You are on page 1of 26

Melt Flow Rate TestingPart 1

Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


259
From: Plastics Technology
Issue: July 2013
Though often criticized, MFR is a very good gauge of the relative average
molecular weight of the polymer. Since molecular weight (MW) is the driving
force behind performance in polymers, it turns out to be a very useful number.
Click Image to Enlarge

Melt indexers, or extrusion plastometers, are common lab tools used to determine
the melt flow rate (MFR). The test, while often disparaged, gauges the relative
average molecular weight (MW) of the polymer. Since MW is the driving force behind
performance in polymers, it turns out to be a very useful number. (Photo: Tinius
Olsen)
Melt flow rate testing can be characterized as the Rodney Dangerfield of material test
methodsit gets no respect. People from all parts of the industry downplay the value
of the test or denigrate its usefulness outright. Those who instruct industry
professionals on processing are quick to point out that the melt flow rate (MFR) value
for a material is a single point on a curve that characterizes viscosity as a function of
shear rate. Since plastics are non-Newtonian, their viscosity varies with shear rate.
The melt flow rate test moves the molten material at a single flow rate, therefore a
single shear rate, and fails to capture the full range of the behavior of a material as a
function of changing shear rate. To make matters worse, the shear rate is not even
controlled. While the load on the material, or the shear stress, is constant during the
course of the test, the shear rate is an output of the test. The MFR itself is a reflection
of the shear rate used during the course of the test, but it is a result of the test and
not a controlled input.
Academicians do not like the units. How do you get from grams/10 minutes to
anything meaningful in terms of fundamental polymer behavior? I have written
extensively on the relative relationship between MFR and average molecular weight.
This has earned me occasional irate e-mails from university professors insisting that I
explain how you convert from g/10 min to the proper units for molecular weight,
which are grams/mole. It is a good question and one that we will answer as
completely as possible over the course of these next few articles.
In fact, melt flow rate testing is a poor tool for gauging processability, for reasons that
we will explain in detail later in this series. But it was never intended to be a
measurement of processability; this is an interpretation that has been applied by

some in the processing community. The notion that a melt flow rate tester is some
type of poor mans capillary rheometer is fundamentally incorrect. In addition, the
relationship between MFR and average molecular weight is a relative one. There are
a lot of factors that can skew this relationship and make interpretation tricky.
For example, adding ingredients to a compound such as glass fibers, impact
modifiers, and certain additives can alter the MFR of a material without changing the
average molecular weight of the polymer one bit.
But if the test is so useless, why does the value appear on so many material data
sheets? Not only is it a line item on the majority of published data sheets, it is often
the key characteristic that distinguishes one grade of material from another within a
given polymer family. In materials as diverse as polycarbonate, acetal, and
polystyrene, the MFR may be the only value on the data sheet that varies
significantly from grade to grade within a particular product offering. The reason is
simple. Assuming that all other factors are kept constant, the MFR is a very good
gauge of the relative average molecular weight of the polymer. Since molecular
weight (MW) is the driving force behind performance in polymers, it turns out to be a
very useful number.
Flow rate in a polymer is related inversely to viscosity. High-viscosity materials flow
with greater resistance and therefore more slowly under any particular set of
conditions than low-viscosity materials do. Therefore, higher-MW polymers have
lower MFR values and lower-MW polymers have higher MFR.
Practitioners of injection molding prefer the latter because it is easier to fill
demanding flow paths in a mold with what we refer to as high-flow materials.
Extruders and blow molders are more likely to prefer higher-MW materials because
they provide higher melt strength, a factor that makes it easier to control the shape of
a parison or a complex profile, die-swell considerations notwithstanding. End users
prefer higher-MW polymers whether they know it or not, because higher MW
correlates with better product performance. Impact resistance, fatigue performance,
environmental stress-crack resistance (ESCR), and barrier properties (to name a
few) all improve with higher MW.
My first exposure to the importance of MW as a material-selection criterion came on
the molding floor over 30 years ago. We were molding traffic-light signal housings out
of a high-MW injection grade of polycarbonate. The nominal MFR of the material was
5 g/10 min. The geometry of the part, coupled with the age of our molding machines,
made this a very challenging part to produce. Because the performance of the part
was quite critical, we conducted a falling-dart impact test once every hour to make
sure that the process was under control. At the conclusion of the run we would
randomly select 20 parts from the lot and repeat the impact test. The typical result
was that 20 out of 20 parts passed.
One day we decided to see what would happen if we used a lower-MW grade. The
reasoning went something like this: If we can use a material that flows more easily,
we can reduce the melt temperature of the material and the pressures associated
with injection and packing. This will reduce the stress on the material and improve the
impact performance of the part or at least compensate for the reduced impact
strength of the lower-MW polymer.
When we sampled a grade with a nominal MFR of 10 g/10 min we did, in fact,
observe that we could reduce the melt temperature of the material by 40 F (22 C),
and our pressure during first-stage filling declined by 10%. But when we ran our
impact performance evaluation on 20 parts made from this material, only four of them
survived the test. This large change in behavior occurred despite the fact that the
notched Izod impact values published on the data sheets for these two grades were

the same. This disparity between actual performance and the expectations created
by the data sheet occurs every day in our industry, and we will address the reasons
for this in a later article.
In our next article we will describe the MFR test procedure and discuss some of the
strengths and weaknesses of the test. We will also explain the reason why so many
material suppliers use the property, not only as a published line item but also as a
key parameter for quality certification from lot to lot.

Melt Flow Rate Testing Part 2


Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


187
From: Plastics Technology
Issue: September 2013
To fully appreciate the strengths and weaknesses of the melt-flow-rate (MFR)
test it is important to know something about the way the test is performed.
Click Image to Enlarge

Because the viscosity of a polymer varies with flow rate, or shear rate, a complete
characterization of viscosity must allow for measurements at multiple shear rates and
the production of a graph that captures this relationship. A capillary rheometer can
provide such data, but not a melt-flow-rate tester. Here is a capillary rheometer
output for two polypropylenes.
In order to fully appreciate the strengths and weaknesses of the melt-flow-rate (MFR)
test it is important to know something about the way the test is performed. The
methodology is covered by ASTM D 1238, while the corresponding international
standard is ISO 1133. There are small differences between the methods, but they
essentially perform the same function. (Link to Part 1 top right of this page.)
Both establish the rate at which a polymer flows under very specific conditions
through an instrument with a very specific geometry. For the ASTM method, the

cylinder into which the material is loaded has a diameter of 0.376 in., and at the
bottom of the cylinder is a removable insert with an even smaller opening or orifice,
as it is usually designated. While a few materials are tested using a non-standard
orifice, the standard opening has a height of 0.315 in. +0.001 in. and a diameter of
0.0825 in. +0.00025 in.
The tolerances here suggest that the dimensions of the flow path are considered to
be criticaland they are. The MFR instruments come with a go/no-go gauge for the
orifice diameter that must be used regularly to ensure that the opening is within
specification. Orifices are available in different materials and some are more durable
than others, particularly against the effects of aggressive cleaning that can enlarge
the opening. In addition, cleaning the top and bottom surface of the orifice can
reduce the height. These are factors that can detract from the accuracy of the
measurements.
Assuming the physical equipment is in good condition, the test involves first
establishing the appropriate temperature for the material being tested. The
prescribed temperatures are polymer-specific: Polycarbonate (PC) is typically tested
at 300 C, polyethylene at 190 C, etc. For some polymers there are two or even three
recognized conditions that may be used and it is usually the resin supplier that
decides which one it uses. Probably the best example of this is ABS, where one of
three different temperatures can be used to perform the test. So when performing
incoming quality-control tests it is important to use the same conditions as your
material supplier if you expect to get the same results. Just as with orifice geometry,
temperature calibration is very important.
The other key input parameter is the mass that is placed on the material sample once
it has been loaded into the cylinder and brought up to the specified temperature. This
is also a polymer-specific setpoint and it can be a single number agreed upon by all,
such as 1.2 kg for PC, or it can be one of two or three values, as in the case of ABS.
For each of the three temperatures that can be used to test ABS there is a particular
mass that accompanies that temperature. You can use 200 C with a 5 kg load, 230 C
with 3.8 kg, or 220 C with 10 kg. And, yes, each condition will give you a different
result for the same compound. The actual MFR value provided by the test is
expressed in grams/10 minutes and is governed by the test conditions and the
composition of the material being tested.
Here is the important fundamental regarding this test: The load is the input that drives
the material through the orifice; the flow rate is the output. Consequently, the MFR
test is a constant-shear-stress test, not a constant-shear-rate test. By definition it is a
pressure-limited configuration. In this respect it is different from a capillary rheometer,
a device for measuring viscosity that can control and vary the flow rate of the test
while measuring the force required to achieve that flow rate.
Capillary rheometry is a controlled-shear-rate test and can provide a true
measurement of viscosity, or resistance to flow. And because the viscosity of a
polymer varies with flow rate, or shear rate, a complete characterization of viscosity
must allow for measurements at multiple shear rates and the production of a graph
that captures this relationship. Figure 1 shows a capillary rheometer output for two
polypropylene materials.
This is where one of the common criticisms of the MFR test comes in. Critics point
out that while capillary rheometry provides a complete picture of the relationship
between viscosity and shear rate across a wide range of conditions that mirror many
different processes, the MFR captures only one point on the curve.
This is true. The question is, So what? This critique implies that the MFR test was
designed to provide some indication of processability. That was never its primary

purpose. Instead it is intended to provide a simple way of measuring the relative


average molecular weight (MW) of the polymer. As MW decreases, MFR increases.
Since MW weight drives performance in polymeric materials, this is something that
should be of interest.
When a polycarbonate supplier creates a range of grades distinguished primarily by
their MFR, it is identifying the products according their average MW. Molecular weight
influences impact performance, fatigue resistance, creep resistance, environmental
stress-crack resistance, and barrier properties. The higher the MW, the better the
performance. When a supplier establishes a specification range around a published
nominal MFR value for a grade, it not doing so because it is concerned with how the
material will process in a particular piece of equipment. The supplier monitors the
MFR because it knows this is an indicator that the average MW of the material is
under control and within the agreed-upon range.
The actual MFR value certainly has implications for processing. No one would
pretend that a polycarbonate with an MFR of 4 g/10 min will flow as far through the
same flow path and under the same molding conditions as one with an MFR of 20.
But the real difference in viscosity is not as large as these numbers would suggest,
because the difference in the flow rate automatically means a difference in shear
rate, and with higher shear rates come lower viscosity. The shear rate at which an
MFR test is performed is actually proportional to the MFR value itself. Multiplying the
MFR by approximately 2.2 will give the shear rate at which the test was performed.
So a 4-melt material is tested at a shear rate of about 8.8 sec-1 while the 20-melt
material is tested at 44 sec-1. Not only are these shear rates different, they dont
represent the shear rates experienced by polymers under most melt processes.
In the next article we will explore the quantitative relationship between MFR and MW,
the utility of using low shear rates to measure MW, and the reasons why some
processors are convinced that variations in their process are caused by lot-to-lot
variation in the MFR of the material.

Know How
Melt Flow Rate TestingPart 3
Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


167
From: Plastics Technology
Issue: October 2013
There is a well-established relationship between something called the weightaverage molecular weight of a polymer and a parameter known as the zeroshear viscosity.
Click Image to Enlarge

FIG. 1. You cannot measure viscosity at a shear rate of zero because viscosity is a
measurement of resistance to flow. In order to measure resistance to flow you have
to make the polymer flow, and as soon as you do so the shear rate becomes some
non-zero value. However, logarithmic plots of viscosity vs. shear rate tend to level off
as the shear rate approaches zero, so liberties are taken to a extrapolate the curve
back to the y-axis to arrive at the value for zero-shear vis

FIG. 2. This graph shows an actual viscosity/shear rate curve for a commercial acetal
copolymer with a nominalMFR of 9 g/10 min. The measurements cover shear rates
from 1.4 to 1400 sec-1 and illustrate the early portion of the behavior shown in Fig. 1.

FIG. 3. The arrows show the shear rates at which the MFR are measured for the two
materials. Because the shear rate for the test performed on the 22-MFR material is
significantly higher than that used for the 4-MFR material, the differences in viscosity
are exaggerated. This is an admitted imperfection of the MFR test. However, it also
illustrates the principle that differences in viscosity caused by variations in molecular
weight are most effectively measured at low shear rates.
A couple of years ago I received an indignant e-mail from a gentleman challenging
my statement that melt flow rate measurements usually provide a good assessment
of the relative average molecular weight of a polymer. How, he demanded, do you
convert a flow-rate measurement expressed in units of grams/10 minutes, to a
measurement of molecular weight, which is given in grams/mole? Researchers have
addressed this question.
There is a well-established relationship between something called the weightaverage molecular weight of a polymer and a parameter known as the zero-shear
viscosity. While the exact relationship is somewhat dependent upon the polymer, the
general equation that can be found in the literature is:
ho = kMw3.4
where ho is the zero shear viscosity, Mw is the weight-average molecular weight, and
k is a constant that is specific to the polymer being evaluated. The exponent 3.4 does
not apply universally; however, the values do tend to fall in a range between 3.2 and
3.9. In any event, this relationship clearly shows that relatively modest changes in
molecular weight (MW) result in large changes in melt viscosity when this viscosity is
measured at a very low shear rate.
Zero shear rate is a concept only a mathematician could love. In practice you cannot
measure viscosity at a shear rate of zero because viscosity is a measurement of
resistance to flow. So to measure resistance to flow you have to make the polymer
flow, and as soon as you do so the shear rate becomes some non-zero value. But
logarithmic plots of viscosity versus shear rate tend to level off as the shear rate
approaches zero, so we can extrapolate the curve back to the y-axis to arrive at the
value for zero-shear viscosity (see Fig. 1).
Figure 2 shows an actual viscosity/shear rate curve for a commercial acetal
copolymer with a nominal melt flow rate (MFR) of 9 g/10 min. The measurements

cover shear rates from 1.4 to 1400 sec-1 and illustrate the early portion of the
behavior shown in Fig. 1. The viscosity does not change by a statistically significant
amount at shear rates between 1.4 and 7 sec-1 before entering what is referred to as
the non-Newtonian region of the curve. Figure 2 also pinpoints the shear rate at
which the MFR test is performed on this material, about 20 sec-1. It can be seen that
the MFR test is performed at a point very close to the plateau leading back to the
zero-shear point on the curve.
This illustrates the usefulness of the MFR test in providing a relative measurement of
the average MW of a polymer. Because the MFR test is performed at relatively low
shear rates, the results approximate the zero-shear viscosity, despite the fact that the
shear rate is not controlled. The fact that the shear rate is not controlled in an MFR
test means that the test actually exaggerates the real differences in melt viscosity
that occur as a function of molecular weight.
This can be shown by revisiting the viscosity vs. shear rate curves for the two
polypropylenes shown in last months column. The MFRs for the two materials are
often interpreted as indicating that the viscosity of the 4-MFR material is 5.5 times
greater than that of the 22-MFR resin.
However, if we compare the viscosity of these two materials at any specific shear rate
we can see that this is not the case. The largest difference in viscosity between the
two materials is observed at the lowest shear rate, and even at this point the ratio of
the two viscosities is not quite 3.5:1 (3460/1005=3.44).
Note that as the shear rate becomes higher, the difference in viscosity between the
two materials decreases. At the highest shear rate on the graph of 10,000 sec-1 the
ratio has declined to 18/13 or 1.38:1. The faster the materials flow the more similar
their viscosities appear to be. This is caused by orientation of the long polymer chain,
a phenomenon often referred to as shear thinning. This behavior allows processors
to move polymer melts long distances through relatively restrictive flow paths,
particularly in injection molding.
The reason that the MFR values suggest a larger difference than the viscosity
measurements indicate is because the shear rate for the two MFR tests is not the
same. The arrows in Fig. 3 show the shear rates at which the MFR tests are
conducted for the two materials. As mentioned before, these shear rates are not
controlled, they are a result of the rate at which the two materials flow under a
constant load. But because the shear rate for the test performed on the 22-MFR
material is significantly higher than that used for the 4-MFR material, the differences
in viscosity are exaggerated. This is an admitted imperfection of the MFR test.
However, it also illustrates well the principle that differences in viscosity caused by
variations in MW are most effectively measured at low shear rates. As the shear rates
get higher, materials with substantially different MW start to look very similar.
At the extremes, commercial raw materials can be found with MFR as low as 0.1 g/10
min and as high as 500 g/10 min. However, for the vast majority of commercial
compounds where an MFR is published, the values fall between 1 and 50 g/10 min.
This means that the shear rates of MFR tests typically fall between 2.2 and 110 sec1.
The exaggerated differences in melt viscosity that the MFR test produces have an
unfortunate effect on the processing community, particularly those processors
involved with injection molding. Processors tend to think of the MFR values as being
representative of the real differences in the flow of the materials.
When molded parts fail to perform as expected, one of the changes that can be
made to improve performance is to increase the MW of the polymer being used to
produce the parts. This involves changing to a lower MFR material. Often this change

is dismissed out of hand by the processor because of the anticipated increase in


processing difficulty suggested by the difference in the MFR values. The viscosity
curves show that the real differences in viscosity are not as large as suggested by
the MFR values.
But it is difficult to get people to think beyond the MFR numbers. Many performance
problems that could have been solved with higher-MW materials have instead been
approached with complicated changes in part design and process conditions simply
because the MFR numbers looked too intimidating.
Next month we will explore the reasons for the perception of MFR as a barrier to
processing and explain why processes that exhibit sensitivity to lot-to-lot fluctuations
in MFR have fundamental problems related to the molding machine or the way the
process is set up.

Melt Flow Rate TestingPart 4


Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


121
From: Plastics Technology
Issue: November 2013
Few molders perform the test in-house. Of those that do, most don't
understand why they are doing it or what they are measuring.
Click Image to Enlarge

Is your injection machine running pressure-limited? If not, normal lot-to-lot variations


in resin MFR probably should not affect the process or product quality significantly.
While melt flow-rate testing is not extremely difficult, and the equipment is not very
costly, few molders perform the test in-house. Of those that do, a relatively small
percentage of them understand why they are doing it or what they are measuring. As
we have mentioned previously, most material suppliers specify and control their
products to a melt flow-rate (MFR) specification because they know that it is related
to the average molecular weight (MW) of the polymer they are producing. If the MFR
is consistent, the average MW is consistent.
But many molders who check their incoming materials to verify the certifications they
receive believe that they are testing to ensure process consistency. This belief is
based on a poor understanding of the relationship between the MFR value and the
actual viscosity of the material at processing conditions.
The capillary rheometry curves we have shown in previous columns (see Editor's
Pick top right) illustrate that the viscosity of a polymer is dependent upon the shear

rate. We also know that a component of the shear rate is the volumetric flow rate of
the material. The other component is the geometry of the flow path, which should be
a constant for any given mold. Of course, the shear rate varies with location in the
flow path. It is different in the runner than it is at the gate, for example.
But for any given location within the flow path there is the expectation that the crosssectional area of the flow path remains the same. Therefore, the only process
variable that can influence the shear rate is the flow rate of the material as it is filling
the mold. Flow rate is controlled by a machine parameter known to processors as
injection speed. On todays machines the technician selects the injection speed by
finding the appropriate screen and entering a setting that is usually given as a linear
distance per unit time such as in./sec or mm/sec.
Lets assume for a moment that the setup document for a particular mold running in a
particular machine calls for a first-stage injection speed of 3 in./sec. If the stroke
distance from the beginning of the injection process to the point at which the machine
transfers from first to second stage is 6 in., then the fill time should be 2 sec. Very few
processors ever check to see if this is actually the case, even though fill-time clocks
accurate to two or three decimal places are standard equipment on modern molding
machines. Processors tend to enter the setpoint and trust that the machine is doing
what it is told to do.
There are a number of factors that could prevent the machine from responding the
way it should, but in the end these all boil down to a lack of pressure needed to
achieve the desired velocity. Processes that run without an abundance of first-stage
injection pressure are called pressure-limited.
Viscosity is defined as resistance to flow. It can be thought of as the product of the
pressure applied to the fluid and the time over which this pressure is applied. To
move a fluid of higher viscosity a certain distance, it is necessary to either use more
pressure over a particular period of time or apply the same pressure over a longer
period of time. The SI units for viscosity, Pascal-seconds (Pa-s), reflect this
relationship.
If a molding machine is to run in a velocity-controlled mode it must be set up to
provide more pressure than is needed to deliver the desired volume of material to the
mold in a fixed time. In this way, when the viscosity of the material increases due to a
change in molecular weight, the pressure will increase proportionally but the time will
remain the same. If the time needed to deliver the material remains the same, then
the shear rate remains the same and will limit any change in viscosity to the inherent
effect of the higher MW. The latter contribution is relatively small when the material is
flowing at velocities typical of first-stage injection.
If the pressure is limited, either by the design of the molding machine or by the
manner in which the machine has been set up, then when the material viscosity
increases, the time required to deliver the material to the mold becomes longer. In
other words, the injection speed slows down all by itself. A slower injection speed
translates to a lower shear rate and this exaggerates the difference in the viscosity of
the materialjust like the MFR test does.
The MFR instrument is pressure-limited. A constant load is applied and the operator
simply observes the behavior of the material. If it is true that the molding machine in
the plant notices the difference in the MFR of the material, it is because the
machine is set up like the MFR tester, to be pressure-limited.
To put it another way, a velocity-controlled process minimizes the effect of changing
viscosity on the stability of the process, while a pressure-controlled process
exaggerates this effect. Normal MFR variations only influence the process if the
machine runs with a limit on the available pressure.

Lot changes in raw material are not the only possible cause of a change in the
viscosity. If regrind is blended into the virgin material, then fluctuations in the amount
of regrind or changes in the MW of that regrind can cause variations.
In addition, many polymers exhibit changes in melt viscosity due to variations in their
moisture content. For example, a 30% glass-filled PET polyester dried to 50 ppm
may have a viscosity 10-15% higher than that same material dried to 200 ppm. Both
of these materials are considered to be adequately dry for most applications, but they
may process differently, depending upon how the machine is set up. This effect is
even more pronounced in nylons.
That is the reason for a lot of the folklore in the industry about nylon materials being
too dry.
In short, if the molding process is properly established, typical fluctuations in MFR for
a given grade of material should have no significant effect on the process or the parts
coming out of that process, because the conditions of the MFR test do not resemble
those of the molding process until you reach the pack and hold phase. If you are
doing things correctly, by that point your mold cavity or cavities should be almost (but
not completely) full. A robust process, properly established in a capable molding
machine, should handle changes in viscosity much larger than the typical lot-to-lot
variation observed in a given grade of material.
So if the MFR test is not intended as a gauge of processability, then what is it
supposed to be used for? Next month we will address this topic.

Know How
Melt Flow Rate TestingPart 5
Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


116
From: Plastics Technology
Issue: December 2013
There are two points in the manufacturing supply chain where a determination
of average molecular weight (MW) is important. The first is when the material is
first received by the molder. The second is after molding.
How should melt-flow-rate (MFR) testing be used? There are two points in the
manufacturing supply chain where a determination of average molecular weight
(MW) is important. The first is when the material is first received by the molder. The
second is after molding.
Checking the incoming material is useful because it establishes that the material was
produced within the specified range for the MW of the product. If a certification is
provided with each lot of material, one of the few properties that is likely to be listed is
the MFR for that lot, along with a minimum and maximum specification value. Most
molders simply check the certification and file it, assuming that the value is correct.
However, there are several good reasons for verifying the number in-house.
First, the MFR value published on the certification is often a snapshot from one point
in a production run that may last for hours or even days. As with any manufacturing
process, there is variation in resin polymerization and compounding, and the sample
that was tested to provide the number on the certification may not be representative
of the material that is received by the molder.

Second, routine checking tends to create an open line of communication between the
molder and the material supplier or distributor. This is always useful when trying to
resolve problems or optimize the consistency of a material. I have had the
experience of working with material suppliers to narrow a MFR specification range to
improve the consistency of molded part performance.
That was 20 years ago, and the responsiveness of resin suppliers today may not be
what it was then. But the point is that if we had not been collecting data in-house and
correlating it to the feedback we were getting from our customers about the
functionality of the molded parts we were shipping, we would not have known what to
ask for.
Third, accuracy is always a concern for those who are new to MFR testing. Being
able to check in-house results against certification values or a certified sample of
material helps the people performing the tests build confidence in their technique.
While these tests focus on the quality of the raw material, the tests on the polymer in
the molded parts is a reflection of the quality of the molding process. Small changes
in the MFR of the polymer during the molding process are to be expected. While
there are exceptions, the MFR typically increases during processing, indicating that
the MW is decreasing under the influence of the heat and shear of processing. There
is little consensus regarding the allowable increase in MFR between raw material and
molded part, although the numbers typically quoted fall into a fairly narrow range,
between 20% and 50%.
The largest amount of work done to answer this question more precisely was
performed back in the 1970s by the major material suppliers. These companies at the
time were the source for all of the good research and advice regarding material
properties and processing. By correlating property retention to the change in the way
the material flows at the low-shear conditions of the MFR test, researchers came up
with guidelines for processors.
However, at that time much of the work was done by measuring the actual melt
viscosity rather than the MFR. This is a subtle but important difference. While the
MFR tester is not an appropriate instrument for making true measurements of
viscosity, it is possible to correlate the flow rates measured in a capillary rheometer to
those observed in a MFR tester. By establishing an empirical equivalence, the
researchers created instrument constants for the MFR tester that could be applied to
various materials in order to report the result using viscosity units. To this day, the
certifications for some commercial grades of raw material are still reported in
viscosity units (poise) rather than a true flow rate (grams/10 minutes).
Using these viscosity measurements, the upper limit for an allowable change from
raw material to molded parts was associated with a viscosity reduction of 30%. This
has often been restated as a maximum allowable increase in MFR of 30%. However,
these two statements are not equivalent. MFR is the reciprocal of viscosity. A 30%
viscosity reduction equals a 42.9% increase in MFR. This has been rounded off to
40% to provide a safety factor. This change translates to a 10% reduction in the
polymers weight-average MW.
People tend to adjust these guidelines in a somewhat self-serving way. Material
suppliers will often quote values of 25-30% for maximum recommended increase in
MFR during processing, in order to shift the burden to the processor. Some
processors will lobby for a number as high as 50% to give themselves a little more
allowance for the things that happen during processing.
And one large OEM holds its molders to a maximum allowable increase of 20% and
sets this as an acceptance criterion for all shipments. The OEMs argument is that if
the number exceeds 20% it begins to see failures during its qualification tests. Yet the

change in MW associated with a 20% MFR increase is quite small. If parts fail under
these circumstances, it is a reflection on the part design and perhaps the nature of
the qualification tests, not the process.
As with all guidelines, it is easy to over-interpret their significance. By setting a 40%
upper limit for MFR increase during processing, we are not saying that if the value
reaches 41% all the parts will fail, and if it stays at 39% everything will work as
planned. The functionality of a product always depends upon a complex interaction of
design, material properties, processing, and application conditions.
But what can be demonstrated is a decline in material properties as the change in
MFR due to processing increases. It is common to receive two parts, one identified
as good and one that has failed. Tests often show that both parts display an
excessive increase in MFR but the change for the part that is still working properly is
much smaller than for the one that has failed.
This simply shows that the part design is robust enough to compensate for some
level of degradation. But the fact that the MFR of the two parts can vary so much is
also an indication that the process conditions used to produce the parts are not
consistent. This is typical; you cannot manage what you dont measure. And too few
processors measure MFR before or after processing.

Know How
Melt Flow Rate TestingPart 6
Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


109
From: Plastics Technology
Issue: January 2014
Once degradation has been established, the discussion inevitably turns to how
it happened. It might be expected that the answer to this question is widely
known. Not so. The good news is that the influences that cause polymer
degradation during processing are few.
Click Image to Enlarge

Improper drying is a major culprit in MFR and molecular-weight changes during


processing. Inadequate moisture removal and excessive drying temperatures can
cause degradation in susceptible polymers.
While last months article was being written, tests were being run on two sets of
polycarbonate parts. The results demonstrated exactly the situation that we outlined
in that article. Both the good and bad parts exhibited an excessive increase in
melt flow rate (MFR) that is typically an indicator of polymer degradation. The good

parts were judged to be good because they had performed well in the application.
The bad parts showed obvious cracking that alerted the end user to the problem.
The molder had not provided a sample of the raw material but it had provided
documentation showing the specific grade of material that was used in both runs. The
raw material had a nominal MFR of 8.5 g/10 min. The good parts gave a MFR of
22.3 g/10 min, or an increase of 159%, while the cracked parts produced a melt flow
rate of 66.4 g/10 min, an increase of nearly 670%. These results both exceed the
benchmark boundary of 40% MFR increase during processing, but obviously the
cracked parts were molded from a material that had undergone a lot more abuse.
When presented with results like this, it is easy to point the finger at the molding
process. This assumes that the raw material used to produce the parts was, in fact,
as represented. It also assumes that the material was made to the published
specification. The value of 8.5 g/10 min represents the published nominal value, not
necessarily the value for the lot of material in question. However, if made properly, a
grade of material with a nominal value of 8.5 g/10 min. should not exceed 10.5 g/10
min. This still places both of these results well outside the appropriate range.
Once degradation has been established, the discussion inevitably turns to how it
happened. It might be expected that the answer to this question is widely known. But
the fact that it is asked as often as it is suggests otherwise. The good news is that the
influences that cause polymer degradation during processing are few.
The universal concern for all polymers is the combined effect of heat and time while
the material is in the melt state. Polymers are typically organic and they have a
limited tolerance for the temperatures and shear stresses associated with molding
processes, particularly injection molding. Polymers like polyethylene and
polypropylene will withstand the processing environment quite well as long as the
antioxidant package in the material holds out.
However, polymers such as polyesters and nylons are more susceptible to thermal
degradation. The rate at which this degradation occurs depends upon the actual melt
temperature of the material in the injection cylinder and the time that it spends there,
also known as the residence time.
If you ask material suppliers about the residence time they prefer for their products,
you will get answers that typically fall in the range of 5-10 min, depending upon the
polymer. Unfortunately, because molds often run in machines with barrels that are
larger than would be considered optimal, residence times often exceed these
recommended values considerably.
The longest residence time I have personally witnessed is 51 min, and I have heard
of a situation where the residence time exceeded 60 min. At these time frames even
melt temperatures that remain well within the suppliers recommended range may not
ensure the survival of the material.
The other major culprit is moisture. This is not a universal concern. Some materials
such as polyethylene and polystyrene are non-polar and cannot absorb an
appreciable amount of water. Other materials like ABS and acrylic are hygroscopic,
but the water will only produce cosmetic problems, it will not chemically attack the
polymer. However, a small class of polymers that includes polyesters, polycarbonate,
and nylons actually react with excess moisture in a way that breaks chemical bonds
within the polymer chain, thus reducing the molecular weight of the polymer.
This process is known as hydrolysis and it can only be prevented by drying the raw
material to an appropriate level before the material is introduced into the heating
cylinder of the molding machine. For some materials the maximum allowable
moisture content for processing is as low as 200 parts per million (0.020%), and
optimal levels can be as low as 50 ppm.

As with most chemical reactions, hydrolysis occurs more rapidly at higher


temperatures. Therefore, in a situation where melt temperatures are elevated or
residence times are extended, improperly dried material will degrade even more
quickly than it would under optimal processing conditions. Many processors and end
users assume that thermal and hydrolytic degradation will be visually evident.
However, this frequently is not the case.
Thermal degradation may result in some discoloration, but colorants can mask this
effect, and dark colors such as brown and black can make color shifts undetectable.
Hydrolytic degradation is often associated with a visual defect known as splay or
silver streaking. But by the time splay appears on the part surface it is already too
late. Splay is often caused by volatiles produced when excess moisture in the
material boils. The moisture that causes the damage to the polymer molecule has
been consumed in the chemical degradation reaction and will not produce a visual
defect.
PET is an excellent example of this. PET can be molded with a substantial amount of
excess moisture and still produce parts with a good surface appearance because it
reacts with the water very efficiently. An alert processor may notice a reduction in the
viscosity of the material and the quality-control people performing mechanical tests
will certainly observe a decline in mechanical strength and ductility, but the parts will
look fine.
There is one more factor that can influence the molecular weight of a polymer during
processing and that is the very drying process used to remove the moisture from the
material. Since drying involves elevated temperatures, extended drying times can
potentially influence the integrity of the material, particularly if the drying
temperatures exceed supplier recommendations. This is unlikely to occur in most
amorphous polymers like polycarbonate, ABS, and acrylic, because the
recommended drying temperatures are usually just 20-30 F below the softening
point. Any significant excursion in the drying temperature will quickly produce a
sintered mass of material that will no longer flow through the hopper.
However, in semi-crystalline polymers, the recommended drying temperature may be
200-300 F below the melting point. In these cases, significantly elevated drying
temperatures can occur without producing any noticeable problems in conveying
resin through the material-handling system. But studies have shown that nylon
particularly those grades that are not heat stabilizedwill undergo an increase in
MFR if dried for extended times at temperatures that exceed supplier
recommendations.
These MFR changes occur even before the material is introduced into the heating
cylinder of the molding machine. The resulting brittle condition of parts made from
this degraded material is often referred to as overdrying. However, it has nothing to
do with the actual moisture content of the pellets. It is attributable to oxidation,
another chemical reaction that results in a reduction in the molecular weight of the
polymer.
Unfortunately, this factor has received far too much attention. It is a real concern with
nylons, and studies have shown that the oxidation process is driven by drying
temperature much more than by drying time. In most polymers, oxidation occurs at a
much slower rate than in nylons and can take days or even weeks, particularly if
proper temperature control is in place.
But this has not stopped material suppliers from including maximum drying-time
recommendations in their literature. Some of these recommended times are
remarkably short, as little as 8-12 hr, and they send an unfortunate message to
processors that damage to the polymer is likely if these drying times are exceeded.

There is no science behind these relatively new recommendations. When it comes to


preserving the molecular weight of a polymer, the risks of not drying enough far
outweigh those associated with drying for a longer period of time. In fact, differences
in the effectiveness of drying distinguished the good from the bad polycarbonate
parts mentioned at the beginning of this article. While the good parts could certainly
have been better, they were at least dried to a level that allowed for a functional part.
The bad parts were not, as evidenced by splay on some of those parts.
Our next article will look at the limitations of the MFR test in detecting resin
degradation and why it does not always work.

Melt Flow Rate TestingPart 7


Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


107
From: Plastics Technology
Issue: February 2014
Here is why the melt flow test is not universal.
Click Image to Enlarge

Two different PBTs were evaluated in the melt state over time. The viscosity of the
materials was measured periodically using a capillary rheometer operating at a
constant shear rate. One material shows excellent stability and the viscosity changes
very little over the duration of the test. But the other initially undergoes a rapid
decrease in viscosity typical of chain scission, followed by an increase in viscosity,
which has been attributed to crosslinking promoted by the oxidation that caused the
initial breakage of the polymer chains.

As useful as the melt flow rate (MFR) test can be, it is not universally applicable.
Fundamentally, the test is intended to document differences in molecular weight.
There are several factors that can make interpretation of the results difficult. The first
involves polymers that do not initially degrade through chain scission, the chemical
reaction by which the polymer chains become shorter.
PVC is a good example. Degradation in PVC begins with a process known as
dehydrochlorination, an impressive term for the loss of hydrogen chloride. As the
chlorine and hydrogen atoms separate from the polymer backbone, double bonds
form, producing a change in color. But the oxidation that initiates chain scission may
not begin until the loss of hydrogen chloride has proceeded to a significant degree.
Therefore, the polymer may have undergone considerable modification without giving
any evidence of a reduction in molecular weight.
A second problem is introduced by competing reactions that can increase and
decrease the molecular weight of the polymer at the same time. Polymers such as
nylons and polyesters can undergo a series of complex reactions that can produce
this type of behavior.
The accompanying chart shows a comparison of two different PBT polyesters
evaluated in the melt state over time. The melt viscosity of the materials was
measured periodically using a capillary rheometer operating at a constant shear rate.
One of the materials shows excellent stability and the viscosity changes very little
over the duration of the test. However, the other material initially undergoes a rapid
decrease in viscosity typical of chain scission followed by an increase in viscosity
which has been attributed to a process of crosslinking promoted by the oxidation that
caused the initial breakage of the polymer chains.
MFR tests performed on samples taken from various points along this time line would
produce very different impressions of the condition of the polymer. A sample taken
20-25 minutes into the evaluation would show significant polymer degradation, a
sample taken at 40 minutes would show no change, and a sample taken at 55
minutes would appear to show that the molecular weight had increased despite the
fact that a part molded from material in this condition would be extremely brittle.
Impact-modified materials can produce a similar effect for the same reason.
Butadiene rubber and EPDM rubber are both commonly used as impact modifiers.
Both of these materials are capable of undergoing crosslinking when exposed to
elevated temperatures. This can cause the MFR of these materials to actually
decrease during processing, a behavior that is frequently observed in materials like
ABS, PPE/HIPS blends, and impact-modified nylon.
Other constituents such as flame retardants can further complicate the response of
the material to processing. In general, the more ingredients there are in the
compound, the more difficult it becomes to make a straightforward interpretation of
the MFR test results.
One excellent example of this principle is the influence of fillers. Because fillers do
not melt at processing temperatures, they tend to increase the melt viscosity of a
polymer. This increase is detectable at any shear rate, but it is more noticeable at low
shear rates where the lack of orientation makes the filler particles or fibers much
more resistant to flow. The problem with comparing the MFR results of pellets and
parts when fillers are present pertains to the influence that the process has on the
integrity of the filler. This is especially true of reinforcements such as glass fiber. We
have established that processing can reduce the length of the polymer chain, and
this is the primary area of interest when we perform MFR tests on molded parts and
the related raw materials. But high-shear processes like injection molding also cause
fiber length to be reduced, and this reduction also contributes to the change in the

MFR measurement. Whenever MFR tests are performed on glass-filled materials, the
measurement is a composite of the response from the polymer and the fiber. As the
fiber content increases, the contributions from this ingredient become more
important.
It is common for fiber lengths to be reduced by 50% during molding. These shorter
fibers flow through the MFR tester more easily than the longer fibers in the original
raw material. Therefore, it is not possible to apply the 40% rule discussed earlier to
filled materials. Even at a 10% loading of glass fiber, the increase in MFR may be as
high as 75% before we begin to have concerns about polymer degradation. By the
time we get to a 30% glass content, the allowable MFR increase may be as great as
200%. The deviations between these benchmarks and the 40% value used to
evaluate unfilled materials are designed to make allowances for the effect of typical
fiber-length reduction on melt viscosity. The uncertainty becomes even greater when
working with long glass fibers, while particulate fillers like talc may introduce a
smaller variance from an unfilled product.
Ideally, we would not perform MFR tests on filled materials; we would instead employ
solution-based measurements that work by dissolving the polymer in a solvent and
removing the filler so that only the polymer is being considered. However, the MFR
test is so accessible and easily performed that many suppliers of filled materials still
use MFR as their quality-control parameter for average molecular weight. Therefore,
it is useful to have a way of allowing for the contribution of fiber-length reduction on
the MFR of a filled material.
Acetal, more commonly known these days as POM, presents a different problem.
MFR is often used to distinguish between grades of this material. However, it has
been shown that the MFR test is not useful for detecting process-induced polymer
degradation in acetal/POM. In fact, even much more sophisticated techniques such
as gel permeation chromatography (GPC) very often do not show any measurable
change in the molecular weight (MW) of this polymer.
This is because the rate of MW reduction in acetals is so rapid at processing
conditions that any polymer chains that begin to unzip reach a MW that is so low that
they no longer influence the melt viscosity of the polymer. There is no significant
amount of intermediate chain lengths in a degraded acetal. There are other analytical
techniques that can detect degradation in POM, but typical MW determinations such
as MFR, solution viscosity, and GPC are ineffective.
However, degradation of acetal that occurs in the solid state takes place slowly
enough to produce a substantial amount of intermediate chain lengths. A good
example is sustained exposure to hot water, particularly if the aqueous environment
is mildly acidic. In this situation the reduction in average MW is gradual enough to
produce the type of lower-MW constituents that the MFR test is designed to detect. In
one case, an acetal homopolymer exhibited a 100% increase in MFR from as-molded
parts to parts that had been immersed in 90 C water for 1000 hr.
While the vast majority of commercial thermoplastics list MFR on their data sheets,
some materials make no mention of this property. In our next installment we will
discuss the reasons for this and discuss some of the special problems related to
measuring melt flow in nylons.

Know How
Melt Flow Rate TestingPart 8

Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


98
From: Plastics Technology
Issue: March 2014
Here are the steps to take in cases where the MFR is not provided by the
polymer supplier.
Click Image to Enlarge

The relationship of MFR to moisture content for unfilled nylon 6 is nearly linear
across the range of 400 to 2000 ppm. Once the moisture content drops below 400
ppm the MFR values begin to decline more rapidly.
The number of commercial materials that employ a melt flow rate (MFR) value as
part of the published property profile attests tothe usefulness of the test. Virtually
every grade of material within many polymer families provides this data point, and
certifications will almost always list the MFR value for every lot. But there are some
polymers where the availability of a MFR value is inconsistent across the base of
suppliers, and there are a few polymer families where a MFR value is rarely provided
by any supplier.
The reasons for this vary. Some of it is convention. Amorphous PET, for example,
does provide a value that is related to molecular weight (MW). But instead of using
MFR, the manufacturers of these materials have chosen to use intrinsic viscosity (IV).
This test involves creating a dilute solution of the polymer in the appropriate solvent

and then comparing the flow rate of this solution to the flow rate of the pure solvent
through a piece of glassware of a standardized geometry. The greater the difference
between these flow rates, the higher the IV of the material and the higher the
average MW of the polymer. Performing the IV test is more complicated because of
the need for a delicate apparatus and the use of some very noxious chemicals.
However, this test can be conducted practically at room temperature. This removes
the need for drying the material as a preparatory step, something that is required
when testing in the melt state.
Filled materials are another example where MFR testing is frequently avoided. This
seems particularly true of filled and reinforced semi-crystalline resins such as
polyesters, nylons, and polyphenylene sulfide (PPS). In PPS, for example, where
virtually all molding and extrusion grades of the polymer are highly filled, it is
essentially impossible to find a MFR value on the data sheets. The one exception is
the unfilled materials that are used either in coatings or as the feedstock for the filled
compounds.
The avoidance of the MFR test for the filled materials is related to last months
column. The presence of the filler decreases the MFR. The value derived from the
test is more dependent upon the filler content and type and less upon the MW of the
polymer. The presence of the filler can also increase the variability of the test results,
making it difficult to arrive at a meaningful specification. However, the avoidance of
the MFR test in filled and reinforced materials is far from universal. Many suppliers of
glass- and mineral-filled materials, including polypropylenes and polycarbonates, still
use the MFR test as a quality-control metric and to provide values for their literature.
PPE/PS alloys represent an interesting example of how conflicted the industry can be
over the use of this test. Anyone who worked with GE Plastics (now Sabic Innovative
Plastics) in the 1970s and 1980s knows that their Noryl data sheets never used to
provide MFR values and instead relied upon capillary rheometer data. If you were to
question the technical people at GE about this they would tell you that the MFR test
does not tell you anything about PPO materials. But in the 1990s the MFR numbers
began to appear on their data sheets and today they are commonly found even for
some filled grades. Some other suppliers have followed suit while others do not
provide MFR values for their PPE/PS materials.
The polymer where there seems to be a general consensus on avoiding the use of
the MFR test is nylon. Regardless of the type of nylon, MFR tests are almost never
done and values are rarely provided on the data sheet. If any measure of MW is
provided, it usually comes in the form of a value known as relative viscosity. This is a
value based on the same type of solvent-based test used for intrinsic viscosity, but
the units in which the results are reported are often different.
Nylon presents some special challenges for the MFR test. Because it is a
hygroscopic material that can hydrolyze in the melt state, all samples of nylon that
are tested in the melt state must be dried before testing. This is also true of materials
like polycarbonate and polyester. But nylon creates additional problems because of
the way that the moisture that remains after drying influences the flow of the polymer.
Most materials that can hydrolyze must be dried to moisture contents of less than
0.020% (200 ppm). The viscosity of many of these materials is influenced by the
exact moisture content of the sample being tested.
However, between about 30 ppm and 200 ppm moisture content, the MFR of these
materials tends to be essentially constant. But unfilled nylons are considered to be
adequately dry once the moisture content is below 0.20% or 2000 ppm. It turns out
that across the range from 0-2000 ppm the viscosity of nylon varies significantly.
Figure 1 shows the relationship of MFR to moisture content for an unfilled nylon 6.

The relationship is nearly linear and it spans the entire range between 400-2000
ppm. Once the moisture content drops below 400 ppm the MFR values begin to
decline more rapidly.
Anyone who has ever molded nylon knows this. Very dry nylon has a higher melt
viscosity and presents challenges in filling long, thin flow paths. This is the reason
that suppliers of nylon advise processors to leave a little bit of moisture in the
material. Remember that when we perform the MFR test we are often comparing the
result for raw material to that of molded parts and we are looking for a certain
threshold in the change from pellets to parts to tell us if we have done a good job of
preserving the MW of the polymer. But the data in Fig. 1 show that the result
obtained from a single sample of material can vary by more than 40% simply as a
function of the moisture content of the sample. This makes obtaining a consistent
result very difficult and it makes interpretation of a pellet-to-part comparison quite
challenging.
There is a way around this. It involves creating a calibration plot such as in Fig. 1.
Then when the MFR test is performed, the moisture content of the sample is
measured at the same time and the result is then normalized to a particular moisture
content such as 0.10% (1000 ppm). Therefore, a MFR of 13.3 g/10 min measured at
1500 ppm would be reported as a normalized value of 12 g/10 min at 1000 ppm. A
value of 10.4 g/10 min reported at 400 ppm would also be reported as a normalized
result of 12.0 g/10 min at 1000 ppm.
In this case two tests that appeared to produce significantly different results are
shown to represent essentially the same values when the effects of moisture content
are considered. For most people this is just too much work. The alternative is to
perform the solvent-based test. But for those who do not wish to tackle the
challenges associated with the solution testing, the much simpler MFR test can still
be made to work even with nylons.
In our next installment, we will discuss the meaning of a related property that is
beginning to appear on an increasing number of data sheets; the melt volume rate.
We will review the relationship of this measurement to the melt flow rate and the
reasons for its increasing popularity.

Melt Flow Rate TestingPart 9


Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


75
From: Plastics Technology
Issue: April 2014
Understanding the value of melt-volume rate.
Click Image to Enlarge

Melt volume rate (MVR) may be a more valuable measure than the melt flow rate
(MFR). (Photo: Tinius Olsen)
If you spend a lot of time looking at data sheets you have probably noticed a trend
over the last several years in how the melt flow rate (MFR) is reported. Along with, or
instead of, the MFR, a new line item is appearing called the melt volume rate (MVR).
Whereas the MFR is reported as a mass per unit time, such as grams/10 min, MVR
is given as a volume per unit time, such as cm/10 min or, regrettably, in./10 min. I
say regrettably because one of the areas where the U.S. is out of step with the rest of
the world is in its use of the old inch-pound system of units as opposed to the metric
system or SI units. We have spent billions of dollars over the last 20 years to adopt
ISO standards for just about everything we do in industry, but we just cannot seem to
let go of our antiquated system of units, with notable exceptions such as the
automotive industry, where the change to SI units has been effectively mandated.
But there are a few areas where we have all become comfortable with the
alternatives and one of those is the use of grams as a measure of mass and g/cm as
a measure of density. When you go into most plants, if people are weighing parts
they are recording the results in grams. And when you look up the density of a
material on a data sheet the units are g/cm. So it is somewhat of a mystery as to
why we would go out of our way to backtrack to a unit of flow that uses cubic inches
instead of cubic centimeters. Imagine the confusion that would result if suddenly one
day we started publishing the density of polyethylene in lb/in. and a 0.955 density
HDPE became a 26.437 HDPE. Converting from in./10 min to cm/10 min is a
simple multiplication by 16.4.
If the material supplier provides both MFR and MVR data, once you make this
conversion you will notice that the numbers are similar but usually not identical. The
ratio of the two valuesthe MFR divided by the MVRis a measure of the melt
density of the material in g/cm. This is an important property because it is a different
number than the solid-state density that is typically provided on the data sheets.
For example, unfilled and unpigmented polycarbonate has a solid-state density of
1.20 g/cm while the melt density of the material is 1.08 g/cm, or 90% of the solid
state density. This value has practical importance. First, melt density is a property
input for flow-simulation software. It is also an important conversion factor for
determining the capacity of an injection unit on a molding machine. Many people in
the industry are aware of the fact that the shot sizes are given in terms of generalpurpose PS.
But when converting from PS to, for example, PE, the tendency is to multiply by the
ratio of the density values provided on the data sheet. So a 30-oz barrel would be
considered to hold 27.5 oz of 0.955 HDPE. But these are solid-state densities and
polymers are not solid when they are in the barrel, or at least we hope not. The melt
density of HDPE is only 80% of the solid-state density, so the barrel capacity in
HDPE is actually a little less than 24.5 oz.

There are some sensible reasons for converting to MVR. First, all of the equations in
rheology that contain a term for flow rate use a volumetric flow rate, not a mass flow
rate. At a more practical level, MFRreferred to by ISO standards as a melt mass
flow ratedoes not correct for the density of the material and therefore distorts the
real rate at which the material flows through the instrument. Denser materials appear
to be flowing faster simply because the material collected from the instrument is
heavier.
Consider two PCs, one white and the other black. The white pigment is likely to be
TiO2, a pigment with a density that is over three times higher than that of the
polymer. If it is added at a loading of 4%, it can raise the density of the final
compound by approximately 10%. The black pigment may be carbon black, which
has essentially the same density as the polymer and therefore does not influence the
density of the final compound. MFR tests that show a 10% higher value for the white
material would appear to suggest that the white material is of a lower average
molecular weight, when in fact the two materials would be identical when viewed in
terms of MVR.
For materials with a melt density of 1 g/cm the MVR and the MFR will be the same.
But this rarely happens. So if you have one supplier of a given material that lists only
MFR and the other lists only MVR, you must know the melt density of the polymer in
order to convert.
Fortunately, the MFR test, when run in a particular way, can provide this property.
The piston that is used to force the material through the orifice has two scribes on it
that are exactly 1 in. (2.54 cm) apart. The instrument can be set up to begin
monitoring the test when the piston travel reaches the first scribe line and the test will
end when the piston reaches the second line. Since the diameter of the cylinder is
closely specified at 0.95504 + 0.00254 cm, the volume associated with this
displacement can be readily calculated at 1.819 cm.
By weighing the extrudate that comes through the orifice during this displacement,
the melt density of any material can be calculated. For example, unfilled,
unpigmented PC has a melt density of 1.08 g/cm, so the mass of material displaced
by this 1-in. travel should be 1.965 g with allowances for small variations due to the
actual size of the bore and the precision of the operator performing the test.
We have spent a lot of time (nine monthly columns so far) on the topic of MFR
testing. We have one more step in this process and that is to review some limitations
of the method. Over the years there have been different attempts to address these
limitations. Some of these have not gained traction in the industry while others are
still in the early stages of development and show a great deal of promise in
expanding our understanding of the practical aspects of polymer flow. We will cover
those in the last installment of this series.
ABOUT THE AUTHOR
Michael Sepe is an independent materials and processing consultant based in
Sedona, Ariz. with clients throughout North America, Europe, and Asia. He has more
than 35 years of experience in the plastics industry and assists clients with material
selection, designing for manufacturability, process optimization, troubleshooting, and
failure analysis. Contact: (928) 203-0408 mike@thematerialanalyst.com.

Know How
MATERIALS: Melt Flow Rate TestingPart 10
Materials Know How

By: Michael Sepe from Michael P. Sepe LLC


79
From: Plastics Technology
Issue: May 2014
Two areas where the melt flow rate test it is not useful are related to
processing.
Click Image to Enlarge

Melt elasticity is most significant in low-shear conditions, such as extrusion and blow
molding, where it presents itself as die swell. It also accounts for cosmetic defects in
injection molding around gates and at weld lines.
Melt flow rate testing, like any method of evaluating material characteristics, is helpful
when it is used for its intended purpose. Two areas where it is not useful are related
to processing. One of these is the need for an accurate assessment of flow
characteristics under typical processing conditions. This is generally thought of
among processors as a measurement of viscosity.
Viscosity in polymers is dependent upon temperature and shear rate, and melt flow
rate (MFR) testing typically fixes the temperature at a constant value for a given
polymer and cannot control shear rate. Capillary rheometry is typically used to
capture the temperature dependence and shear-rate dependence of viscosity, and
curves of viscosity versus shear rate at multiple temperatures are required in flow
simulation software so that the actual inputs provided by the analyst can be
interpreted by the software. Applicable shear rates can vary from relatively low for
processes like blow molding to very high for processes like injection molding, so
characterizing behavior across a wide range of shear rates is important.
While the predictions made by simulation software continue to improve over time,
there are still some problems that have defied a precise solution. Part of the reason
for this is that even with the enhanced capabilities of a capillary rheometer, these lab
instruments do not replicate what is happening in many real-world processes,
particularly in a process involving a closed mold. The MFR test and capillary
rheometry both utilize an open system and a flow path where all of the material is
molten.
However, in the real world of injection molding, for example, a frozen layer begins to
develop along the flow path almost immediately and becomes thicker with time,
comprising a larger and larger percentage of the flow path cross-section as the
material cools. This means that the size of the effective flow path, which is the most
important term in the equations that describe polymer flow, is continually changing in
a manner that the lab instrumentation does not recognize or account for.
In addition, these instruments do not account for shear-induced flow imbalances that
can cause a temperature gradient to develop within the layers of flowing material.
Much of the industry has been in denial regarding this phenomenon since it was first
discussed by John Beaumont in the late 1990s, and there are important questions
that remain regarding its magnitude and its effect on processing and part quality. But
ongoing work makes it clear that this is a factor that must be considered in order to
arrive at a precise assessment of the way in which polymers flow. Even if the MFR
tester could be instrumented to measure this behavior, it cannot produce or control

the shear rates needed create the conditions where the phenomenon could be
measured.
The second shortcoming of lab instruments designed to characterize viscosity is their
inability to account for or measure melt elasticity. Most of us have heard at some
point in our careers that polymers are viscoelastic materials. This means that they
simultaneously exhibit the characteristics of elastic solids and viscous fluids, and the
compound response of any material is some combination of these two factors that
varies with temperature and shear.
Viscoelastic behavior gets a lot of attention when polymers are in the solid state in
their end-use environment. The viscous flow component, which appears to be a
minor contributor to overall performance, is a consideration in phenomena such as
creep, stress relaxation, and fatigue.
In the molten state, when most of us think of the polymer as a fluid, some remnants
of elastic behavior remain. The elastic characteristics tend to be most important in
high-viscosity systems and therefore appear at their greatest when melt temperature
and shear rates are low and molecular weight is high. In blow molding and extrusion,
melt elasticity manifests as die swell.
But it also appears in processes such as injection molding and plays a large part in
causing many of the cosmetic defects that plague processors, particularly around
gates and at weld lines.
An attempt was made in the early 1990s to develop an instrument that could
measure melt elasticity simply. Bryce Maxwell, one of the more creative minds from
the early era of plastics and a professor at Princeton, developed an instrument that
he called a melt elasticity index (MEI) tester. This device was intended to provide a
simple means of capturing the elastic properties of a melt just as the MFR tester is
designed to simply measure viscous flow. The MEI tester also captured the elastic
recovery properties, something that much more sophisticated rheological instruments
failed to accomplish.
It consisted of a heater that allowed the operator of the device to control the
temperature of a material sample that was positioned on a flat plate and heated to
the test temperature. A second disk-shaped plate was then placed on top of the
sample and pressed downward. As the polymer flowed, it formed a specimen of
controlled thickness between the two plates. A drive motor then rotated the upper
disk, shearing the specimen with respect to the lower plate. After a specified amount
of deformation was applied, the upper disk was released from the drive and it then
rotated in the opposite direction, propelled by the elastic recovery of the specimen. A
dial and pointer on the top disk provided a means of quantifying the material
response. This was observed and documented with a video camera.
Dr. Maxwell demonstrated the device at a few SPE ANTEC meetings, but it never
went through the rather lengthy vetting process needed to bring it into the world of
established laboratory instrumentation. It was, however, capable of distinguishing
between materials of different melt elasticity, even when the melt viscosity of the
materials appeared to be comparable, and it did so in a matter of minutes and at
conditions similar to those associated with some processing methods.
If developed, it would have likely represented the same type of lower-cost alternative
to parallel-plate or cone-and-plate rheometry that the MFR test represents when
compared with capillary rheometry. This would permit processors to make useful, if
approximate, measurements of melt elasticity that would provide another piece of the
puzzle that is polymer flow.

We have now completed our tour of the MFR test. Hopefully, we have been able to
clarify those things for which it is intended to be used and have dispelled some of the
myths surrounding the numbers that come out of the test.
However, I suspect that confusion will continue for many in the industry. Just as I was
finishing up this article, I received an e-mail from a compounder that provides
PC/ABS alloys to the automotive industry in India. The compounders customers are
complaining that sometimes the parts that they mold exhibit cracks in thin sections
and they claim that this happens when the MFR of the raw material declines from 6
g/10 min to 4 g/10 min. This assertion reflects a poor understanding of the
relationship between MFR and the manner in which polymers flow when subjected to
real-world processing conditions.

You might also like