FTIR Spectrometer with the Everest ATR
Very large samples can be measured, for example, paintings, with no contact between the
instrument and the artwork using other dedicated reflectance accessories. The distorted
spectra that may result from anomalous dispersion, or the possible scattering effect as the
light interacts with the sample, are both disadvantages of the specular reflectance sampling
technique. It is difficult to get reasonable matches with most spectral libraries when there
are distorted spectra, as these are usually collected in transmission.
The specular reflectance spectrum collected with the Everest accessory from a white piece
of smooth, white plastic is shown in Figure 2. Here, the collected spectrum, in red, looks
quite different from the best match value, in blue. The Thermo Scientific™ OMNIC™
Paradigm Software calculated a match value of only 52.5. This is very low, suggesting that
the two spectra are from different materials.
Figure 2. Specular reflectance spectrum (red) from a white piece of plastic and the library
spectrum from the best search result.
Much of the spectral distortion resulting from the specular reflectance measurement is
mathematically eliminated with the Kramers-Kronig correction in the OMNIC Paradigm
Software. A spectrum is also created that looks as if it were collected in transmission.
The corrected spectrum, in red, overlaid with the top match value, in purple, is shown in
Figure 3. The match value increases dramatically to 93 once the Kramers-Kronig is applied,
indicating that the sample is most likely a polycarbonate plastic.
Figure 3. Spectrum after Kramers-Kronig corrections and library spectrum from the best
search result.
The OMNIC Paradigm Software can speed up future analyses by automatically creating a
workflow, similar to a Standard Operating Procedure, ensuring consistent results from one
analyst to the next. The actual processing steps completed for the spectral reflectance
spectrum, including the Kramers-Kronig correction, are displayed to the right-hand side of
Figure 4. The workflow is automatically created for future use (Figure 5) by clicking on the
Create Workflow button at the bottom right-hand side of Figure 4.
Figure 4. Comparison of the original spectrum from the white plastic, the KK corrected
spectrum and the best library match.
Lab managers may sometimes require additional functionality in the workflow in order to
improve it. As shown in Figure 6, a more robust workflow is achievable with a few additional
features (or Tiles):
1. Instruction tile – Sample information, such as the name and batch number, needs to
be entered by the operator.
2. Save Results tile – The spectra and results are archived to a specified file location
3. Repeat tile – A loop is added so that multiple samples can be analyzed at once
Several useful features are included with OMNIC Paradigm Software to make analysis
easy. Spectral distortions can be removed from diffuse reflectance measurements through
the application of the Kramers-Kronig correction, which significantly improves the library
match value from around 50 to greater than 90. The steps can be simplified for any scientist
or technician in the lab as the process of collecting a spectrum, applying the Kramers-
Kronig correction, and successfully searching a library can be automated into a Workflow.
This information has been sourced, reviewed and adapted from materials provided by
Thermo Fisher Scientific – Materials & Structural Analysis.
For more information on this source, please visit Thermo Fisher Scientific – Materials &
Structural Analysis.
From manufacturing to recycling, industries can benefit a lot from the use of polymer analysis
by XRF.
A recent interview with Christopher Shaffer, XRF, XRD, and OES Business Development
Manager, from Thermo Fisher Scientific, outlines the uses and benefits of this powerful
technique throughout the entire polymer value chain, looking at where XRF fits into these
processes and providing a number of example scenarios and applications of this.
To become more stable, this outer shell electron will relax, release energy and take the
place of that missing electron. This process emits energy known as fluorescence energy,
which is what we measure via XRF.
There are two styles of x-ray fluorescence: energy dispersive and wavelength dispersive.
Energy dispersive XRF is a more direct analysis of the fluorescence energy emitted from
the sample. We use an x-ray source, an excited sample and a solid-state detector to collect
the fluorescence energy. The detector is used for the discrimination and determination of
this collected energy.
Wavelength dispersive XRF takes fluorescence energy and uses a crystal to diffract and
divide this into its component wavelengths. Using the detector and the crystal, a 2 theta
angle can be determined for that individual element.
XRF reaches the detection limits of an ICP, which can be as low as PPP, PPT or even PPQ,
but with XRF, there is a lot less sample preparation required. We can typically run a sample,
as received, with minimal or no sample preparation, whereas in ICP, we must dissolve the
sample into an aqueous solution. Dissolving a sample in this way will lead to different
degrees of dilution, potentially impacting ICP’s detection limits.
For liquids, we use disposable films with different films as support. The films we use are
determined by what type of solution we are using.
The best way to measure materials depends on what type of solution is being used. For
example, gravimetric or volumetric measuring techniques work very well if we have an
aqueous solution. If you have a nonaqueous solution - like an oil solution - then gravimetric
measuring is the method to use.
We tend to use the same support, films and disposable cells as we do in liquids for loose
powders. We simply insert an amount of sample into a cell gravimetrically when working
Regardless of the loose powder sample type, we tend to use the least absorbing film. This
is typically polypropylene.
The advantages of this approach are that it is a quick, inexpensive means of conducting
heavy metal analysis. The disadvantages of this method lie in its inability to offer an
adequate analysis of light elements, primarily due to the film’s absorption rate. There are
also issues with inconsistent packing on the sample’s surface, prompting differences in
detection between particles and creating variation.
When preparing polymers, we tend to use a hot press to heat the polymer and press this
under about 10 tons of pressure. This method is ideal for good sample preparation, offering
good surface adherence and surface smoothness. It also doesn’t require a film and is
suitable for elements ranging from carbon to uranium.
The downside to this method of polymer preparation is that this is a more expensive
technique requiring a specific hot press to be available in the lab. These presses can only
reach 180 °C, and while the process does not need a melt temperature, this does need to
sufficiently soften the polymer for pressing this into a solid disc.
The majority of sample preparation with XRF is fast and relatively simple compared to other
techniques because we can measure many samples directly. Most techniques also require
linear calibrations, but if we are using XRF without standards, we can use semi-quantum
fundamental parameters to achieve quick analysis for single or multiple element
procedures.
XRF is also useful when looking to meet environmental regulations in terms of analyzing
the polymer as a waste.
In exploration, we can use XRF for mineral identification and quantification, for example, in
blast holes and pouring samples. We can analyze the elements present from PPM to 100%
levels, either to look for trace elements that identify the presence or likelihood of oil-rich
ores or to just look at the sample’s general composition and its major and minor
components.
In a typical exploration application, we would take a sample, run the XRF analysis and
determine its elemental composition. This would typically be expressed as oxides – a
stoichiometric expression of elemental concentrations.
We take that information and use another technique known as X-ray diffraction (XRD) to
determine the phase assembly or any crystalline components in the sample.
When we perform a search on a database for XRD, there will be hundreds of thousands of
possible candidates populating our search results. We can use our XRF results to minimize
the search criteria by knowing which elements to search for. This aids in the processing and
quantification of the XRD data.
The combination of these two techniques allows us to gain a complete picture of the
elemental composition and the mineralogy of the samples we are looking at.
Types of samples can range from gasoline to catalysts, additives and everything in
between. We often use different ISOs, norms or regulations that dictate which elements we
are going to look for and the concentration ranges we need to achieve. Using this
information, we can better determine which X-ray will best fit our analysis needs.
For example, when looking at crude oil, we typically monitor sulfur, chloride and other
contaminants in the crude oil. Sulfur determination, in particular, can involve very high
concentrations, depending on where the crude oil comes from - from thousands of PPM up
to 5% or 6% in value.
XRF can easily accommodate these limits using a long-range, linear regression. This can
Chloride determination is key to managing corrosion and emission issues in the refinery.
We need to monitor this and ensure that chloride concentrations are low enough so that we
don’t start corroding the refinery pipes or ore.
When working with other heavy metals, such as vanadium and nickel, we need to
determine and quantify these because they can poison the catalyst and render this
ineffective in the cracking process.
If these metals move through the refinement process and enter oil products, this can create
corrosion issues in the engine blocks. Again, XRF is excellent at detecting all of these
substances from low PPM to high percentage ranges.
Image Credit:Shutterstock/oilandgasphotogragher
For example, in gasoline, the regulations state that sulfur levels must be 10 PPM or less,
while in diesel, this is 15 PPM or less. These regulations continue to be set lower and lower,
Using energy-dispersive XRF, we can reach a detection limit of two PPM of sulfur, while
using wavelength dispersive allows us to reach as low as 0.1 PPM - well within even the
lowest regulations on either technique.
The other thing we have to consider is repeatability as this determines how reliable our
results are. In one example application, we ran the same solution in 20 different cells,
measuring an average concentration of 7.3% and a typical variation of less than 0.3 PPM.
This example confirmed very high precision from sample to sample, even using different
cells. In fact, the cells were typically where we saw most of the variation.
Chrome and nickel are some of the more common wear metals. Their presence in oil can
indicate where a part is wearing, how fast it is wearing and where failures could potentially
come from.
Contaminants like sulfur and fluoride generally come from environmental conditions. These
need to be monitored to help minimize external contaminants in the oil.
Another essential part of the refinement process is the cracking catalyst – the catalyst used
to break down the crude oil into its component hydrocarbons.
One of the main elements we need to monitor in this instance is the total chloride. This
should be done in fresh regenerator spent alumina catalyst because the effectiveness of
the catalyst may be affected if chlorine levels are too high.
We can also use XRF to evaluate suppliers, undertake R&D on new catalyst formulations,
audit controls and evaluate recycling of the bulk catalyst. The best method for preparing
these samples is via what is known as fusion - the addition of lithium tetraborate and
metaborate mixtures and fusing this into a solid glass disc. This then creates homogenous
samples which offer long dynamic ranges.
The analysis of polymer additives is another key focus in this area. These additives are
inorganic elements added to polymers to give them different properties, such as stabilizers,
flame retardants, antioxidants or colorants.
Some polymers also develop undesirable properties as they break down. When
organometallics break down, for example, polymers will degrade and start releasing certain
elements that are very toxic or environmentally damaging.
We can use XRF to monitor both these aspects - polymer additions and the breakdown of
the polymer with respect to the environmental regulations.
When working with polymers, typical elements we can investigate via XRF include
magnesium, calcium, iron, zinc, and more. These elements can be easily analyzed by
EDXRF or WDXRF, with the best choice of technique depending on which elements are in
the concentration range.
If we don’t have an exact matrix match, as long as we can get a liquid with the same
density, we can often use those liquids to create different calibrations of our actual polymer
samples.
In XRF, we only need to calibrate the system once, and we can maintain these calibrations
for the instrument’s life using something known as drift correction.
Drift correction uses samples that do not change over time. They are used to monitor the
degradation of an X-ray tube or any other components and then factor in that change to
ensure the calibration is relevant through the entire lifespan of the instrument.
Image Credit:Shutterstock/XXLPhoto
RoHS and WEEE stipulate that we have to measure different types of chromium and
bromides. XRF can only tell us the total amount of chromium or the total amount of
bromides, so we regulate to a level below chromium and the bromides, such that even if
they were in these forms, they would be below the detection limits.
The US equivalent to RoHS (a European standard) is ASTM F2617. This standard requires
monitoring the same elements at the same concentration ranges as RoHS.
We can meet all of these easily with a benchtop XRF system, with limits of detection a
factor of a hundred or more, less than their restricted levels. Overall, XRF offers a
For example, if you want to start looking at fluorine, sodium and magnesium at lower
concentration ranges, energy dispersive XRF will not provide that value for you – you
should consider moving to a wavelength dispersive XRF instead.
The ASCM6247 standard defines lower and upper limits for many of these elements in
process control: fluorine from 100 to 300 PPM, sodium from 25 to 200 PPM, magnesium
from 10 to 600 PPM, etc.
These are limits that energy-dispersive XRF will not meet, but wavelength dispersive XRF
will. We can still achieve all the limits required for heavy metal analysis in WDXRF.
In general, if an energy dispersive XRF system – which is lower cost and available as a
smaller benchtop system – is suitable for the analytical techniques and samples required,
this is likely a more appropriate choice. However, if an application requires detection that
EDXRF cannot accommodate, then it will be necessary to step up and move to a WDXRF
system.
UniQuant is a peak-based analysis tool, offering much greater accuracy than a scan-based
tool.
In a peak-based tool, we move to the known 2 theta location (or keV location), measure this
for a set amount of time - from 4 seconds and 12 seconds - and measure on a background
position.
Peak to background precision is therefore much better because we are getting much better
counting statistics through the analytical time. This ‘peak-hopping’ method is a much more
accurate method for semi-quantitative analysis.
The total count time for these types of techniques is anywhere from 14 to 20 minutes,
depending on how many elements we are looking at and what power of system we have.
We typically look at elements from fluorine to uranium but using WDXRF; we can add other
elements such as carbon, boron and nitrogen.
This calibration is maintained through the use of drift correction. It is also manipulable,
meaning we can shrink the total analysis time by eliminating elements we know won’t be
there.
Alternatively, if we have a particular element that we need to measure with very high
precision, we can increase the individual count time on that element without increasing the
count time across the board.
With the sample loaded, we can continue to do the analysis. If we don’t have a method for
the sample, we can use UniQuant - the standard routine. A range of programs are
available, including vacuum mode.
The analysis is started with the single press of a button, and the system provides updates
at each stage of the analysis.
Once the first condition has been evaluated and elements detected, we can automatically
label them and then expand around that with further conditions. As the keV level continues
to be increased, further conditions may reveal additional elements present but not currently
being excited enough to be displayed.
The system progresses from condition to condition, increasing keV and exciting different
elements as it does. The elements immediately to the left of the major excitation pump will
be excited most efficiently.
Once the analysis is complete, the user is returned to the main menu to look at the results
and perform any calculations that need to be done.
This information has been sourced, reviewed and adapted from materials provided by
Thermo Fisher Scientific – Materials & Structural Analysis.
For more information on this source, please visit Thermo Fisher Scientific – Materials &
Structural Analysis.
Disclaimer: The views expressed here are those of the interviewee and do not
necessarily represent the views of AZoM.com Limited (T/A) AZoNetwork, the owner and
operator of this website. This disclaimer forms part of the Terms and Conditions of use
of this website.
Fast, Dynamic Analysis with Benchtop XRD. Video credit: Thermo Fisher Scientific –
Materials & Structural Analysis
Key observables for polymer materials include the type of polymer – for example,
polyethylene (PE) or polypropylene (PP) – or its crystallinity. Polymers also exhibit varying
microstructures, much like those found in common metal alloys, like steel. These
microstructures also influence polymer materials’ mechanical properties.
It is important that users and producers in highly industrialized environments are able to
rapidly screen products to ensure they possess appropriate properties.
One straightforward and quick way to perform this screening lies in the use of X-ray
diffraction (XRD) combined with whole pattern Rietveld refinements. This approach
represents an ideal solution for QC/QA or research-related applications.
In a recent study, the Thermo Scientific™ ARL™ EQUINOX 100 Diffractometer was used to
evaluate a number of key polymer properties.
Thermo Scientific ARL EQUINOX 100. Image Credit: Thermo Fisher Scientific – Materials &
Structural Analysis
Additional qualitative phase analyses confirm the presence of Polyethylene (PE) and
Polypropylene (PP) type materials.
In this example, a calculation of the amorphous content (in w%) used a density of 0.85
g/cm3, confidently providing results for both PE and PP samples with high and low D
values.
It should be noted that the intergrowth structure of α-PP and β-PP is central to the resulting
material properties, prompting this observable to be widely used in industrial applications.
These results were achieved via Whole Pattern Rietveld refinement with MDI JADE 2010,
which affords users a robust and standardless method for the quantification of amorphous
content.
As the system is low wattage, this does not necessitate the use of an external water chiller
or additional peripheral infrastructure. The instrument can therefore be easily transported
between laboratories or from the laboratory to the field.
The obtainable data quality from the benchtop ARL™ EQUINOX 100 is comparable to that
of a common high-power floor-standing instrument.
In the example presented here, both qualitative and quantitative analysis was undertaken
using MDI JADE 2010 and the ICDD PDF4+ Organic database.
Acknowledgments
Produced from materials originally authored by Ju Weicai, Raphael Yerly and Dr. Simon
Welzmiller from Thermo Fisher Scientific.
This information has been sourced, reviewed and adapted from materials provided by
Thermo Fisher Scientific – Materials & Structural Analysis.
For more information on this source, please visit Thermo Fisher Scientific – Materials &
Structural Analysis.
Polymers used in the production of plastics are typically processed at higher temperatures,
working with these in a molten state. Properly comprehending their melting, deformation
and flow is central to effectively processing and transforming these into end products.
Polymers are regarded as viscoelastic materials, meaning that they exhibit both viscous
(liquid-like) and elastic (solid-like) properties. Their high molecular weight and chemical
structure cause polymer melts to exhibit a complicated flow and deformation behavior.
Too much elasticity may result in flow anomalies and undesirable effects during a range of
common processing steps,1 for example, a melt stream may swell as it exits the narrow die
of an extruder.
Figure 1 displays a series of other examples of flow anomalies resulting from the elastic
properties of polymeric fluids.
Figure 1. Typical flow anomalies of viscoelastic polymer fluids. Image Credit: Thermo
Fisher Scientific – Materials & Structural Analysis
Two tools frequently employed in the measurement of polymer melt viscosities are capillary
viscometers and melt flow indexers. These instruments do not offer any insight into the
viscoelastic properties of the tested sample, however.
In contrast, the use of a rotational rheometer with the capacity to perform rheological tests
with small oscillatory mechanical excitations can offer users a robust and comprehensive
investigation into these properties.
This article offers an overview of the various rheological tests which can be conducted
using rotational rheometers. It also outlines how the results of these tests relate to a range
of processing conditions and final product properties.
This is because of polymer melts’ high elasticity and their tendency to result in edge failures
when exposed to the large deformations typical in a rotational rheometer.
The Cox-Merz rule states that when complex viscosity (Iη*I) is derived from oscillatory
frequency sweep measurements plotted against the angular frequency (ω), this will be
identical to the steady-state shear viscosity determined via rotational testing plotted against
the shear rate.2
The Cox-Merz rule is an empirical rule applicable to a wide range of polymer melts and
solutions.
Figure 2 compares viscosity data acquired from rheological tests in rotation versus data
acquired in oscillation mode.
Figure 2. Comparison of viscosity data obtained from a steady state shear (red symbols)
and an oscillation frequency sweep test (green symbols). Image Credit: Thermo Fisher
Scientific – Materials & Structural Analysis
Viscosity begins to decrease as the end of the Newtonian (zero shear viscosity) plateau is
reached. At this point, the shear viscosity (red symbols) drops abruptly and ceases to
display a smooth and continuous progression.
This observable drop is the result of sample fracture at the edge of the measuring
geometry, a phenomenon caused by secondary flow fields.1 It should be noted that the
oscillation frequency sweep (green symbols) provides higher data quality and can
accommodate a broader frequency range.
The oscillatory frequency sweep’s enhanced testing range is due to the small amplitudes of
the imposed oscillatory shear. Therefore, conducting an oscillatory frequency sweep and
applying the Cox-Merz rule is the method of choice when looking to acquire shear viscosity
data for polymeric materials.
When maintained within its LVR, the material’s microstructure remains unchanged. Its
rheological properties will also remain constant and unaffected by the applied stress of
deformation; for example, the storage and the loss modulus (G’ and G’’, respectively) or the
complex viscosity.
Should a critical deformation or stress value be reached, this will prompt the material’s
microstructure to change, and in turn, its rheological parameters.
In the example presented here, the rheometer software has automatically calculated the
end of the linear viscoelastic range of this LDPE melt, determining this to be equal to a
deformation of 55%.
Should it be necessary to perform a frequency sweep test over a wider frequency range of
several orders of magnitude, it is advisable to undertake a number of amplitude sweeps at
various frequencies to ensure the selected deformation is within the LVR throughout the
whole frequency range.
Figure 3. Storage modulus G’, loss modulus G’’ and the complex viscosity Iη*I as a function
of the deformation γ for a LDPE melt at 1 Hz and 190 °C. Image Credit: Thermo Fisher
Scientific – Materials & Structural Analysis
For example, the acquisition of shear rate-dependent viscosity data via oscillatory
frequency sweep experiments can be processed using the Cox-Merz rule to quantify a
material’s flow resistance during high shear processing applications such as injection
molding or extrusion.
In contrast, low frequency/shear data (for example, zero shear viscosity / η0) can be
employed to calculate the average molecular weight (Mw) of a polymer melt. This is
achieved via the following formula:
η0 = K • MW 3.4 (1)
Within this equation, the prefactor (k) is dependent on the polymer’s molecular structure.3
Equation 1 is applicable to polymers with a linear chain structure and a molecular weight
above a critical value (Mc).
Figure 4. Shear rate depending viscosity of a polystyrene melt and typical applications.
Image Credit: Thermo Fisher Scientific – Materials & Structural Analysis
Frequency sweep data also offers a means of directly measuring a polymer’s viscous and
elastic properties. These properties are represented by the storage and the loss moduli (G’
and G’’, respectively) and are generally measured at varying frequencies and timescales.
Data from these measurements can illustrate the general structure of a material, alongside
valuable information on its molecular weight (Mw) and molecular weight distribution (MWD).
Figure 5 illustrates an example of this crossover point shifting, highlighting where G’ = G’’
and the MW or MWD differ for an otherwise identical polymer melt.
Figure 5. Storage modulus G’, loss modulus G’’ and the complex viscosity Iη*I as a function
of the angular frequency ω for a polystyrene melt at 190 °C. Image Credit: Thermo Fisher
Scientific – Materials & Structural Analysis
Flow anomalies caused by the elasticity of polymer melts may adversely affect product
quality in extrusion and other polymer processing applications.
Figure 6 compares storage modulus data as a function of the applied frequency for a
number of polyethylene samples with differing Melt Flow Indices (MFI).
Each of these three PE samples was processed under identical conditions using a 16 mm
parallel twin screw extruder.
Once they reached the end of the extruder barrel, the melts were forced through a vertical
rod capillary die with a diameter of 1 mm and an L/D ratio of 10. A laser micrometer was
employed to measure the die swell of the extrudate.
A die-swell of 0.5 mm was found to result in a total strand diameter of 1.5 mm for the PE
with the lowest molecular weight and highest MFI (~20). The strand exited the extrusion line
as an even strand with no observable surface defects (Figure 6).
The PE sample exhibiting medium MFI (~2) displayed an uneven surface structure with a
changing diameter, while the PE sample exhibiting the lowest MFI (~0.2) and highest
molecular weight displayed obvious indications of melt fracture under the extrusion
conditions also utilized for the other samples.
An examination of the rheological data highlights that the three samples are notably
different in terms of their elasticity (G’). This is particularly prevalent at the lowest frequency
(10-2 Hz), where the values in G’ differ in one or more orders of magnitude.
Storage modulus also offers a sensitive indicator of the elasticity incorporated by a high
molecular weight tail.
Figure 7 compares the results of a total of three frequency sweeps performed on a low
molecular weight LDPE and two blends of an identical LDPE that also featured a small
weight fraction of a high molecular weight PE.
Figure 7. Storage modulus G’ and loss modulus G’’ as a function of the angular frequency
ω for a low molecular weight polyethylene melt and two blends at 190 °C. Image Credit:
Thermo Fisher Scientific – Materials & Structural Analysis
G’ shows clear differences for these three melts in the low frequency range, and a small
fraction of 1 wt% of high molecular weight PE can also be detected.
It should be noted that it is rarely possible to view these small differences when employing
Gel Permeation Chromatography (GPC) or related techniques to ascertain molecular
weight distribution.
MFI results achieved via a capillary viscometer would also fail to reveal differences between
the three samples.
Storage modulus data acquired via oscillatory frequency sweep experiments remains the
most sensitive indicator of a high molecular weight tail in a polymer melt.
It is important to acquire this data because even small amounts of high molecular weight
fraction can lead to flow anomalies that negatively impact final polymer strand quality.
Figure 5 displays rheological data obtained over an angular frequency range from less than
10-2 rad/s to over 104 rad/s. Acquiring rheological data over such a wide range requires
more than a single frequency sweep test.
Both the low and high frequency regions are limited by time issues (for example, duration of
a single oscillation) or rheometer specifications (for example, maximum frequency).
TTS leverages the principle that both temperature and frequency (time) affect the
viscoelastic behavior of polymer melts in similar ways,3 meaning that it is possible to
perform a number of frequency sweeps over a smaller range at different temperatures.
A data set (at one temperature) is selected as a reference, allowing a master curve to be
generated by shifting other results towards the reference curve.
It is, therefore, possible to acquire rheological data over a much wider frequency range
when using the TTS principle, particularly when compared to a single frequency sweep
experiment.
TTS is applicable to many polymer melts and blends, but this is generally only a viable tool
when used over a limited temperature range.4
Figure 8a displays results from a number of frequency sweep tests conducted at various
temperatures.
The TTS principle was applied to these results, with 190 °C chosen as a reference
temperature, generating a master curve (Figure 8b) which includes viscoelastic data over
almost 8 orders of magnitude in frequency.
It is possible to divide the master curve into three regions. At low frequencies, the sample is
in the terminal region prompting the polymer melt to exhibit predominantly viscous behavior.
Material behavior in the terminal region is controlled by long molecule chain relaxation
processes, while G’ and G’’ typically have slopes of 2 and 1 in a double logarithmic plot in
this region.
At medium frequencies, however, a transition takes place with a crossover between G’ and
G’’. In this scenario, viscoelastic behavior is primarily driven by the polymer’s molecular
weight distribution.
At the highest frequencies, the sample exhibits predominantly elastic behavior, with G’
larger than G’’. In this scenario, the polymer’s behavior is controlled by the rapid relaxation
motion of the shortest polymer chains.
G’ and G’’ data obtained over a wide frequency range can also be employed in the
calculation of molecular weight and molecular weight distribution for a significant number of
linear thermoplastic homopolymers.
This is achieved by ensuring that the tested frequency range includes data from the low
frequency terminal region to the end of the high frequency plateau region.
DMTA testing involves a material being exposed to oscillatory mechanical excitation while
its temperature is continuously altered.
Data acquired throughout this process can be used to identify characteristic phase
transitions, for example, glass transition, melting and/or crystallization within the polymer
matrix.
DMTA can also be used to evaluate final product performance and to investigate the
robustness of specific application-based properties, such as brittleness, stiffness, damping
or impact resistance.
Figure 9. Storage modulus G’, loss modulus G’’ and tanδ as a function of temperature for a
polyetheretherketone. Image Credit: Thermo Fisher Scientific – Materials & Structural
Analysis
This was tested from below its glass transition to slightly under its melting temperature,
supported by the use of a specialized solids clamping tool designed for rotational
rheometers.5
A lab scale injection molding system was used to prepare the rectangular sample.6 A
material’s glass transition can be identified using different metrics under rheological testing,
with the most common metric utilizing the maximum in the loss modulus G’’.
The initial decrease in the storage modulus G’ or the maximum in the tanδ (G’’/G’) can also
be used to highlight glass transition.
In the example presented in Figure 9, the maximum in G’’ is found in the center of a wider
transition range, while the onset of the G’ decrease is close to the beginning of the
transition. The maximum in tanδ is found close to the end of this range.
sweep test.
Figure 10. Generalized behavior of polymer sample in a DMTA test. Image Credit: Thermo
Fisher Scientific – Materials & Structural Analysis
Semi-crystalline polymers will always transition from a glassy region at low temperatures,
shifting to a rubbery plateau and ultimately into their melt state when subjected to higher
temperatures.
The step height from the glassy region to the rubbery plateau is dependent on the
polymer’s degree of crystallinity - as the degree of crystalline domains in the polymer
increases, this, in turn, causes a decrease in step height between the two regions.
It should also be noted that low molecular weight polymers do not exhibit a rubbery plateau.
In these examples, once the glass transition is complete, the material will change to a soft
melt with G’ decreasing in line with increasing temperature.
Cross-linked polymers do not melt at all, however. These remain in a rubbery state until the
onset of thermal decomposition.
Rather, extensional flows tend to occur in polymer melt processes in the form of film
blowing, fiber spinning, injection molding or foam extrusion.
Figure 11. Extensional viscosity as a function of strain rate for a non-branched HDPE (top)
and highly branched LDPE (bottom). All tests were performed at 150 °C. Image Credit:
Thermo Fisher Scientific – Materials & Structural Analysis
These tests were conducted using the Sentmanant Extensional Rheometer (SER) fixture
designed for use with rotational rheometers.7
The plot on the left displays the extensional behavior of a non-branched high-density
polyethylene (HDPE) sample - no observable strain hardening occurred for this type of
material.
The plot on the right displays the results of identical experiments and was performed with a
highly branched low-density polyethylene (LDPE) sample.
The red curves in the images indicate transient shear viscosities multiplied by three
according to the Trouton ratio for uniaxial extension.8 This shear viscosity data was
acquired via rotational step experiments.
The extensional behavior of the branched LDPE sample was notably different from the
behavior observed in shear flow, unlike the linear HDPE sample. The LDPE sample
displayed shear hardening behavior throughout extensional testing, particularly at higher
deformation rates.
This behavior cannot be measured and evaluated using standard rotational rheological
measurements.
Conclusion
Knowledge of a polymeric material’s viscoelastic properties is key to optimizing formulations
and blends, adapting a process to accommodate the properties of a specific material, and
minimizing the risk of flow anomalies.
The use of rotational rheometers to perform appropriate rheological tests facilitates the
investigation of a polymers’ viscoelastic behavior; from the melt-state to the solid-state and
all points in between.
References
1. C.W. Macosko, Rheology: Principles, Measurements, and Applications, Wiley-VCH;
New York (1994).
2. W.P. Cox and E.H. Merz, Journal of Polymer Science, 28, 619 (1958).
3. J.D. Ferry, Viscoelastic Properties of Polymers, 3rd ed., John Wiley & Sons, N.Y.
(1980).
4. J. Dealy, D. Plazek, Time-Temperature Superposition – A Users Guide, Rheology
Bulletin, 78(2), (2009).
5. C. Küchenmeister-Lehrheuser, K. Oldörp, F. Meyer, Solids clamping tool for Dynamic
Mechanical Analysis (DMTA) with HAAKE MARS rheometers, Thermo Fisher
Scientific Product Information P004 (2016).
6. Thermo Scientific HAAKE MiniJet Pro, Thermo Fisher Scientific Specification Sheet
(2014).
7. C. Küchenmeister-Lehrheuser, F. Meyer, Sentmanat Extensional Rheometer (SER)
for the Thermo Scientific HAAKE MARS, Thermo Fisher Scientific Product Information
P019 (2016).
8. F. T. Trouton, Proc. R. Soc. A77, 426−440 (1906).
Acknowledgments
Produced from materials originally authored by Fabian Meyer and Nate Crawford from
Thermo Scientific.
This information has been sourced, reviewed and adapted from materials provided by
Thermo Fisher Scientific – Materials & Structural Analysis.
For more information on this source, please visit Thermo Fisher Scientific – Materials &
Structural Analysis.
Polymers comprise repetitive units of smaller molecules known as monomers that combine
to form long chains. These polymer chains in a molten state or solution include random
coils with strong intermolecular interactions through entanglements. As a result, the flow
and deformation behavior is complex. Rotational rheometers are commonly used to
characterize flow behavior and assess the processability of various plastics.
Polymer melts have a high elasticity due to their structure, which often restricts rheological
measurements in a rotational mode where secondary flows and edge fractures can occur.
As a result, frequency-controlled oscillatory tests are used to gain insight into the
rheological properties of various processing parameters.
According to the equation, Cox and Merz discovered that for linear, unfilled polymers, the
dynamic viscosity η as a function of the shear rate ẏ can be correlated with the complex
viscosity |η*| as a function of the angular frequency ω according to the equation (1).2
However, the so-called Cox-Merz rule does not apply to multiphase liquids, such as
suspensions or chemically crosslinked and gelled systems.3
Therefore, acquiring rheological data in oscillation could take less than a second, minutes,
or even days or weeks, based on the chosen frequency. On the other hand, the maximum
frequency of commercially available rheometers is usually limited to around 100 Hz (628
rad/s) by the inertia of the measuring system itself.
To address this problem, the time–temperature superposition (TTS) principle can generate
master curves with a broader frequency range. This article offers a guideline for the
generation of master curves. It also verifies whether the TTS principle can be applied to a
data set by employing the van Gurp-Palmen plot.
A deformation amplitude Y of 0.5 % was chosen within the sample’s linear viscoelastic
range. At 180, 200, 220, 240, 260, and 280 °C, frequency sweeps were performed from 300
rad/s to 1 rad/s. Each frequency sweep was performed with a fresh sample to ensure no
thermal degradation occurred. Matching sample discs were prepared in advance for all
rheological measurements using a HAAKE Minijet Pro Mini Injection Molding System.
G' and G" intersect at an angular frequency of 1.78 rad/s at the chosen reference
temperature of 220 °C. PMMA exhibits this crossover at a much higher frequency of 144
rad/s at 280 °C. PMMA, on the other hand, shows no crossover of G' and G" at 180 °C. G'
and G" appear to be different parts of the same curve depending on the temperature.
This observation is used when implementing the TTS principle. The individual frequency
sweep data sets are shifted horizontally and vertically towards a reference temperature T0
to generate a master curve covering a frequency range much larger than the limited range
The TTS employs temperature-dependent horizontal and vertical shift factors, aT and bT,
respectively.5 The horizontal shift factor aT specifies the shift of the measured moduli data
along the x (frequency)-axis towards the data set acquired at the reference temperature T0.
The vertical shift factor bT is characterized by the density ρ(T) of the polymeric material at a
given temperature in comparison to the reference ρ0 (T0), as shown in equation (2).5
Equation 2
Figure 2. Rheological frequency sweep data at temperatures of 180 °C, 220 °C, and 280
°C for PMMA. Image Credit: Thermo Fisher Scientific – Materials & Structural Analysis
Modern rheometer software, such as the HAAKE RheoWinTM Rheometer Control Software,
shifts rheological data and generates master curves to obtain information about viscoelastic
properties over a wide frequency range. The TTS principle cannot be applied to
thermorheologically complex materials in which two or more relaxation mechanisms have
different temperature dependencies.5
The rheological data in Figure 2 were used to generate the PMMA master curve in Figure 3.
Additional measurement data was collected at 200 °C, 240 °C, and 260 °C to allow for
more overlap and higher precision. The reference temperature was set at 220 °C.
Figure 3. G’ and G’’ master curve at a reference temperature of 220 °C obtained from
frequency sweep data at 180 °C, 200 °C, 220 °C, 240 °C, 260 °C and 280 °C for a PMMA
melt. Image Credit: Thermo Fisher Scientific – Materials & Structural Analysis
TTS was used to prolong the original angular frequency range of each measurement data
set (1 to 300 rad/s) to 0.03 to 2600 rad/s. This range is broad enough to include shear rates
encountered in many standard polymer processing techniques, such as extrusion (1...1.000
s-1) and injection molding (10...10.000 s-1).6
The master curve could be lengthened to even higher frequencies/shear rates with
additional frequency sweeps at temperatures below 180 °C.
Van Gurp–Palmen disclosed that the phase angle δ as a function of complex shear
modulus |G*| for each isothermal frequency sweep used to develop a master curve must
superpose into a single continuous curve to ascertain the applicability of the TTS principle
for a given polymer fluid.7 Figure 4 depicts the van Gurp–Palmen plot of the individual
PMMA frequency sweep measurements.
Figure 4. van Gurp-Palmen plot for the single PMMA frequency sweep tests performed at
different temperatures. Image Credit: Thermo Fisher Scientific – Materials & Structural
Analysis
The van Gurp–Palmen plot can be used to determine whether a polymer sample, such as
the PMMA grade used in this study, is thermorheologically simple.
The rheometer software automatically obtains information on the complex viscosity |η*| of
the sample based on the G' and G" master curves (3).
Equation 3
Numerous material-dependent parameters can be acquired from a master curve for the
complex viscosity, using a suitable model fitting, such as the Carreau–Yasuda curve fit
described in equation.(4).8
Equation 4
In Equation 4, η∞ is the viscosity at an infinite shear rate, η0 is the zero-shear viscosity, and
λ is a time constant that represents a characteristic relaxation time of the sample. The
power law index n describes shear thinning behavior at high shear rates, whereas the
transition factor a refers to the process of polymer chain disentanglement within the
polymer network, and the transition from the Newtonian plateau to the material’s shear
thinning behavior.
Figure 5 depicts the various Carreau–Yasuda curve fit parameters on an example master
curve.
A Carreau-Yasuda curve fitting was executed on the complex viscosity data of the master
curve to acquire these material-specific parameters for the examined PMMA sample. The
master curve was generated, and the curve fitting was performed using the respective
HAAKE RheoWin Software functionalities.
The HAAKE RheoWin Software calculated the Carreau–Yasuda parameters for the PMMA
master curve at 220 °C, as shown in Table 1.
Table 1. Carreau–Yasuda curve fit parameters for a PMMA master curve at a reference
temperature of 220 °C. Source: Thermo Fisher Scientific – Materials & Structural Analysis
Parameter Value
η0 100.200 Pas
λ 0.81
a 0.74
n 0.20
The zero-shear viscosity η0 and the transition factor are essential parameters for polymer
processing. It is well known that η0 is proportional to a polymer’s average chain length. The
distribution of different chain lengths relates to the transition factor. As a result, changes in
zero-shear viscosity after processing or recycling can be directly correlated with a polymer’s
molecular mass or polydispersity.
Conclusion
The creation of a master curve from frequency sweep data collected at different
temperatures and using the TTS principle for a PMMA sample was outlined in this article.
The van Gurp-Palmen plot was developed to provide a quick and easy way to determine
whether a set of frequency sweep data is suitable for TTS shifting.
The Carreau–Yasuda curve fitting model was discussed as a method for obtaining various
polymer-specific parameters from master curves, which can be used to characterize the
References
1. Alger, M–Polymer Science Dictionary (2017)
2. Cox, W. P. and Merz, E.H.–Correlation of Dynamic and Steady flow Viscosities–J.
Polym. Sci. (1958)
3. Geissle, W. and Hochstein, B.–Validity of the Cox-Merz rule for concentrated
suspensions–J .Rheol. 47 (4) (2013)
4. Dealy, J.M. , Read, D.J., Larson, R.G.–Structure and Rheology of Molten Polymers–
2nd edition (2018)
5. Plazek, D.–Time-temperature superposition–A users guide – Rheol. Bul. 78 (2) (2009)
6. Carnicer, V. et al.–Microfluidic rheology: A new approach to measure viscosity of
ceramic suspensions at extremely high shear rates–J. Oceram 5 (2021)
7. Van Gurp, M., Palmen, J,–Time-temperature superposition for polymeric blends–
Rheol. Bul. 67 (1) (1998)
8. Malkin, A. and Isayev, A.–Rheology–Concepts, Methods, and Applications–3rd edition
(2017)
This information has been sourced, reviewed and adapted from materials provided by
Thermo Fisher Scientific – Materials & Structural Analysis.
For more information on this source, please visit Thermo Fisher Scientific – Materials &
Structural Analysis.
In this interview, AZoM talks to Dirk Leister from Thermo Fisher Scientific about the role of
torque rheometers in the field of polymer compound development.
R&D customers in the sectors of novel polymer compounds or polymer additives are often
limited by the amount of material available to develop formulations or do pilot-scale
production. Our standalone twin-screw compounders offer flexible compounding
configurations for small sample amounts making them particularly well-suited for R&D work.
They are also available in different grades of steel for the product contact parts so they can
be used in more regulated environments such as pharmaceuticals, cosmetics and food
applications.
Our rheometer solutions help to characterize polymer compounds in a melt or in their final
shape to optimize processing or ensure quality requirements.
For example, some recent work went into the area of “green tires” to increase fuel efficiency
for automobiles. The goal was to develop new rubber compounds from ecologically friendly
resources with the aim to reduce roll-resistance and wear; the reduction in resistance helps
save fuel and make the tires themselves last longer.
Another hot research topic involves novel battery technologies for electric cars. The focus
here is on battery separator films and electrode materials. Both critical battery components
can be manufactured using a twin-screw extruder. The lab-scale size of our equipment
makes production of those materials fast and resource efficient.
In both examples, it’s important that the new materials are thoroughly compounded at the
start. Then, after compounding, extensive testing of the material properties is required. This
is where the unique modularity of the HAAKE PolyLab Torque Rheometer comes into play;
it can handle both kinds of tasks.
As I said before, the HAAKE Polylab Torque Rheometer blends the worlds of compounding
and rheological material testing into one instrument. That’s how the platform can serve
multiple research and quality control tasks. The modular design allows it to combine four
instrument capabilities in one system solution.
The HAAKE PolyLab Torque Rheometer has a strong, reliable drive and measuring unit,
and it's flexible enough to combine with different processing devices. These are batch
mixers, single screw extruders, twin-screw compounders, and rheological testing solutions
under process-relevant conditions. That means the same system can be used to compound
new materials and, with just a few set-up changes, test those same new materials for
relevant material properties.
This flexibility helps customers to establish workflows specific to their needs and to rapidly
adapt to changing requirements in the future. By utilizing different modules, customers can
leverage their workflows in both R&D and QC environments. For example, the batch mixers
can be used for recipe development in R&D and quality determinations in QC. The mixers
also help to understand important material properties like melting, degradation and flow
behavior to optimize final products. Single screw extruders can be used to produce test-
specimen like sheets and films to test additive dispersion in a final product.
The last, but not least of the four applications is rheological testing under process-relevant
conditions. Different capillary dies are available depending on the material to be tested and
the desired shear rate range; these allow the customer to reliably determine rheological
behavior. This is of prime importance for machine and tool design, as well as for the
optimization of final product.
For example, the HAAKE PolyLab Torque Rheometer saves time and reduces operating
errors as users learn just one software system to perform many different activities. That
makes it faster and easier for users to learn how to use the software and do the different
tasks. It also reduces the chance for error by not having to operate different software
packages for standalone mixers, extruders and rheometers.
Often software methods developed in the R&D lab are complex and involve multiple steps
until a full compounding or mixer test is conducted. The HAAKE PolyLab Torque Rheometer
provides consistency from R&D to QC with a mode of software operation that guides users
in QC to set up the system correctly and run pre-programmed methods by pressing a single
button. The software system is also available in multiple languages for businesses and
processes that span location sites in multiple countries.
Another way the HAAKE PolyLab Torque Rheometer saves time is through its ability to
work with small sample volumes. It not only saves on material costs, but also allows the
customer to produce and characterize more samples in less time.
Polylab V4
The flexibility of the HAAKE PolyLab Torque Rheometer shows through when it comes to
transferring work from the lab to the production environment. With an extremely broad
range of available downstream equipment, such as feeders for liquid and solid materials,
dies and take-off devices, users can set up entire compounding lines in the lab that have
the same functionality as real-world production environments. Because the HAAKE
PolyLab Torque Rheometer allows users to do all work under the conditions of the process
environment, the results obtained initially allow for easy scale-up to the production area. In
fact, the HAAKE PolyLab Torque Rheometer provides data from the analysis of material
properties that can be used directly for quality control purposes.
And of course, the HAAKE PolyLab Torque Rheometer is an open system – different
sensors can be attached to it to get additional information. For example, conductivity
measurements in the mixer can help to analyze the dispersion of carbon black in rubber
compounds in an online process. Plug-and-play functionality means the system
automatically detects attached accessories and helps to minimize set-up errors. So it’s
faster to do multiple applications on one instrument, and the changeover from one
application to the next is quick and easy.
Powder injection molding (PIM) is one field where the HAAKE PolyLab Torque Rheometer
is deployed frequently. For PIM feedstock, polymers are compounded with a high load of
metal or ceramic powders (80% or more). The HAAKE PolyLab Torque Rheometer enables
efficient compounding of those materials.
Additive manufacturing, which includes 3D printing, makes up another field where we have
seen a steep rise in interest for specialized material in the form of polymer filament to be
used with 3D printers.
Many customers have asked to explore possibilities with the HAAKE PolyLab Torque
Rheometer beyond polymers. For example, we are working with industry partners and
universities in the areas of food extrusion and cosmetics formulation where the system is
used as a continuous mixing device.
Such demand encourages us to continue developing the capabilities of the HAAKE PolyLab
Torque Rheometer; we’re always developing new accessories and enhancements to
ensure the system can meet future market demands. In fact, we just published an
application compendium that shows many specific applications for the HAAKE PolyLab
Torque Rheometer; it can be found on our website and even on your AZO Materials web
pages. It’s called Flexible solutions for advanced material development.
We plan to grow the capabilities of the flexible torque rheometer system. There are many
new applications to come for the HAAKE Polylab Torque Rheometer. As an open system,
we may develop different sensors to provide users with additional information during
operation. For example, simultaneous conductivity measurements while compounding new
rubber compounds would give users a quality measurement of the homogeneity of the
rubber compound.What are the plans for the future of torque rheometers and related
polymer instruments at Thermo Fisher Scientific?
We’re always striving to increase the range of our equipment. As another example,
additional die designs can help in setting up a small-scale production system to more
effectively mimic a large-scale system. And those are just a few ideas of what can be done
with the HAAKE PolyLab Torque Rheometer.
Dirk has over 20 years of experience in Sales and Marketing for laboratory and
manufacturing equipment especially in the pharmaceutical and life sciences industries.
Disclaimer: The views expressed here are those of the interviewee and do not
necessarily represent the views of AZoM.com Limited (T/A) AZoNetwork, the owner and
operator of this website. This disclaimer forms part of the Terms and Conditions of use
of this website.
Request a Quote
Methods
Sample Preparation
The Thermo Scientific™ HAAKE™ MiniLab Micro Compounder with co-rotating screws
(Figure 1) was used to prepare the mixtures of PET with additives at 270 °C with a screw
speed of 50 rpm. For 15 minutes, the sample was mixed by re-circulating. During the
mixing process the pressure drop was monitored in the slit capillary of the backflow channel
(Figure 2).
Figure 2. HAAKE MiniLab Micro Compounder backflow channel built as slit capillary with
two pressure sensors.
Rheological Test
The rheological tests were performed using 20 mm parallel plates and a gap of 1.4 mm on
a Thermo Scientific™ HAAKE™ MARS™ Rheometer with an electrical heated oven at 270
°C under nitrogen atmosphere. All samples had been primarily tested in an amplitude
sweep to determine the linear viscoelastic range. New test specimens were used for
frequency sweeps from 0.1 to 46 Hz. The deformation for all tests was 0.5%, thus in a safe
Results
In the recirculation mode, the pressure profile can be monitored over a period of time by the
pressure difference of the two pressure sensors, built in backflow channel (see Figure 2). At
the start of the test, material is filled into the micro compounder. This causes a pressure
peak. After all the material is filled in and the temperature equilibrated, the pressure profile
over time can refer to a reaction of the polymer. A decrease of the pressure over time
shows an alteration of the material. For plain PET, for instance, this can be a reaction of the
polymer with water (moisture) where the polymer degrades. A reduction of the pressure is
in line with a lower viscosity of the PET. When the pressure increases over time it is a sign
of a condensation reaction of the PET increase in the chain length or branching which
results in a higher viscosity. The samples for the rheological test were prepared with
material that had been recirculated for 15 minutes in the HAAKE MiniLab Micro
Compounder. The final pressure value can be correlated with complex viscosity |η*| of a
dynamic oscillatory test done with a rheometer. For the plain PET illustrated in Figure 5
after the loading peak, the pressure drop shows a decomposition of the PET. After 15
minutes, pressure is nearly constant with a value of approximately 18 bar. In Figure 6, the
frequency sweep for the same sample indicates that the loss modulus G" is significantly
higher than the storage modulus G'. The slight bumpy curve of G" is because of the fact
that the phase shift δ is nearly 90° and the smallest changes have big influences on G".
The complex zero shear viscosity |η*| is 200 Pas.
Figure 7 shows the PET with 1% 1,2,4-Benzenetricarboxylic anhydride after the loading
peak, a pressure increase which correlates with the condensation reaction of the PET. After
15 minutes, the pressure continues to increase with a value of approximately 15 bar.
Compared to the plain PET it is marginally less of an indication of a lower viscosity.
A look at the frequency sweep in Figure 8 for the same sample illustrates that G' and G" are
getting closer. This goes along with a lower δ of approximately 85° at low frequencies. The
PET gains more elasticity. The |η*| is 150 Pas at low frequencies. Compared to the plain
PET, the additive is accountable for the lower pressure and the lower |η*| on the one hand,
but on the other hand the additive brought about a reaction of the PET.
meta-Dioxazolinebenzene.
Conclusion
The HAAKE MiniLab Micro Compounder is a suitable instrument to screen the effects of
various additives. Only a small quantity of sample (7 g) is needed. Just viewing the
pressure dependence provides a first revelation of the functionality of the additives. The
time required for one test is reasonable. If additional rheological or other physical tests
have to be done, the transfer of the polymer melt into the HAAKE MiniJet Pro System is
possible. Different test specimens can be prepared rapidly and are reproducible. Additional
studies on the molecular weight and distribution, either by comprehensive rheological tests
for example Time Temperature Superposition compared with GPC data, could really
establish the assumptions.
This information has been sourced, reviewed and adapted from materials provided by
Thermo Fisher Scientific – Materials & Structural Analysis.
For more information on this source, please visit Thermo Fisher Scientific – Materials &
Structural Analysis.
ARL EQUINOX 100 X-ray Nicolet iS50 FTIR Spectrometer HAAKE MiniLab 3 Micro Compounder
Diffractometer