You are on page 1of 23

TREATISE ON EPDM

P. S. RAVISHANKAR*
EXXONMOBIL CHEMICAL COMPANY, 5200 BAYWAY DRIVE, BAYTOWN, TX 77520

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)

ABSTRACT
EPDM is the designation given for a saturated polymer chain of the polymethylene type, consisting of ethylene,
propylene, and diene. It has been and continues to be one of the most widely used and rapidly growing synthetic rubbers.
Technology development has been driven by novel catalysts and processes that have expanded the range of products while
delivering improvements in manufacturing such as energy efficiency and environmental footprint. A broad overview of the
EPDM process and product technology is presented with focus on more recent developments. The range of topics includes
polymer chemistry and physics, characterization, applications, and the manufacturing process. Technology platforms based
on both the ZieglerNatta catalysts and metallocene catalysts are compared and contrasted. [doi:10.5254/rct.12.87993]

CONTENTS
I. Prologue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
II. Introduction to EPDM . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
B. Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
III. Polymer Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. Molecular Weight . . . . . . . . . . . . . . . . . . . . . . . . . . .
B. Molecular Weight Distribution and Long Chain Branching
C. Crystallinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Ethylene Content . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E. Composition Distribution . . . . . . . . . . . . . . . . . . . . . .
F. Sequence Distribution . . . . . . . . . . . . . . . . . . . . . . . . .
G. Solution Properties . . . . . . . . . . . . . . . . . . . . . . . . . . .
IV. Characterization Tools . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. Molecular Weight . . . . . . . . . . . . . . . . . . . . . . . . . . .
B. Molecular Weight Distribution and Long Chain Branching
1. Branching Index . . . . . . . . . . . . . . . . . . . . . . . . . .
2. Rheology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
C. Composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Crystallinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E. Sequence Distribution . . . . . . . . . . . . . . . . . . . . . . . . .
V. Compounding and Vulcanization . . . . . . . . . . . . . . . . . . . . .
A. Choice of Diene . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
VI. Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. Thermoplastic Vulcanizates . . . . . . . . . . . . . . . . . . . .
B. Wire and Cable Insulation . . . . . . . . . . . . . . . . . . . . . .
C. Oil Modification . . . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Automotive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Body Sealing . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2. Coolant Hose . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3. Molded Parts . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E. Roof Sheeting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
*Corresponding author. Ph: 281-834-5020; email: p.s.ravishankar@exxonmobil.com

327

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

328
328
328
329
330
330
330
331
332
332
332
333
333
333
334
334
335
336
336
336
337
337
338
338
339
340
340
340
341
341
341

328

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)
F. Miscellaneous Applications
VII. Manufacturing Process . . . . .
VIII. Epilogue . . . . . . . . . . . . . .
IX. References . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

341
342
346
346

I. PROLOGUE
First produced industrially in the early 1960s, EPDM has experienced impressive growth over
the past five decades. This growth has been supported by a cycle of polymer substitution, which has
led to a need for better balance of properties, which in turn has led to innovations in catalyst and
process development. It is to be expected that a polymer of such vintage would have been reviewed
by many over the years. This indeed is the case.18 The current author has a unique background over
his career, which combines expertise in both process and product. This review article, reflecting the
authors background, is being presented as a process/product treatise on EPDM.
The review begins with the product section in which the polymer physics related to the
chemical structure of EPDM is briefly discussed. This is followed by a discussion of the
characterization tools and end-use applications as they relate to polymer structure. Detailed review
of EPDM compounding and vulcanization can be found elsewhere.6 In the process section, two of
the major manufacturing platforms in use today are discussed and compared based on published
literature from various EPDM producers.
II. INTRODUCTION TO EPDM
A. PRODUCT

EPDM is the ASTM designation given for a saturated polymer chain of the polymethylene
type, consisting of ethylene, propylene, and diene. The structure is shown in Figure 1. It belongs to
the family of polyolefins, occupying the rubbery composition range between the two most common
plastics in use todaypolyethylene and polypropylene. This is shown in Figure 2 as a plot of
crystallinity versus composition. Thus, the development of EPDM using the Nobel Prizewinning
catalyst technology by Ziegler and Natta (ZN technology) was a natural evolution, following the
production of plastics using that technology. The first copolymer (EPM) of ethylene and propylene,
designated VistalonTM 404 EPDM Rubber, was produced at the Baton Rouge ExxonMobil
Chemical facility in 1961, and it is still in production today.
The structure of EPDM shown in Figure 1 suggests that the polymer, without unsaturation on
the backbone, will be ozone resistant, and the thin chain architecture (large contour length per
unit volume in undiluted state), especially at higher ethylene content, will lead to low
entanglement molecular weight and therefore high plateau modulus, resulting in high ller
acceptance.1 These are some of the factors that make EPDM unique, compared with other
synthetic rubbers.
Like most other rubbers in use today, EPDM derives mechanical strength through network
formation enabled by the chemical cross-linking step also called vulcanization or curing. This, in
fact, is a key differentiator of rubber over plastics. Presence of a diolefin as a third monomer
improves the cross-linking response with peroxide cures and permits direct sulfur cure, depending
on the choice of the diolefin. Following the commercial production of EPM, it took several more
years to develop a terpolymer (EPDM) that would allow sulfur vulcanization that was in vogue for
other polymers at that time. A variety of dienes were explored and commercialized by the various
EPDM producers, although 5-ethylidene-2-norbornene (ENB) eventually became the industry
standard1,2 for sulfur vulcanization.

TREATISE ON EPDM

329

FIG. 1. Structure of EPDM.

The ability to use sulfur vulcanization allowed EPDM to enter an era of new opportunities.
EPDM was entering a world dominated by batch equipment originally developed for natural rubber
and styrene butadiene rubber (SBR). The advent of EPDM made it possible to use a low specific
gravity rubber with continuous vulcanization and the associated cost savings for the parts producer.
It was also an ideal replacement for chloroprene rubber in applications not requiring oil resistance.
B. PROCESS

EPDM is produced in a continuous process. Early in the development of EPDM, both the slurry
and solution processes appeared attractive with the first generation of ZN catalysts. Although the
slurry process is still in commercial use,4 the development of vanadium-based ZN catalysts
soluble in hydrocarbons, such as hexane, made the solution process a good choice and was adopted
by most manufacturers. While gas phase polymerization is very common for plastics, the technique

FIG. 2. EPDM crystallinity as a function of composition (for illustrative purposes only).

330

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)

is more challenging for rubbery polymers, as evidenced by the only successful commercialization
of this technology9 by Dow Chemical Co., which has since been decommercialized.
Recent advances in the development of metallocene catalysts have resulted in the conversion of
about 20% of EPDM capacity to technology based on these catalysts, and this figure is expected to
continue to grow.10 For this reason, throughout this review, comparative references will be made to
both types of EPDM platformsthe polymers produced by the conventional ZN technology will
be designated cEPDM and the polymers produced by the metallocene technology will be
designated mEPDM.
III. POLYMER PHYSICS
The major EPDM structural parameters that control the properties in various applications are
molecular weight (MW) as measured by Mooney viscosity or gel permeation chromatography
(GPC), molecular weight distribution (MWD) and long chain branching (LCB) as measured by
GPC or Mooney relaxation area (MLRA) or other rheological measures, crystallinity (ethylene
content), and level of unsaturation (diene content). Secondary parameters such as sequence
distribution, compositional tailoring, and so on, are also important. Catalyst and process
development to enable synthesis of various EPDM structures has been primarily driven by such
attribute requirements. Note that unlike polybutadiene type rubbers, EPDM typically does not have
intricate microstructural parameters related to the configurational details of the insertion of the
monomers into the polymer chain. Although even for EPDM, one could envision such
microstructural details since they relate to the tacticity of the propylene insertions, most EPDM
catalysts tend to be atactic in their propylene insertions.1
A. MOLECULAR WEIGHT

EPDM has been commercially produced over a wide range of MW. Number average MW (Mn)
can range from as low as 5000 to as high as 250 000. The low end of this range commercially
produced by some manufacturers can even be a liquid.11 At the other end of the spectrum, the
Mooney viscosity of the rubber can be so high that the rubber cannot be manufactured without
adding extension oil. In such instances, MW characterization of the reactor product requires
modified Mooney measurement techniques, to allow the use of commercially available Mooney
viscometers. In general, the peak MW in the MWD corresponds to the Mooney viscosity (ML),
which can be related to Mn using the empirical relationship
ML kMn n

Where, the exponent n is independent of MW for linear polymers and the constant k increases with
MWD. Thus, broader MWD polymers lead to lower Mn at a given ML. This relationship suggests
that at a given ML, broad MWD polymers that possess better processibility than narrow MWD
polymers will have a lower cure state because of the lower Mn. This indeed is the case.1
B. MOLECULAR WEIGHT DISTRIBUTION AND LONG CHAIN BRANCHING

MWD of EPDM as measured by GPC can display a variety of features. The inherent MWD is
primarily determined by the choice of polymerization catalyst.1 For single site (vanadium or
metallocene) catalysts in well-mixed continuous-flow stirred-tank reactors (CSTR), in the absence
of LCB and other features such as bimodality, the theoretical and experimentally observed MWD is
the most probable Mz:Mw:Mn 3:2:1 (ref 1). For catalysts that produce multiple sites, the Mw/Mn
ratio can be as high as 50. Other features that may be evident in a GPC trace include the presence of

TREATISE ON EPDM

331

LCB and evidence of bimodality that can be introduced for mEPDM through the use of multiple
reactors12 and multiple reactors in combination with the choice of catalyst for cEPDM.13
LCB in EPDM has been thoroughly investigated over many years.1,1416 For most mEPDM,
even in the presence of a diolefin such as ENB, the only available mechanism for LCB is the endgroup copolymerization, leading to trifunctional branch points.3,17 Branching of this type cannot
reach gel point.1416 Therefore, most commercially available EPDM produced with metallocene
catalysts are, by definition, free of gels as they exit the reactor, although gels can be produced even
in the mEPDM process through free-radical cross-linking, if they are present during polymer
synthesis, recovery, and storage. On the other hand, for cEPDM, the major branching mechanism
involves the exocyclic double bond on the diolefin. Coupling of chains across the enchained diene
that occurs through a cationic mechanism arising from acidic ZN catalyst systems leads to
tetrafunctional cross-links akin to what happens during vulcanization of rubber. Branching of this
type has a finite gel limit1416 and needs to be carefully controlled to avoid gel formation in the
reactor. This is even more critical for diolefins that possess two double bonds, both of which can
participate in polymerization. In this case, the approach to gel point and the postgel region can be
modeled18 using a numerical fractionation technique in which the method of moments is
applied to successively larger generation of molecules that are produced through a branchon-branch mechanism. The experimental MWD data for such EPDM are qualitatively
similar to the predictions based on this modeling approach.18,19 Upon reaching the gel point,
the gel fraction separates and stays within the reactor, while the sol fraction, which is still very
highly branched and multimodal in MWD, exits as the product of polymerization. Thus,
only a visual inspection of the reactor internals can ensure a gel-free polymerization. High
MW EPDM with substantial levels of vinylnorbornene (VNB) that are free of such macrogel
has been commercially produced by ExxonMobil for more than a decade.19 More recently,
other producers of EPDM have followed.20,21
While LCB can be inherent in the production of a variety of cEPDM and needs to be controlled,
its presence leads to desirable rheological characteristics, such as improved elasticity, that are not
obtainable with substantially linear mEPDM,22 unless the MWD can be altered, such as through the
introduction of bimodality.10,12 The lack of a cationic cross-link mechanism to couple across the
enchained ENB in mEPDM can be compensated by the introduction of branching through
polymerization using a minor amount of a diolefin such as VNB in addition to ENB.4,1921
C. CRYSTALLINITY

Examination of Figure 2 clearly shows that EPDM spans the composition range between
polyethylene (PE) and polypropylene (PP). PE is a highly crystalline plastic, with the
crystallinity originating from the methylene sequences. While atactic PP is not crystalline, most
commercially important PP possess crystallinity related to the tacticity of the propylene
insertions. Thus, it is to be expected that commercial EPDM that span the composition range of
5090 mol% ethylene (~4080 wt%) will exhibit crystallinity related to the methylene
sequences at higher ethylene content and will tend to an amorphous state as the ethylene content
decreases.1 In general, EPDM compositions below about 60 wt% are considered devoid of
crystallinity (amorphous) as evidenced by the lack of a melting peak, and compositions above
70 wt% are considered crystalline. It is customary to refer to the composition range 6070 wt%
as semicrystalline. Presence of crystallinity in EPDM has important consequences to
properties such as green strength of the shaped article prior to vulcanization, mixing and
processing of the rubber, cold ow, tack, and, for high MW rubber, ability to hold
extension oil without exudation.1 Strain induced crystallization can also lead to enhanced
mechanical properties of both the gum and compounded rubber.23

332

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)
D. ETHYLENE CONTENT

The range of ethylene content achievable for EPDM is a function of catalyst type. While
metallocene catalysts can typically polymerize both ethylene and propylene with ease and
therefore can span the entire range of compositions between PE and PP, cEPDM is constrained
to 5090 mol%. This is primarily a limitation of the ZN catalysts inability to insert propylene
next to propylene in the growing polymer chain. Besides the influence on crystallinity, the range
of ethylene content has other important consequences to the application of EPDM. With increase
in ethylene content, entanglement MW decreases, leading to a higher plateau modulus. The
higher plateau modulus, in turn, allows higher filler loading with retention of physical
properties, thus lowering compound formulation costs in applications such as industrial hose.
EPM polymers used as viscosity improvers in motor oils display improved thickening of the
oil as the ethylene content of the polymer increases at constant MW, until a limit is
reached where the increased crystallinity affects the ow ability of the motor oil. Novel
polymers have been commercially available for more than two decades for this
application, which tailors the intramolecular distribution of the ethylene content, leading
up to desired and unique properties.24,25
E. COMPOSITION DISTRIBUTION

Intermolecular compositional distribution (CD) can be achieved either by choice of catalyst or


through reaction engineering. For example, ZN vanadium catalysts combined with diethylaluminum chloride cocatalyst is known to produce multiple active catalyst sites.1 Each catalyst site
tends to display vastly different propensities toward incorporation of certain monomers into the
growing chain. ZN titanium catalysts that are normally insoluble in the polymerization solvent
also display this behavior. The resulting product is a superposition of polymers that can vary from
PE to atactic PP and the compositions in between. Presence of such broad CD (usually confounded
with broad MWD) can be revealed by fractional precipitation of the polymer dissolved in a good
solvent, with each fraction displaying vastly different composition and MW.26
For catalysts that tend to exhibit kinetics typical of single active species, for example,
vanadium catalysts combined with ethylaluminum sesquichloride or for most commercially used
metallocene catalysts, CD tends to be very uniform.26 If a broader, bimodal CD is desired in EPDM
produced with these catalysts, it can be accomplished through reaction engineering such as through
the use of reactors operating in series or parallel mode.27
F. SEQUENCE DISTRIBUTION

Even for a homogeneous polymer of a certain average composition, the arrangement of


monomer units along the chain can vary. This is referred to as the monomer sequence
distribution. When the polymer is prepared in a well-mixed CSTR, the sequence distribution is
predominantly determined by the choice of catalyst. For cEPDM, vanadium-based catalysts
tend to possess reactivity ratios between 0.1 and 0.5, leading to homogenous products. Titanium
catalysts with reactivity ratios in the 15 range tend to produce disperse (blocky) polymers.1
For mEPDM, the catalysts tend to be highly engineered and the desired reactivity ratio can be
part of the choice of catalyst design. More recently, it has been experimentally determined28
that even for single site catalysts with similar reactivity ratios, the distribution of methylene
sequences is narrower for mEPDM compared with cEPDM as shown in Figure 3. Such
differences at the molecular level can lead to profound differences in the behavior of the
polymer in certain applications.

TREATISE ON EPDM

333

FIG. 3. Distribution of methylene sequence lengths in ZN and metallocene polymers (each peak can be calibrated to
methylene sequence length using calibration standards).

G. SOLUTION PROPERTIES

Structural details of the polymer lead to different solution properties when dissolved in good
solvents. This effect is typically reflected in the inherent viscosity (IV) at a given MW (see figure 8
in ref 1). IV is related to MW and polymer composition through the MarkHouwink relationship.29
At a given MW, for linear EPDM, IV decreases with broadening MWD. In general, presence of
LCB lowers the IV, leading to a route for characterizing LCB through a branching index. Such
measurements are also critical in the design of new EPDM plants since the viscosity of the
polymer cement is directly related to the IV of the polymer that the plant is designed to
produce.
IV. CHARACTERIZATION TOOLS
A review of characterization tools for polymers, including EPDM, can be found in the
literature.30 Discussion here is focused only on EPDM.
A. MOLECULAR WEIGHT

While GPC is used to measure the various moments of the MWD, the industry standard for
routine measurement of MW is the Mooney viscosity. ASTM D1646 describes this measurement. It
is typically reported as ML (1 4) at 125 8C in Mooney units, where, L designates Large rotor
and the two numbers in parentheses refer to the preheat time and test time at the indicated
temperature. However, Mooney viscosity values greater than about 100 cannot generally be
reliably measured under these conditions. In this event, a higher temperature can be used
(i.e., 150 8C), with eventual longer shearing time (i.e., 1 8 @ 125 or 150 8C). However, this
would extend the useful range of measurement capability only by a small extent. Many
commercial EPDMs are produced at even higher MW in the reactor. An option to
characterize these polymers with the same Mooney machine is to carry out the Mooney
measurement on the reactor samples using a nonstandard (smaller and thinner) rotor
[Mooney small thin (MST)]. Details of this technique can be found in the patent literature.31
Typically, when such high MW EPDM is produced in the reactor, the polymer cannot be
recovered and nished downstream without lowering the viscosity, such as by using
extension oil. Empirical correlations can be used to relate the Mooney viscosities of the oilextended polymer and the neat polymer.

334

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)
ML1 4 @ 125 8C ConstantMST5 4 @ 200 8C100=100 PHRa

PHR stands for parts of oil per hundred parts of rubber, and the exponent a is dependent on the
degree of LCB in the polymer. In general, for a given MST and PHR, the value of ML decreases as
the level of LCB in the polymer increases. At the low MW end of the Mooney spectrum, it is difficult
to use the Mooney machine with EPDM less than about 10 ML. The range can be extended to a
small extent by lowering the measurement temperature to 100 8C. However, this temperature is too
low to prevent the influence of crystallinity on measured Mooney values for crystalline EPDM. For
such EPDM, it is preferable to use techniques such as melt index or melt flow rate that are used in the
plastics industry. For very low (liquid) MW EPDM, Brookfield viscosity measurements are used to
characterize MW.11 For blends of polymers, the Mooney viscosity can be obtained using the socalled log blending relationship,
log ML nA log MLA nB log MLB

where all logarithms are base 10 and ML is the Mooney viscosity of a blend of two polymers A and
B, each having individual Mooney viscosities MLA and MLB, respectively. The fraction of polymer
A in the blend is nA, while the fraction of the polymer B is nB.
B. MOLECULAR WEIGHT DISTRIBUTION AND LONG CHAIN BRANCHING

GPC using a filtered dilute solution of the polymer in trichlorobenzene is the standard
technique used to characterize MWD.1 For a typical linear EPDM, ratios of the various moments of
the MWD are sufficient to describe the breadth of the distribution. Presence of LCB and/or use of
multiple reactors to produce bimodal EPDM can significantly alter the MWD. In the absence of
LCB or at low levels of LCB (such as mEPDM with ENB), bimodality can be readily observed and
characterized through deconvolution of the GPC trace. GPC techniques32 use a series of three
detectorsdifferential refractive index (DRI), low angle laser light scattering (LALLS), and
viscosity (V)in tandem to gather a complete set of characterization data with a single injection of
the sample into the GPC column. When this technique is used for high MW EPDM, an appropriate
detector result should be used to characterize the various moments of the MWD. The DRI analysis
assumes that GPC columns separate macromolecules strictly by hydrodynamic size and therefore
give accurate values for Mn, regardless of the presence of LCB. It underpredicts Mw when LCB is
present because branched high MW chains elute where linear low MW chains are expected. The
LALLS detector, on the other hand, measures the true MW regardless of chain architecture and
gives true Mw even when LCB is present. However, the Mn value is more reliable from the DRI
signal as a result of the low sensitivity of the LALLS detector at lower MW (<100 000). The
viscosity detector is useful in obtaining the solution viscosity of the polymer as an integral part of
the GPC test in place of an independent laboratory test such as IV measured in decalin. Careful
control of GPC parameters such as polymer concentration, flow rate, particle size in the column
packing, and so on, is essential when EPDM containing a significant fraction of polymer at MW
higher than about 1 million (such as VNB EPDM19) is to be characterized using this technique, to
avoid erroneous results from breakdown of the high MW fraction during the test.
1. Branching Index. It is customary to express LCB in relation to a linear polymer
through a measure such as IV or MW or some combination of both, expressed as a ratio of
the properties measured on the branched polymer and calculated or measured for a linear
equivalent. One such measure used extensively in the patent literature is described here in
some detail. The relative degree of branching is determined using a branching index factor
(BI). This factor can be calculated using three laboratory measurements of polymer

TREATISE ON EPDM

335

properties in solutions.1 In general, these are (i ) Mw, GPC LALLS, weight average molecular
weight measured using a LALLS detector in combination with GPC; (ii ) weight average
molecular weight (Mw, DRI ) and viscosity average molecular weight (Mv,GPC DRI ) using a
DRI detector in combination with GPC; and (iii ) IV measured in decalin at 135 8C.
Historically, the first two of the tests were obtained from a single GPC injection, and the
third was obtained in an independent IV test. With the advent of the triple detection GPC,
all three of these results can be obtained from a single GPC measurement, as discussed
earlier.
An average branching index is defined as12,27:
BI

Mv;br 3 Mw;DRI
Mw;GPC LALLS 3 Mv GPC DRI

where, Mv, br (IV/k)1/a and a is the MarkHouwink constant ( 0.759 for EPDM in decalin
at 135 8C, when the IV used in this equation is the IV measured in decalin).1 From this
equation it follows that the branching index for a linear polymer is 1.0. For branched
polymers, the extent of branching is defined relative to the linear polymer. Since at a
constant number average molecular weight Mn, (MW )branch . (MW )linear, BI for branched
polymers is less than 1.0, and a smaller BI value denotes a higher level of branching. It is
important to note that when the IV value from a GPC viscometer is used in place of IV
measured in decalin, k and a values appropriate for the GPC solvent should be used in the
equation above.
BI values for mEPDM that typically fall in the range 0.851.0 are considered lightly branched
or linear for most purposes. For cEPDM with cationic branching through the diene, the BI values
range from 0.5 to 0.9. With branching through polymerization that is sufficiently controlled to avoid
gel region, BI can be as low as 0.1 and remain free of macrogel.19
2. Rheology. While GPC yields detailed information regarding MW and MWD, real time
tests for MWD characterization use rheological tools akin to Mooney viscometer for MW testing.
One such tool that is available as part of the Mooney test (and therefore does not require an
independent test for MWD and LCB determination) is the MLRA.33 This is the area under the
relaxation curve obtained as the polymer is allowed to relax at the end of the Mooney test, with the
rotor stopped. While this parameter is a very useful tool, it has certain limitations. For one, because it
is run at the end of the Mooney test, MLRA is a very strong function of Mooney. Therefore, it is
essential that measured MLRA values for different polymers be corrected to a constant Mooney
basis using linear interpolation before meaningful comparisons are made. In addition, MLRA
values reflect the total contribution from inherent MWD, LCB, and any modality present.
Individual contribution of each of these elements can only be deciphered by comparing MLRA of
families of EPDM with known features.34
Although MLRA as a rheological tool was incorporated into ASTM D1646 in 1993, each
EPDM manufacturer has chosen to use a different characterization tool for this purpose.4,22,35,36
Most of these rheological measures require an independent test performed on another instrument
besides the Mooney machine, such as the rubber processibility analyzer or a similar small amplitude
oscillatory shear instrument.37 The results obtained from these tests include viscoelastic properties
such as elastic modulus (G 0 ), loss modulus (G 00 ), and tan d ( G 00 /G 0 ). Typically, a frequency sweep
is performed (ASTM D6204), with the dependence of parameters such as phase angle d and
complex viscosity on frequency used as a measure of LCB. These rheological parameters tend to be
especially sensitive to the presence of LCB at low frequencies, although confounded with MW,
unless the data are presented in a MW independent form such as the Van Gurp plot or ColeCole
plot.38,39 Other techniques to remove the MW dependence of phase angle d measured at low

336

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)

frequencies have been proposed and used by some producers of EPDM.35 Careful sample
preparation technique, especially for very high MW rubbers, is essential to obtain reproducible
results. It is feasible to combine the tools discussed here to distinguish the nature of LCB such as
between cEPDM and mEPDM.40 Recent advances in rheological characterization include the
ability to employ large amplitude oscillatory strain (LAOS) and Fourier transform (FT) rheometry
to study nonlinear viscoelasticity, especially of compounded rubber and complex gum rubber such
as natural rubber.41
C. COMPOSITION

Ethylene content is determined by FT infrared (FTIR), ASTM D3900. The reported ethylene
value from this test for EPDM ignores the presence of diene and therefore should always be
corrected as such to infer the actual content of ethylene and propylene in the presence of diene.
Diene [ENB and dicyclopentadiene (DCPD)] content incorporated in the polymer is determined by
FTIR, ASTM D6047. Other dienes can be measured via proton nuclear magnetic resonance
(NMR). In some instances, such as for VNB, it is possible to develop a procedure similar to ASTM
D6047 using the FTIR technique. These methods only measure available unsaturation. In cases
such as VNB, where a significant level of unsaturation is consumed in branching reactions, the total
diene content can be measured by refractive index, using a calibration for a diene such as ENB with
the pendent unsaturation intact. Test methods based on FTIR can also be used for quantification of
antioxidants such as Irganox-1076, typically added to EPDM for warehouse stability, using the
ester carbonyl absorbance at a wave length of 1741 cm1. For oil-extended EPDM, the extension oil
should first be removed by solvent extraction prior to the use of these analytic techniques. The oil
content itself can be determined using ASTM D297, which specifies a variety of solvent extractions
of rubber for various purposes. Extractions involving a refluxing apparatus such as Soxhlet or
Kumagawa, with boiling methyl ethyl ketone, are useful in oil content determination without errors
related to overextraction of low MW rubber along with the oil. Residual impurities in EPDM that
originate from the catalyst residues can be quantified using inductively coupled plasma emission
spectroscopy (ICPES). While this technique is useful to quantify a very large number of impurity
species, it is not conducive to quality assurance testing in a manufacturing plant. For this purpose, it
is possible to use X-ray fluorescence spectrometry with suitable calibration standards for specific
elements, based on ICPES.
D. CRYSTALLINITY

As noted earlier, crystallinity in EPDM arises from the backbone methylene sequences. The
standard technique used to measure crystallinity is differential scanning calorimetry (DSC). ASTM
D3418 describes the typical procedure used to measure the crystallinity of EPDM. However,
variations in details of this procedure such as sample annealing time, heating and cooling rates, and
so on, can have a profound impact on the reported crystallinity parameters and should be taken into
account when comparing results across various laboratories. In addition, EPDM with broad or
bimodal CD can exhibit multiple melting peaks. Therefore, presentation of a single crystallinity
value is seldom useful.
E. SEQUENCE DISTRIBUTION
13

C NMR can reveal sequences up to five carbons.1 It allows the establishment of regio and
stereo chemistry (type of insertions) of propylene in the backbone. When combined with kinetic

TREATISE ON EPDM

337

FIG. 4. Dienes commonly used to produce EPDM.

models and experimental data, it allows the estimation of the reactivity ratio for a given catalyst
system used to synthesize the EPDM.
V. COMPOUNDING AND VULCANIZATION
EPDM derives the majority of its useful properties through compounding with other
ingredients, such as fillers and plasticizers, followed by vulcanization that promotes cross-linking
of the polymer chains. Structural features of the polymer can impact all aspects of the compounded
rubber. For example, typically, polymer is a relatively expensive component of the final compound.
Therefore, polymer properties that lead to higher filler loading without the loss of compound
attributes are desirable. As mentioned earlier, higher ethylene content leads to higher plateau
modulus and higher filler acceptance. Rubber compounding itself is a time and energy intensive
process.6 Therefore structural features that reduce filler incorporation time with uniform dispersion
of the ingredients at the lowest mixing energy are desirable. Typically, the compounded rubber is
shaped into the desired article prior to vulcanization. In this regard, polymers that result in low
compound Mooney viscosity, good processing characteristics (extrusion, molding, calendering,
etc.), and good green strength for shape retention prior to vulcanization are attractive.
The majority of EPDM, as with other synthetic rubbers, is vulcanized with sulfur-accelerator
systems. The primary requirement is the presence of unsaturation in the polymer. For this reason,
EPM copolymers cannot be sulfur cured. In applications that require vulcanization of EPM
copolymers or enhanced EPDM properties such as good heat resistance, low compression set at
high temperatures (.125 8C), and high cure rate, peroxide vulcanization is preferred.3 Other, less
commonly used cure systems for thermoset applications (but, widely used for thermoplastic
vulcanizates) include phenolic resin cure and hydrosilation cross-linking.6,4244 A more detailed
discussion of EPDM compounding and vulcanization can be found elsewhere.6
A. CHOICE OF DIENE

The structures of commonly used dienes in EPDM are shown in Figure 4.


Nonconjugated dienes with pendent unsaturation are preferred over conjugated dienes to
preserve the benefits that EPDM derives from having a saturated backbone.1 Thus, a diolefin
where one of the two double bonds is polymerized into the polymer backbone with ease will
be beneficial. Many such diolefins were investigated early on in the history of EPDM, and
ENB was chosen mainly because of the decision by Union Carbide Corporation to

338

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)

commercialize this diene.2 The typical range of diene composition in EPDM is 012 wt%
(~04 mol%). It is an expensive raw material, when compared with the other two
monomers, ethylene and propylene, by as much as a factor of 3 to 4. Considerations that
influence the choice and level of diene for a given application include (i ) choice of
vulcanization chemistry; (ii ) type of vulcanization (batch vs continuous); (iii ) desired
balance of properties; and (iv) desired compound rheology. It has been observed10 that
mEPDM reaches the same cure state as cEPDM at a substantially lower diene content,
resulting in significant cost savings. One could speculate that if other variables such as Mn
are comparable for the two polymers, this observation may be attributable to a more uniform
distribution of diene along the growing chain in the case of mEPDM, leading to higher use
in the cross-linked network. Considerations that impact the manufacturing process for a
given choice and level of diene include (i ) ease of incorporation in the polymer backbone as
measured by the conversion of the diene; (ii ) degree of poisoning to the catalyst system as
measured by the catalyst productivity and propylene conversion; (iii ) ability to fractionate
and recycle the unconverted diene; and (iv) environmental considerations. While ENB is
widely used by most producers of EPDM, some other dienes are also in limited commercial
use. For example, ExxonMobil produces commercial grades containing VNB which,
although inactive in sulfur vulcanization, is almost a factor of three better than ENB in
peroxide vulcanization.45 DCPD is also available from several EPDM producers as an
alternate diene choice to ENB. It is important to control the polymerization operating
window for these dienes to avoid the formation of macrogel.4,19 EPDM containing more
than one diene is available from some producers.4,46,47
VI. APPLICATIONS
EPDM is used in a wide variety of applications.28 In this section, major application areas will be
addressed from the polymer structure-property point of view.
A. THERMOPLASTIC VULCANIZATES

Thermoplastic vulcanizates (TPVs), also referred to as thermoplastic elastomers or


dynamically vulcanized alloys (DVAs), exhibit rubber-like behavior and yet can be
processed in thermoplastic processing equipment.48 They overcome a major disadvantage of
thermoset (vulcanized) EPDM in that they can be recycled, just like plastics. While DVAs
can be produced from many combinations of vulcanized rubber and plastic, those produced
from EPDM and PP are aimed directly as a replacement for thermoset EPDM. They are
especially useful in applications requiring high temperature performance as a result of the
high melting point of isotactic PP, commonly used as the plastic phase. They can be
produced in a variety of hardnesses (soft rubber-like to hard plastic-like) and can be made to
serve a variety of markets, from consumer articles to highly engineered automotive parts.
EPDM/PP TPVs are typically produced in a reactive extrusion process using a
continuous twin-screw extruder.9,4951 The individual feeds of rubber and plastic go through
a mixing section, followed by the introduction of curatives, fillers, and plasticizers. The
simultaneous vulcanization and mixing of the EPDM produces a final product in which 12
l sized particles of vulcanized EPDM are dispersed in a PP matrix.43,50 The entire process is
rapid, with a residence time in the extruder on the order of minutes. The requirement for the
degree of cure of the EPDM phase is driven by the application, with highly engineered
applications demanding a high degree of cure (.96%, i.e., <4% sol fraction). Such a high
degree of cure can be accomplished using medium diene, ultrahigh MW EPDM, which are

TREATISE ON EPDM

339

FIG. 5. MV wire and cable construction.

typically oil extended up to 100 parts of oil per hundred parts of rubber. Since TPVs contain
substantial amounts of oil in addition to the extension oil added to the rubber, amorphous to
semicrystalline EPDM is preferred to avoid exudation. Novel EPDM structures that produce
better TPVs at lower MW and oil-extension levels have been proposed.52
TPVs that are nonhygroscopic and colorable using an alternate cure system such as
hydrosilation and peroxide have been commercially produced for some time.44,53,54 EPDMs
containing VNB as the diene are described as advantageous over EPDMs containing ENB
for TPVs using these cure systems.
B. WIRE AND CABLE INSULATION

This EPDM application is typically segmented into high (.35 kV), medium (535 kV),
and low voltage (<5 kV).55 Low voltage cables such as appliance cables are typically highly
filled and require a crystalline EPDM of medium diene content and high MW (~60 ML (1
4) 125 8C). High voltage cables tend to be very specialized. The medium voltage (MV)
segment has been historically served by cross-linked polyethylene with increased penetration
by EPDM insulations owing to the improved flexibility, dimensional stability at higher
service temperature, water tree resistance, and service life.56 However, the rubber
requirements for this segment are stringent, requiring a long qualification cycle.57 The
insulation layer (Figure 5) sandwiched between the insulation and conductor shield is
typically coextruded in a multihead extruder.58 Thus, the polymer choice is dictated by the
need for fast and smooth extrusion and fast cure rate to accommodate continuous
vulcanization (typically using peroxide cure). While each cable manufacturer uses a
proprietary formulation around a chosen EPDM, variations of an industry standard
formulation, SuperOhm 3728 is used by most cable producers. The requirements outlined
above are optimally met by a VNB EPDM, which has therefore been the leading polymer in
this application.45,59 The diene content is low (even copolymers can be used if deficiency in
cure rate and state can be tolerated) to provide good heat ageing, and the polymer
crystallinity is optimized to provide required tensile strength without the treeing
phenomenon associated with too high a polymer crystallinity. More recently,
technological innovations in mEPDM and plastomers as well as the cost pressures
on the cable producers have led to substitution of at least some of the cEPDM in their
formulations with these polymers.60

340

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)
C. OIL MODIFICATION

EPM copolymers have been used for a long time as viscosity index improvers in motor oils,
transforming them to multigrade oils.61,62 Two factors that govern the polymer selection are the socalled shear stability index (SSI) as measured according to ASTM D-7109 and thickening
efficiency (TE). SSI is an end-use requirement (such as oil change frequency in a car) and specified
by the formulator of the motor oil. Lower numbers indicate a more shear stable polymer. TE is the
ability of the polymer to increase the base oil kinematic viscosity (ASTM D-445) by a factor of two.
Higher numbers indicate lower polymer addition. In general, TE increases with SSI. Since SSI is
specified, the polymer with the highest TE at that SSI will be the most desirable. Typically, for a
given MWD, target Mooney viscosity of the polymer is chosen based on the desired SSI. Besides
MW, the TE is maximized through the choice of polymer composition. Stiff chains with fewer and
smaller side groups tend to have a higher intrinsic viscosity, and therefore higher TE, at a given
MW.61 Narrow MWD polymers tend to be more shear stable, leading to lower SSI at a given TE.
Therefore, at a given MW, higher ethylene content and narrower MWD directionally increase the
TE at a given SSI. However, the choice of ethylene content is typically driven by the choice of base
stock used to formulate the oil, dictated by polymerwax interaction. This leaves MWD as the
major differentiator between various EPMs in commercial use.61 Since the MWD is invariant at a
value of 2.0 for single site catalysts in a CSTR, narrower MWD can be obtained only through
pseudoliving polymerization in a tubular reactor.63,64 An added advantage of the use of a
tubular reactor is the ability to tailor the intrachain compositional distribution, leading to
favorable structures when the high ethylene segments interact with each other in the motor
oil.61 Such polymers have been commercially available for more than two decades.
D. AUTOMOTIVE

The majority of EPDM is used in automotive applications.65,66 These include dense and
sponge weather seals, coolant hoses, power transmission belts, brake parts, and so on. Each of these
parts demands a different EPDM structure, depending on the attribute requirements, compound
formulation, processing steps, and type of processing equipment.
1. Body Sealing. Automotive body sealing is one of the largest applications for EPDM.65,66
It is also one of the most challenging in that the demands are exacting and cost pressures on the
suppliers to the automobile industry are high. The process steps typically involve preparing the
compound by mixing the polymer with fillers in an internal mixer such as a Banbury, sheeting the
compound on a mill, feeding strips of the sheet into an extruder with a die that matches the shape of
the desired profile, and curing the extruded profile in a continuous vulcanization oven (hot air
tunnel). Suitable polymer selection34,67 will allow fast one-pass mixing of the compound, high
extrusion throughput matched by fast cure rate in the continuous vulcanization step. The polymer
and the compound have to be free of imperfections such as microscopic gels,68 especially if the
profile will be a show surface in the automobile. Typically, a gel filter is employed in the production
of these EPDMs to produce a gel-free polymer. For dense weather seals, EPDM with high diene
content for fast cure and high cure state and low ethylene content for good cold temperature
compression set (sealing) and load deflection is preferred. As noted earlier, mEPDM appears to
satisfy these requirements at lower diene content than cEPDM.10,65 The processing requirements
can be met with cEPDM by choosing a catalyst with a very short lifetime so the trailing reactor in a
series reactor configuration can be operated in a manner that produces a low amount of very high
MW polymer, resulting in a highly processible bimodal EPDM.13,34 Sponge seals are even more
demanding, since the cell structure in the sponge has to be carefully controlled for optimum sealing
and collapse resistance.69 Corner moldings produced as part of the seal assembly are typically high

TREATISE ON EPDM

341

hardness compounds and require a very high ethylene, medium diene EPDM of low MW. Recent
developments in the body sealing segment include the migration of thermoset sealing systems to
TPVs based on EPDM and hybrid sealing systems, driven by low system cost, design flexibility,
recyclability, and improved aesthetics.70,71
2. Coolant Hose. Automotive coolant hoses demand a very long service life, lasting over
250 000 km or 10 years, electrochemically resistant compounds, and good low-temperature
sealing.66,72 The production of these hoses is a multistep process that requires significant handling
of the green (unvulcanized) part.66 Thus, green strength is an important polymer requirement.
Conflicting requirements such as good low-temperature sealing and high green strength can be met
through control of compositional distribution or through blending with a very high MW EPDM.
While the North American (NA) producers still predominantly produce sulfur cured hoses,
peroxide cure is more common in Europe.66 Use of VNB as the diene for peroxide cured hoses can
be beneficial and cost effective.73,74
3. Molded Parts. A variety of automotive parts, such as brake cups and diaphragms, are
typically produced using any of the variety of molding techniques.75,76 These formulations also
tend to be very rubber rich and low in plasticizers and fillers.6 This necessitates the use of a low MW
EPDM with medium diene content and completely amorphous structure for good low-temperature
sealing. If cold flow presents an issue, polymers with bimodal CD or cEPDM produced with a
multisited catalyst system that results in a broad CD can be employed. It is essential to use a very
linear EPDM for this application since good tear strength is a key requirement. This requirement
makes this application a good fit for mEPDM.10 Since a majority of these parts are peroxide cured,
use of VNB as the diene could be an advantage.
E. ROOF SHEETING

EPDM roofing membranes are fairly common in commercial buildings in NA. Water
management applications such as geomembranes and pond liners also share some of the same
requirements as roofing membranes. They are produced in very large batches (up to 600 kg)77 and
calendered into sheets that are several meters wide. Several of these sheets are spliced together to
produce even larger sheets (master role). Critical requirements therefore include a balance
between the splice tack and the green strength required to prevent sagging. Polymer design is
crucial in achieving this balance.78 Beyond the MWD and CD, these polymers should be free
of substantial LCB, making them a good t for mEPDM. The diene content is maintained
low to promote good heat ageing and prolonged service life. Another application that uses
calendering as a process step and therefore shares the polymer requirements with roof
sheeting is power transmission belts (PTBs), a relatively new segment for EPDM. Unlike
roong membranes, PTB is typically peroxide cured, and therefore either a copolymer or a
VNB EPDM is preferred.
F. MISCELLANEOUS APPLICATIONS

EPDM is also used in the production of thermoplastic olefin compounds for automotive
bumpers and interior panels. However, it has been mostly replaced by plastomers in recent years.
Adding polar functional groups to the backbone in a postpolymerization step has allowed the use of
a nonpolar polymer like EPDM in blends with polar polymers such as nylon.79 Applications include
impact strength and toughness improvement of engineering plastics, compatibilization, and
promotion of adhesion to polar substrates. EPDM is also used as a minor component in tires and
tubes along with butyl rubber.

342

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)

FIG. 6. Structure of ZieglerNatta and metallocene catalysts.

VII. MANUFACTURING PROCESS


The abbreviations cEPDM and mEPDM used in this review denote EPDM produced with
ZieglerNatta and metallocene catalysts, respectively. The characteristics of these two classes of
catalysts, shown pictorially in Figure 6, are compared in Table I. Both catalyst types belong to the
broad class of chain growth polymerization with coordinative insertion of monomers at a metal
alkyl bond. The commercial use of metallocene catalysts was pioneered in the early 1990s by
ExxonMobil Chemical.80,81 Since then, EPDM based on metallocene catalysts has been available
from several producers, including ExxonMobil Chemical, Dow Chemical Company, and Mitsui
Chemicals, Inc. Each EPDM producer uses a suite of proprietary catalysts.
In general, the ZieglerNatta catalyst is a complex of two components: a catalyst that is a
transition metal halide, such as TiCl4, VCl4, VOCl3, and so on, of which VOCl3 is the most widely
used; and a cocatalyst that is a metal alkyl component, such as (C2H5 )2AlCl diethylaluminum
chloride, or ethylaluminum sesquichloride, (C2H5 )3Al2Cl3. The aluminum alkyls need at least one
halogen to be an effective cocatalyst. The catalyst/cocatalyst combinations produce complexes
soluble in nonpolar hydrocarbon medium (Figure 6a). As discussed earlier, the diethylaluminum
chloride produces multiple active species, leading to a broad MWD and CD, whereas the
ethylaluminum sesquichloride tends to produce single species, leading to a narrow MWD and CD.
A metallocene catalyst, shown pictorially in Figure 6b, consists of a transition-metal atom
sandwiched between ring structures through a bridge system that prevents ligand rotation, locking it
into a single geometry. The nature of the metal center, ligands, and polymerization conditions
dictate the microstructures and physical properties of the polymer prepared. Akin to a cocatalyst
used with ZieglerNatta catalysts, an activator is required to generate the active site with
metallocene catalysts. This is typically an alkyl alumoxane or a noncoordinating anion, depending
on the structure of the catalyst. Typically, an alkyl scavenger is required to sustain metallocene
polymerization because of the high sensitivity of these catalysts to poisons such as water.
Given the vastly different nature of the two catalyst types, it is to be expected that the
manufacturing process can be designed to benefit from their unique attributes. The details of the
manufacturing processes used by each of the global EPDM producers remain proprietary. For
illustrative purposes, the steps involved in the manufacture of both cEPDM and mEPDM using a
solution process are shown in Figure 7 on a comparative basis. In this figure, the steps that are
unique to the mEPDM process are shown in italic.
In both process platforms, it is important to use adsorbent beds to remove moisture and other
impurities from the feed stream. Typically a molecular sieve and/or alumina are used. Catalyst
components are typically delivered to the reactor as a homogeneous solution. This may require a
preparation step depending on the form in which the catalyst components are received. The

TREATISE ON EPDM

343

TABLE I
COMPARISON OF ZIEGLERNATTA AND METALLOCENE CATALYST

ZieglerNatta catalyst

Metallocene catalyst

Ill-defined structures with multiple geometry


Broad molecular weight distribution
is possible through catalyst selection
Poor thermal stability
Low reactor temperatures
Low lifetime
Poor comonomer incorporation
Limited to propylene as comonomer
Low activity
Less expensive

Well defined single geometry


Narrow molecular weight distribution
High temperature stability
High reactor temperatures
Long lifetime
Good comonomer incorporation
Broader comonomer range possible
High activity
More expensive

components can also be either premixed or delivered separately to the reactor. In most instances,
design details that are proprietary to each producer, such as the placement of the nozzles delivering
the catalyst components inside the reactor, the type of agitators chosen to mix the reactor contents,
and so on, can have an impact on productivity of the catalyst as well as product attributes.
EPDM polymerization is an exothermic process. Therefore, the heat of polymerization has to
be removed to allow the operation of the reactor at the desired temperature. This can be
accomplished either through the use of adiabatic liquid-full reactors with chilled feeds or through
the use of jacketed reactors in combination with a boiling pool and an overhead condenser.1 The
choice of reactor temperature is typically dictated by the characteristics of the catalyst and the choice
of reactor pressure by the type of operation employed downstream of the reactor. The influence of
catalyst choice on reaction temperature is typically dictated by reaction kinetics for cEPDM and by
both reaction kinetics and tendency for spontaneous chain transfer for mEPDM. ZN catalysts tend
to shut down at temperatures above ~65 8C as termination rate dominates over propagation rate.1
Metallocene catalysts, on the other hand, exhibit a wide variety of behaviors. In general, unlike ZN
catalysts, the propagation rate increases rapidly with increase in temperature, compared with
termination rate, favoring higher reaction temperatures. However, most metallocene catalysts also

FIG. 7. EPDM manufacturing process scheme.

344

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)

exhibit a tendency to undergo chain transfer through beta hydride elimination.82,83 This process
accelerates with an increase in reaction temperature, thus lowering MW. The best metallocene
catalysts for the production of mEPDM allow the production of high MW EPDM even in the presence
of such a chain transfer process, allowing reaction temperatures well in excess of 100150 8C. Both
chain transfer to monomer and beta hydride elimination result in unsaturated chain ends that can
reincorporate in a growing chain, leading the trifunctional branch points.17 For most metallocene
catalysts, this is the only pathway available to introduce LCB in the polymer with ENB as the diene. A
second reactor can be used in series or parallel operation on either platform to increase the
productivity of the catalyst and/or tailor the product attributes as discussed earlier.12,13
Every EPDM producer develops recipes for process control that use the broad principles
outlined here, leading to maximum throughput at the lowest cost in a robust operating window.
Typically, MW (Mooney viscosity) is regulated through the use of a chain transfer agent such as
hydrogen. For cEPDM that exhibits LCB through the ENB, degree of branching is typically
regulated by acidity control in the reactor through the use of a Lewis base. Other process variables
such as reactor temperature, monomer concentration, catalyst consumption rate, residence time,
nature of reactor agitation, and so on, can all significantly influence the control of polymer properties.
Similar considerations apply to the control of mEPDM polymerization except for branching control,
since LCB through ENB is typically absent. Instead, branching control for mEPDM is accomplished
through control of unsaturated chain ends and their reincorporation in a growing chain.
The polymer cement leaving the reactor is typically quenched to stop the
polymerization. The beta hydride chain transfer process for metallocene catalysts leads to
another important consequence. When combined with the relatively long lifetime of these
catalysts, a single active catalyst site can produce hundreds of polymer chains prior to
termination, leading to catalyst productivity (measured as catalyst efciency in wt/wt basis)
that is two orders of magnitude higher than ZN catalysts. The high catalyst productivity
results in very low levels of catalyst residue in the polymer and obviates the need for a deashing step to remove these residues. The de-ashing step typically consists of contacting the
quenched polymer cement with water or other aqueous media that removes the catalyst
residue by mass transfer from the hydrocarbon cement into the aqueous phase. Extension oil
and antioxidants are typically added to the de-ashed cement prior to polymer recovery.
Polymer recovery for cEPDM is typically accomplished in a manner similar to other synthetic
rubbers, using the wet rubber finishing process.1 This involves flashing the hydrocarbon with steam
and hot water, leaving behind a polymer crumb. The wet crumb that could still have a
signicant level of hydrocarbon along with water is dried by extrusion drying, and the dry
crumb is conveyed to balers, producing bales of rubber. Such a nishing process can be
associated with signicant emission of hydrocarbon into the ambient atmosphere. Some of
these emissions can be captured and sent to a regenerative thermal oxidizer as an emission
control device. The hydrocarbons ashed in the ashing step are recovered through extensive
distillation and drying before being recycled to the reactor. Hydrocarbon recovery and
recycling is the most energy intensive step in the wet rubber nishing process.
The high reaction temperature of the mEPDM process results in several benefits. The lower
viscosity of the polymer cement at higher temperature allows higher cement concentration leaving
the reactor, when compared with cEPDM of the same MW. In addition, the high enthalpic content of
the high temperature cement can be captured through proper choice of postreaction polymer
recovery steps. If the reactor is operated at sufficiently high pressure, multistage flashing of the
hydrocarbon through pressure reduction leads to concentration of the polymer cement. The
concentrated cement is subjected to a devolatilization step for complete removal of hydrocarbons
prior to pelletization. An alternative concentration step is described in the patent literature, which
uses the thermodynamic principle that polymer solutions can undergo liquid/liquid phase

TREATISE ON EPDM

345

TABLE II
ADVANTAGES OF MEPDM PROCESS OVER CEPDM

Advantage

Feature

Intrinsic to catalyst
Lower energy consumption
No deashing/low water emissions
Wider product composition window
Not intrinsic to catalyst
Simplified recycle fractionation
Low contamination
Low MW capability
Very low air (VOC) emissions
Product form

High reactor temperature


High catalyst activity
High C3, broader series reactor options,
higher alpha olefins
L-L separation
Closed finishing process, melt filtration,
no slurry aids
Devol finishing
Devol finishing
Both pellets and bales

separation at the lower critical solution temperature.84 The liquid phase that contains the majority of
the polymer is subject to further concentration and devolatilization. The liquid phase that is devoid
of polymer can be directly recycled to the reactor, thus saving significant energy and costs
associated with the monomer and solvent recovery step. A disadvantage of this closed nishing
process, in which impurities in additives such as extension oil have a direct path to the reactor,
is that the production of high MW oil-extended EPDM presents a challenge. It is also difcult
to reprocess nished product, a common practice in cEPDM process. A comparison of the
two process platforms is shown in Tables II and III.
Recently, at least one commercial producer of cEPDM has announced the use of a process that
appears to combine some of the benefits of the mEPDM type polymer recovery steps with the use of
ZN catalysts.85
TABLE III
ADVANTAGES OF CEPDM PROCESS OVER MEPDM

Advantage
Intrinsic to catalyst
Range of MWD possible
Tailored MWD through series reactors
Easy processing singe reactor products
VNB EPDM
Ultranarrow MWD tubular reactor products
Not intrinsic to catalyst
Produce ultrahigh MW EPDM
Better economics through reprocessing
Easier to purge poisons

Feature

Both single and multisited catalysts available


Low catalyst life time allows MWD tailoring
Long chain branching through ENB
More facile branching control
Lower cost catalyst with low levels
of chain transfer
Not a closed systemcan add extender oil
Not a closed systemcan add back
reprocessible polymer
Extensive fractionation of recycle streams

346

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)
VIII. EPILOGUE

The review of EPDM presented in this article is intended to supplement prior reviews18 on this
subject as well as provide an alternate perspective, capturing the recent product/process
developments. For a polymer that is more than five decades in commercial production, there is a
high level of innovation that continues even today. The innovation cycle typically starts with the
development of novel catalysts, followed by process development, and eventually new products
that meet the market needs by expanding volume (through both substitution and organic growth),
lowering cost, and delivering end-use attributes, leading to novel process and product platforms.
This cycle of innovation that started with the invention of the ZieglerNatta catalysts is being
followed by a similar cycle of innovation that started with the invention of the metallocene catalyst.
This new cycle is currently at the catalyst and process innovation stage8688 and, needless to say,
product innovations that even extend to other polyolefins beyond EPDM, leading to new markets
and applications, will follow in the years to come.
IX. REFERENCES
1

G. Ver Strate, Ethylene-Propylene Elastomers, in Encyclopedia of Polymer Science and Engineering, Vol. 6, Wiley,
New York, 1986, p. 522.

E. N. Kresge, RUBBER CHEM. TECHNOL. 83, 227 (2010).

F. P. Baldwin and G. Ver Strate, RUBBER CHEM. TECHNOL. 45, 709 (1972).

S. Bhattacharjee, H. Bender, and D. Padliya, RUBBER CHEM. TECHNOL. 76, 1057 (2003).

J. W. M. Noordermeer, Ethylene-Propylene Polymers, in Kirk-Othmer Encyclopedia of Chemical Technology, Vol. 10,


Wiley, New York, 2005, p. 704.

S. Jacob, E. Jourdain, and B. Rodgers, Ethylene-Propylene Rubbers, in The Vanderbilt Rubber Handbook, R.T.
Vanderbilt Co., Norwalk, CT, 2010, p. 138.

S. Cesca, J. Polym. Sci. 10, 1 (1975).

S. L. Bell, Production of Ethylene-Based Elastomers by a Solution Process with ExxpolTM Metallocene Catalyst, in
PEP Review 2005-08, IHS Chemical, Englewood, CO, 2005.

G. Williams, T. Clayfield, and R. Johnston, Nordelt MG The Game Changer...for TPV, paper presented at the
Sixth International Conference on Thermoplastic Elastomers 2003, Brussels, Belgium, September 1617, 2003.

10

E. Jourdain, S. Jacob, G. Wouters, Y. Devorest, M. Joshi, M. Bednarik, and M. F. Welker, Amorphous Metallocene
EPDM for Superior Low Temperature Properties, Paper 25, presented at the 176th Fall Meeting of the Rubber
Division, ACS, Pittsburgh, October 1313, 2009.

11

A. U. Paeglis, F. C. Cesare, and D. N. Matthews, Liquid EPDM Elastomers, Paper 80 presented at the 132nd Fall
Meeting of the Rubber Division, ACS, Cleveland, October 68, 1987.

12

P. S. Ravishankar and N. R. Dharmarajan, U.S. Patent 6,271,311 (to ExxonMobil Chemical Patents Inc.), August 7, 2001.

13

S. Datta and E. N. Kresge, U.S. Patent 4,722,971 (to Exxon Chemical Patents, Inc.), February 2, 1988.

14

E. N. Kresge, C. Cozewith, and G. Ver Strate, Long Chain Branching and Gel in EPDM, paper presented at the 125th

Meeting of the Rubber Division, ACS, Indianapolis, May 8-11, 1984.


15

C. Cozewith, G. Ver Strate, and W. Graessley, Chem. Eng. Sci. 34, 245 (1979).

16

G. Ver Strate, C. Cozewith, and W. Graessley, J. Appl. Polym. Sci. 25, 59 (1980).

17

W.-J. Wang, S. Zhu, and S.-J. Park, Macromolecules 33, 5770 (2000).

18

R. Li, A. B. Corripio, K. M. Dooley, M. A. Henson, and M. J. Kurtz, Chem. Eng. Sci. 59, 2297 (2004).

19

P. S. Ravishankar, VNB EPDM Free from Macro-Gel Through Novel Control of Long Chain Branching, Melvin

Mooney Distinguished Technology Award Presentation, Rubber Division, ACS, Akron, April 19, 2011.

TREATISE ON EPDM

347

20

M. Van Duin, M. Dees, and H. Dikland, Advantages of EPDM Rubber Products with 2-Vinyl-5-Norbornene as Third
MonomerPart I: Improved Peroxide Curing Efficiency, Paper 49, presented at the 172nd Meeting of the Rubber
Division, ACS, Cleveland, October 1618, 2007.

21

M. Zuideveld, E. Ijpeij, E. Arts, and G. van Doremaele, High Performance Kelten ACEe Advanced Catalysts for
Innovative EPDMs, paper presented at the meeting of Advances in Polyolens, Santa Rosa, CA, September 2528,
2009.

22

J. G. Pillow, Kautsch. Gummi Kunstst. 51, 855 (1998).

23

B. Scholtens, E. Riande, and J. Mark, J. Polym. Sci. Polym. Phys. 22, 1223 (1984).

24

J. E. Johnston, R. Bloch, G. Ver Strate, and W. R. Song, U.S. Patent 4,507,515 (to Exxon Research and Engineering Co.),
March 26, 1985.

25

G. Ver Strate, R. Bloch, M. J. Struglinski, J. E. Johnston, and R. K. West, U.S. Patent 4,804,794 (to Exxon Chemical
Patents), February 14, 1989.

26

R. Liotta and P. S. Ravishankar, Design of an EPDM Polymer to Achieve Exceptional Processibility and Product
Attributes in Molding Applications,paper presented at the Meeting of the Rubber Division, ACS, Orlando, FL
September 2124, 1999.

27

P. S. Ravishankar and N. R. Dharmarajan, U.S. Patent 7,199,189 (to ExxonMobil Chemical Patents Inc.), April 3, 2007.

28

P. S. Ravishankar, EPDMPast, Present, and Future, Paper 117, presented at the 176th Meeting of the Rubber

Division, ACS, Pittsburgh, October 1315, 2009.


29

G. Scholte, N. L. J. Meijerink, H. M. Schoffeleers, and A. M. G. Brands, J. Appl. Polym. Sci. 29, 3763 (1984).

30

N. P. Cheremisinoff, Polymer Characterization Laboratory Techniques and Analysis, Noyes Publications, Westwood, NJ,

1996.
31

G. J. Wouters, P. S. Ravishankar, K. L. Chasey, and E. P. Jourdain, U.S. Patent 6,686,419 (to ExxonMobil Chemical
Patents Inc.), February 3, 2004.

32

W.J. Wang, S. Kharchenko, K. Migler, and S. Zhu, Polymer 45, 6495 (2004).

33

C. B. Friedersdorf and I. Duvdevani, Rubber World 211, 30 (1995).

34

O. Georjon and E.P. Jourdain, Industrial Scale-up of High Productivity Extrusion with Bimodal EPDM,paper
presented at the International Rubber Conference 2001, Birmingham, England, June 1214, 2001.

35

H. C. Booij, Kautsch. Gummi Kunstst. 44, 128 (1991).

36

P. W. L. J. Meijers, L. R. Maag, H. J. H. Beelen, and P. M. van de Ven, Kautsch. Gummi Kunstst. 52, 663 (1999).

37

N. Nakajima, E. R. Harrell, and E. A. Collins, RUBBER CHEM. TECHNOL. 50, 99 (1977).

38

M. Demaio, S. Caldari, and O. Chiantore, Int. J. Polym. Anal. Char. 5, 85 (1999).

39

C. A. Garcia-Franco, B. A. Harrington, and D. J. Lohse, Rheologica Acta 44, 591 (2005).

40

S. Mitra, M. Jorgensen, W. B. Pedersen, K. Almdal, and D. Banerjee, J. Appl. Polym. Sci. 113, 2962 (2009).

41

J. Leblanc, RUBBER CHEM. TECHNOL. 83, 65 (2010).

42

F. Ignatz-Hoover and B. H. To, Rubber Compounding Chemistry and Applications, B. Rodgers, Ed., Marcel Dekker, Inc.,
New York, 2004, p 505.

43

M. D. Ellul, J. Patel, and A. J. Tinker, RUBBER CHEM. TECHNOL. 68, 573 (1995).

44

R. E. Medsker, D. R. Hazelton, G. W. Gilbertson, S. Abdou-Sabet, K. S. Shen, R. L. Hazelton, and P. S. Ravishankar, U.S.


Patent 6,251,998 (to Advanced Elastomers L.P., Exxon Chemical Patents), June 26, 2001.

45

P. S. Ravishankar and N. R. Dharmarajan, Rubber World 219, 23 (1998).

46

H. Ebata, K. Ichino, and M. Kunizane, U.S. Patent Publication US 2009/0209672A1 (to Mitsui Chemicals Inc.), August
20, 2009

47

K. Masaaki, T. Toshiyuki, T. Tetsuo, and H. Takashi, EP 751 156 (to Mitsui Chemicals Inc.), 2001.

48

T. Abraham and C. McMahan, Rubber Compounding Chemistry and Applications, B. Rodgers, Ed., Marcel Dekker, Inc.,
New York, 2004, p 163.

348

RUBBER CHEMISTRY AND TECHNOLOGY, Vol. 85, No. 3, pp. 327349 (2012)

49

A. Y. Coran and R. P. Patel, RUBBER CHEM. TECHNOL. 54, 116 (1982).

50

S. Shahbikian, P. J. Carreau, M.-C. Heuzey, M. D. Ellul, H. P. Nadella, J. Cheng, and P. Shirodkar, RUBBER CHEM.

TECHNOL. 84, 325 (2011).


51

C. F. Antunes, A. V. Machado, and M. van Duin, Eur. Polym. J. 47, 1447 (2011).

52

M. D. Ellul, P. S. Ravishankar, J. Cheng, and P. E. McDaniel, U.S. Patent 8,178,625 (to ExxonMobil Chemical Patents,
Inc.), May 15, 2012.

53

M. D. Ellul, D. R. Hazelton, and P. S. Ravishankar, U.S. Patent 5,656,693 (to Exxon Chemical Patents Inc.), August 12,
1997.

54

M. D. Ellul and P. S. Ravishankar, New Highly Crosslinked TPEs Based on VNB-EPDM with Improved Compression
Set, Paper 74, presented at the 166th Meeting of the Rubber Division, ACS, Columbus, OH, October 58, 2004.

55

J. H. Dudas and W. C. Cochran, IEEE Elec. Ins. Mag. 15, 29 (1999).

56

S. Boggs and J. Xu, IEEE Elec. Ins. Mag. 17, 1 (2001).

57

M. Breen and V. M. Kothari, Overview of Insulating Material for Industrial Wire and Cable ProductsApplications,

Requirements, and Approval Process, Paper 14, presented at the 165th Spring Meeting of the Rubber Division,
ACS, Grand Rapids, MI, May 1719, 2004.
58

D.-C. Lee, R. Laakso, L. Gross, and J. Muskopf, Olefinic Elastomers for Wire and Cable Applications, paper presented

at the Proceedings of the 55th International Wire and Cable Symposium, Providence, RI, 2006, p 158.
59

J. L. Mead, Z. Tao, and H. S. Liu, RUBBER CHEM. TECHNOL. 75, 701 (2002).

60

G. J. Pehlert, N. R. Dharmarajan, and P. S. Ravishankar, Rubber World 226, 39 (2002).

61

G. Ver Strate and M. Struglinski, Polymers as Lubricating Oil Viscosity Modifiers, in Polymers as Rheology Modifiers,

Vol. 462, D. Schulz and J. Glass, Eds., ACS Symp. Ser., Washington, D.C., 1991.
62

H. G. Muller, Ang. Macrom. Chemie 67, 61 (1978).

63

G. Ver Strate, C. Cozewith, and S. T. Ju, Macromolecules 21, 3360 (1988).

64

G. Ver Strate, C. Cozewith, R. K. West, W. M. Davis, and G. A. Capone, Macromolecules 32, 3837 (1999).

65

E. P. Jourdain and G. J. Wouters, New Metallocene EPDMs for Expanded Attribute Range to Meet New Rubber Market

Requirements, paper presented at the International Rubber Conference, Nurnberg, Germany, 2009.
66

G. Choonoo and G. Vroomen, The Use of EPDM in Coolant Hoses Now and in the Future, paper presented at the

162nd Fall Meeting of the Rubber Division, ACS, Pittsburgh, October 811, 2002.
67

G. Stella and N. P. Cheremisinoff, Designing EPDM for Production Efficiency, paper presented at the International
Rubber Conference, Sydney, Australia, October 1988.

68

A. A. Galuska, Surface Interface Anal. 31, 177 (2001).

69

J. R. Schauder and G. Stella, Rubber World 207, 23 (1993).

70

O. Georjon, E. Jourdain, and A. Sahnoune, Kautsch. Gummi Kunstst. 58, 507 (2005).

71

S. Jacob, M. Hill, E. Jourdain, F. Zacarias, L. Dimitrijevs, D. Evanko, and J. Anderson, Innovative Solutions for Foamed
Thermoplastic Vulcanizate Automotive Weatherseals, paper presented at the Thermoplastic Elastomers Conference,
Akron, September 1214, 2005.

72

R. C. Keller and T. A. Mills, Kautsch. Gummi Kunstst. 44, 1032 (1991).

73

M. C. Bulawa and P. S. Ravishankar, Reduction of Peroxide usage in EPDM Coolant Hose via Polymer Redesign, paper
presented at the 155th Meeting of the Rubber Division, ACS, Chicago, April 1316, 1999.

74

P. S. Ravishankar and J. R. Schauder, U.S. Patent 5,766,713 (to Exxon Chemical Patents Inc.), June 16, 1998.

75

L. Korugic-Karasz and A. Vinci, Bull. Chem. Technol. Macedonia 21, 1 (2002).

76

S. Karam, M. Vincent, and Y. De Zelicourt, Int. Polym. Proc. 13, 209 (1998).

77

EPDM Membrane Production: Materials and Manufacturing Process, Technical Bulletin, EPDM Roofing Association,
http://www.epdmroofs.org/attachments/epdmmembraneproductionmaterialsandmanufacturingprocesses_era.pdf.
Accessed date August 1, 2012.

TREATISE ON EPDM

349

78

S. Datta, P. S. Ravishankar, and L. G. Kaufman, U.S. 5,654,370 (to Exxon Chemical Patents Inc.), August 15, 1997.

79

J. R. Schauder and G. Stella, Exxelore Modifiers for effective Polyamide Performance Enhancement, Polyamide 2001,

Dusseldorf, 2001.
80

H. C. Welborn, U.S. Patent 4,897,455 (to Exxon Chemical Patents Inc.), January 30, 1990.

81

H. W. Turner, G. A. Vaughan, R. A. Fisher, J. F. Walzer, C. S. Speed, B. J. Folie, and D. J. Crowther, U.S. Patent 5,767,208
(to Exxon Chemical Patents Inc.), June 16, 1998.

82

P. J. Chirik and J. E. Bercaw, Polym. Preprints 41, 393 (2000).

83

W. J. Wang, D. Yan, and S. Zhu, Polym. Reac. Eng. 7, 327 (1999).

84

C. B. Friedersdorf, U.S. Patent 7,163,989 (to ExxonMobil Chemical Patents Inc.), January 16, 2007.

85

Kumho Polychem Co. www.polychem.co.kr. Accessed date August 1, 2012.

86

P. S. Ravishankar, U.S. Patent 7,511,106 (to ExxonMobil Chemical Patents Inc.), March 31, 2009.

87

P. S. Ravishankar, U.S. Patent 7,943,711 (to ExxonMobil Chemical Patents Inc.), May 17, 2011.

88

P. S. Chum and K.W. Swogger, Prog. Polym. Sci. 33, 797 (2008).

[Received February 2012; Revised May 2012]

You might also like