Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

OTC-27609-MS

Nano-Enhanced Elastomers for Oilfield Applications

Rostyslav Dolog, Darryl Ventura, Valery Khabashesku, and Qusai Darugar, Baker Hughes Inc

Copyright 2017, Offshore Technology Conference

This paper was prepared for presentation at the Offshore Technology Conference held in Houston, Texas, USA, 1–4 May 2017.

This paper was selected for presentation by an OTC program committee following review of information contained in an abstract submitted by the author(s). Contents of
the paper have not been reviewed by the Offshore Technology Conference and are subject to correction by the author(s). The material does not necessarily reflect any
position of the Offshore Technology Conference, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Offshore Technology Conference is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of OTC copyright.

Abstract
The use of elastomers in oil industry extends over a broad range of applications including seals, packing
elements, reactive rubber elements, stators, and pads. These applications require a variety of property
requirements that may differ for dynamic and static applications or include a need for stimuli-responsive
capabilities in certain tools.
This research details the effect of nanofillers on elastomer properties for oil and gas components. The
effects include enhancement of mechanical properties, wear resistance, thermal conductivity and heat
expansion properties. In addition, effects of nanofillers on rapid gas decompression (RGD) resistance,
chemical resistance to downhole fluids, and resistance to chemical aging at downhole temperatures were
investigated.
Advanced rubber nanocomposites formulations, based on Hydrogenated Nitrile Butadiene Rubber
(HNBR) elastomers, were designed internally. Their properties were assessed using methods and techniques
to qualify elastomers for downhole applications. Mechanical properties of elastomers were evaluated at
room temperature and at 325° F, which is a maximum application temperature for HNBR elastomers. RGD
testing was conducted according to ISO standards.
Results indicated that it is possible to control mechanical properties of elastomers with nanotechnology,
including improving the abrasion resistance of the elastomers by more than 100% in dynamic, wear-
intensive applications, when compared to commercial compounds typically used in the oil industry. Thermal
conductivity was improved by up to 40%, while heat expansion decreased by 30%, providing more
versatility for seal design in dynamic applications which are prone to localized heating. In addition, RGD
resistance in nanocomposites was examined and compared it to control samples. The industrial scale
feasibility for nano-enhanced elastomers was demonstrated by a scale-up study.

Introduction
Due to the diversity of elastomers application in oil industry, a large portfolio of rubber compounds is
required to provide a certain set of properties for each specific need. The broad field of rubber chemistry
is continuously developing to meet the growing demands of the industry for reliable materials that
can withstand harsh downhole conditions, especially in high-temperature/high-pressure oilfields. Nitrile
elastomers (NBR, HNBR), fluoroelastomers (FKM, FEPM, and FFKM) and EPDM are among the most
common matrixes for rubber compounds used downhole (Campion 2005). Conventional reinforcing fillers
2 OTC-27609-MS

such as carbon black and silica have been used for many decades. However, nano-sized fillers such
carbon nanotubes (CNTs), graphene, nanoclays and nanosilica, expand capabilities for rubber compound
development, specifically as they relate to improved performance (Bhattacharyya 2008; Bokobza 2007;
Frogley 2003)
This study demonstrated the ability to adjust the mechanical properties of HNBR compounds with
nanotechnology. A detailed investigation was conducted of how properties of nano-enhanced elastomers
can be affected by changes in the curing system and mixing procedures. The latter often significantly
differ between laboratory conditions and large-scale manufacturing, where achieving high-quality mixing
can be problematic. Furthermore, this study explores how enhanced properties, demonstrated at ambient
conditions, translate to improved performance in downhole applications. In addition, the assessment of
thermal properties of nano-enhanced elastomers and their resistance to wear, downhole chemicals and
explosive decompression provides an understanding of their performance downhole.

Mechanical Properties
Many factors affect the final performance of a rubber product including the compound recipe, the mixing
process and curing conditions. A typical compound recipe consists of raw elastomer, filler, the curing system
and other additives that facilitate the mixing process, improve aging characteristics or provide new abilities
that are not inherent to elastomers. This study demonstrates how CNTs can be used to adjust the mechanical
properties of elastomer compounds, even when used in relatively small amounts. Furthermore, the effect
of mixing time and curative concentration on the properties of nano-enhanced elastomer compound is
investigated.
To study how CNT filler concentration affects the performance of elastomer compounds, a series of
samples (Series A) based on an HNBR elastomer with multi-wall CNTs and 50 phr of N550 carbon black
co-filler and peroxide-based curing system was prepared by melt mixing in a C.W. Brabender internal mixer
at 60 RPM for 10 minutes. Multiple studies in the literature reported that combining CNTs with carbon
black co-filler provided a synergistic effect on mechanical properties of rubber and mitigated the adverse
effect of CNTs on ultimate elongation (Perez 2009; Szeluga 2015).
Tensile testing was conducted at room temperature and at 325° F, according to the ASTM D412 standard,
using a MTS Criterion testing unit. The compression set of the samples was measured for 22 hours at 125°
C by compressing standard "1-inch" rubber button to 25% compression strain in a fixture according to the
ASTM D395 standard. Composition of the compounds and results of Shore A hardness and compression
set measurements are provided in Table 1.

Table 1—Samples with different concentration of CNTs

Sample Carbon black (phr) Nanofiller (phr) Hardness Shore A

A1 N550 (50) 0 70

A2 N550 (50) CNTs (5) 77

A3 N550 (50) CNTs (10) 86

A4 N550 (50) CNTs (15) 86

Fig. 1 graphically represents how tensile properties of nano-enhanced elastomers depend on the
concentration of CNT nanofillers. CNTs significantly improve the modulus and tensile strength of the
specimen, but the effect tends to have a smaller impact on the sample modulus and elongation with each
additional 5 phr of CNTs, especially at high temperature. This result is due to a certain saturation of the
sample with CNTs. As more CNTs are dispersed in the network, the quality of dispersion decreases. The high
OTC-27609-MS 3

cost of CNTs compared to traditional fillers such as carbon black or silica make it important to determine
at which point the incorporation of additional CNTs fails to be economical in each particular system.

Figure 1—Effect of CNTs on mechanical properties of HNBR system at room temperature and at 325° F

The compression set of the samples gradually increased with the increasing CNT loading. This behavior
can be explained by a negative effect of CNTs on the radicals generated by the curing system (e.g. capturing).
On the other hand, reinforcement provided by CNTs can also be attributed to the increase in the compression
set. Highly reinforced compounds have a high modulus and tend to have higher compression set due to
the to the polymer network damage caused by the high stresses necessary to achieve the same degree of
compression compared to the elastomer with lower filler loadings.
Curatives composition and concentration, in this case represented by a peroxide and a crosslinking
co-agent, largely define the crosslinking density of the rubber compound, which translates into different
mechanical properties and compression set. Samples with four different curative concentrations were
prepared and tested (see Table 2), and the ratio between the peroxide and the crosslinking co-agent was kept
constant. Properties of samples B1-B4 were evaluated at room temperature and at 325° F. From the results
of this testing one can conclude that when the concentration of peroxide curing agent exceeds 8 phr, tensile
strength decreases along with elongation at break. The optimal concentration of peroxide in the system
appears to be close to 8 phr, if achieving high tensile strength is the main goal. As expected, increasing the
amount of curing agent and co-agent (TAIC) also increased the modulus of the elastomer while elongation
and the compression set decreased. Rubber chemists usually set a goal to find an optimal amount of curing
components to achieve the highest possible tensile strength, without sacrificing the compression set yet
maintaining reasonable elongations and modulus. The compression set values have to be under 30% (or
even <20%) for the majority of sealing applications in oil and gas industry at or above 250° F.
4 OTC-27609-MS

Table 2—Samples with different concentrations of curing components

Sample Curing agent (peroxide, phr) Carbon black (phr) Nanofiller (phr)

B1 4 N326 (50) CNTs (5)

B2 6 N326 (50) CNTs (5)

B3 8 N326 (50) CNTs (5)

B4 10 N326 (50) CNTs (5)

The graphic summary of tensile testing results of samples B1-B4, provided in Fig. 2, demonstrates
how the concentration of curing agent affects mechanical properties of nano-enhanced elastomers at room
temperature and at 325° F.

Figure 2—Effect of curing agent concentration on mechanical properties of elastomers

Mechanical mixing promotes dispersion of the filler of the polymer matrix. However, mixing can be also
break down polymer chains and subsequently decrease the molecular weight, which negatively impacts
properties of the system. Samples B2, B5-B7 comprise the same peroxide curing system as samples A1-A4,
but they feature a different type of carbon black (N326) and a higher ACN content HNBR polymer (see Table
3). Varying the mixing time enables study of the effect of mixing on dispersion and polymer chain damage
to optimize processing conditions. Results demonstrate that quality of mixing can be steadily improved by
using longer mixing times; up to ~15 minutes of mixing. Further mixing causes chains to break. In most
cases the higher tensile strength and modulus of elastomer preserves the preferred elastomer performance.
OTC-27609-MS 5

Table 3—Samples with different concentration of CNTs

Sample Mixing time (min.) Carbon black (phr) Nanofiller (phr) Hardness Shore A

B5 5 N326 (50) CNTs (5) 80

B2 10 N326 (50) CNTs (5) 82

B6 15 N326 (50) CNTs (5) 81

B7 20 N326 (50) CNTs (5) 82

The graphic summary of tensile testing results of samples B2, B5-B7, provided in Fig. 3, demonstrates
how mixing time affects mechanical properties of nano-enhanced elastomers at room temperature and at
325° F.

Figure 3—Effect of mixing time on mechanical properties of HNBR system

Scale-up study
A scale-up study was conducted to investigate how the mixing efficiency changed during the transition from
laboratory scale mixing, typically conveyed in a C.W. Brabender internal mixer capable of mixing ~50-75
g batch, to mixing significantly larger batches that can be used for production of commercial quantities of
rubber products. On an industrial scale, rubber compounds are usually mixed in large internal mixers or
rubber calendar machines.
A Farrel internal mixer, capable of mixing batches up to ~1-1.5 kg, was used to simulate mid-level
scale-up of the compound mixing process at ~60 RPM. HNBR compounds were mixed in two-step mixing
procedure for at least 10 minutes; one compound contained 50 phr of carbon black, and another contained
50 phr of carbon black and 15 phr of CNTs fillers. HNBR polymer was mixed with CNTs, carbon black,
plasticizer and antioxidant during first step, and the curing system was added during second step. For
comparison, the same samples were mixed at laboratory scale in C.W. Brabender for 10 minutes at ~60 RPM.
6 OTC-27609-MS

At room temperature, the tensile properties of the compound that contained carbon black and CNTs
appeared to be slightly better when it was mixed in the Farrel Mixer. The tensile strength and modulus
were improved by ~10%, while elongation at break decreased ~5%, when compared to sample that was
mixed in C.W. Brabender, see Fig. 4. On the other hand, the compound that contained only carbon black
demonstrated ~10% decrease in ultimate elongation in the sample from the scale-up mixing process, but it
had comparable E25 modulus and tensile strength.

Figure 4—Comparison of compound tensile properties mixed in different mixers at room temperature (scale-up study)

Two competing factors affected the performance of these compounds. Mixing them in the small mixing
chamber of the C.W. Brabender was more efficient than in the much larger chamber of the Farrel mixer.
The small chamber enabled better dispersion of the filler particles, but the higher shear generated in the
small chamber had the potential to damage polymer chains. The Farrel mixer dispersed the filler particles
well and maintained intact polymer chains, leading to overall improvements in performance of elastomer
compounds at room temperature and at 325° F (see Fig. 5).

Figure 5—Comparison of tensile properties of compounds mixed in different mixers at 325° F (scale-up study)

Scaled-up production of elastomer compounds that contain CNTs fillers often raise concerns about
quality of nanoparticle dispersions in large-scale mixing processes and associated scalability of performance
OTC-27609-MS 7

improvements. However, it is demonstrated herein that with this particular compound it is possible to
maintain improved performance in large internal mixers, at least when mixing up to 1-1.5 kg of rubber.

Study of Properties Relevant to Oilfield Applications


The resistance of elastomers to liquid chemicals and gases, explosive decompression (also referred to as
RGD), wear, and their mechanical properties, compression set and thermal properties largely determine
their applicability in oilfield applications. Some properties are more relevant to dynamic applications, while
others are critical for long-term use in static environment. Thermal expansion, thermal conductivity and
heat build-up resistance are of significant importance for dynamic applications such as dynamic seals and
motor stators. A custom system based on an internally designed HNBR recipe was used to assess the effects
on nanofillers and different types of carbon black on the properties relevant to applications in oil industry.
Thermal conductivity determines the rate of heat dissipation through the seal. In dynamic applications,
some particular spots are often greatly affected by shear, causing overheating that can accelerate elastomer
aging and degradation. Higher thermal conductivity can extend the life of the seal in dynamic applications.
Thermal conductivity of CNTs is inherently very large, which makes them very efficient fillers for boosting
thermal conductivity of the seal. The incorporation of an additional 15 phr of CNT filler to an elastomer
compound containing 50 phr carbon black increased the thermal conductivity by ~40%, as shown in Fig.
6. C-Therm TCi Thermal Conductivity Analyzer with Modified Transient Plane Source (MTPS) technique
was used to assess the thermal conductivity of elastomers.

Figure 6—Thermal conductivity of elastomers with different fillers combination

Evaluation of heat build-up resistance was conducted according to ASTM 623 on a Goodrich Flexometer.
Two HNBR elastomer compounds were compared: the temperature of a sample with carbon black filler
(e.g., CB (50phr)) increased by 39° F, while temperature of a sample that was reinforced by carbon black
and CNTs (e.g., CB 50(phr)+CNTs(15phr)) increased the temperature by 47° F. This can partially offset
benefits of improved thermal conductivity.
Thermal expansion coefficients of the same HNBR compounds were assessed using a TA Instruments
BAHR DIL 806 and averaged between 30° C and 150° C. The results are summarized in Fig. 7. Although
it is possible to adjust the compound composition to design an elastomer system with a certain value of
thermal expansion coefficient, Fig. 7 does not show a straightforward general trend. This is due to several
competing factors: inherent heat expansion coefficients of fillers, change in free volume of the polymer
due to addition of nanofiller and increase in material's modulus due to reinforcement, which is inversely
proportional to heat expansion coefficient (Dolog 2016).
8 OTC-27609-MS

Figure 7—Coefficient of thermal expansion of elastomers with different fillers combination

In some dynamic applications elastomer wear can be the dominant mechanism of seal failure. The
elastomer industry takes advantage of a large number of carbon black grades that differ in particle size and
size distribution, particle structure, aggregation and functionalization (acidity). These properties of carbon
black fillers affect the properties of the final elastomer compounds. Carbon black grades that feature smaller
particle size, better-developed structure, and a higher number of functional groups on the surface usually
provide better reinforcement for the elastomer compound. The carbon black grade number is usually a three-
digit number, where first digit represents particle size (lower number = smaller particle size). Carbon black
grades, such as N234, are designed to provide superior abrasion resistance to elastomers. On the other hand,
this grade of carbon black creates a stronger interface with elastomer matrix and can significantly increase
heat buildup in dynamic applications.
Elastomer seals were subjected to a simple wear test against an abrasive surface for 30 minutes. A steady
stream of water was used to cool the surface and provide lubrication, and wear loss was recorded and plotted
in Fig. 8. CNTs provided superior wear resistance. Compound reinforcement with as little as 15 phr of CNTs
provided better resistance to wear than 50 phr of N234 carbon black. The difference in performance was
even more drastic when compared with 50 phr of general-purpose carbon black grades N550 and N326.
It is important to note that the sample that was reinforced by N550 was based on a lower ACN content
HNBR elastomer, which further contributed to the high volume loss due to wear. HNBR polymers with
lower ACN content are typically used to develop rubber compounds with higher elasticity (e.g. elongation at
break) but lower modulus and abrasion resistance. Based on the internal testing results, abrasion resistance
of the elastomer reinforced by CNTs was also more than 100% higher than abrasion resistance of one of the
commercial compounds typically used in the oil industry (not shown on Fig. 8).

Figure 8—Volume loss due to wear


OTC-27609-MS 9

Swelling tests of nano-enhanced elastomers in LVT200 oil and water brines, specifically 3% KCl and
10% CaCl2water solutions, at ambient pressure and 200° F confirmed that the swelling resistance of rubbers
was largely a function of elastomer matrix, and was not significantly affected by the presence of CNTs fillers.
Observed minor improvements in swelling resistance could only be attributed to the increased modulus of
the matrix, due to additional reinforcement provided by nanofiller.
Investigation of nano-enhanced elastomers' resistance to explosive decompression (so called RGD
resistance), and CSR test were proven to accurately reflect on the conditions that elastomers experience in
downhole applications, and were more representative tests than swelling tests.
In oilfield applications, when oil tools are removed to the surface from the downhole environment, there
is a steep drop in pressure. In this case pressured seals, saturated with downhole gasses can suffer from
explosive decompression, when gasses, dissolved in elastomer, escape. RGD test enables recreation of these
conditions in the laboratory and analysis of how susceptible the elastomers are to explosive decompression.
An RGD test was conducted internally following the ISO/FDIS 23936-2:2011 procedure using type -234-
O-rings, see Fig. 9. The test was performed at 200° F and ~1000 psi. Three specimens were tested for each of
the three HNBR samples with three different filler combinations: 1) 50 phr of N550 carbon black, 2) 50 phr
of N550 carbon black and 15 phr of CNTs, and 3) 50 phr of N550 carbon black and 15 phr of carboxylated
CNTs. Each O-ring was compressed to ~15% strain using a specially designed O-ring fixture and placed
into an autoclave where high temperature and pressure were imposed on a sample. The pressure in the vessel
was created with 90% nitrogen gas and 10% CO2 gas mixture. The sample was soaked in a reactor for 48
hours and then pressurized and depressurized every day for four days following the standard ISO procedure
at depressurization rate of ~290 psi/min.

Figure 9—Cross-sections of the most damaged O-rings left to right: 1) 50 phr of N550 carbon black 2) 50 phr
of N550 carbon black and 15 phr of CNTs 3) 50 phr of N550 carbon black and 15 phr of carboxylated CNTs

When the RGD test concluded, samples were depressurized, cooled down and removed from the vessel.
Each O-ring was later cut into four parts, and each cross-section was examined to determine the most
damaged one for each sample. That cross-section of the specimen was further studied under optical
microscope to investigate for possible blisters, holes and cracks. If a split was detected on a surface, the
sample would automatically be considered as failed. The length of the external and internal cracks for the
most damaged cross section was measured and the rating was assigned using guideline from ISO procedures.
Both nano-enhanced elastomer samples were assigned a ranking of "3", although these samples were not
optimized for a specific application, they passed the test according to the ISO guidelines.
Furthermore, the HNBR sample with 50 phr of carbon black and 15 phr of carboxylated CNTs was
subjected to a CSR test in two typical downhole fluids: NORSOK 702010 and HYCAL II for 2 weeks.
Rubber cylinders, 0.5-in. wide and 0.25-in. tall, were compressed in by ~20% and aged at 331° F and ~500
psi. The sealing force and stiffness of the samples were measured prior the test and after 2, 7, 14, and 21
days of aging, graphic summary is provided on Fig. 10. The CSR test is a practical method to measure the
10 OTC-27609-MS

sealing performance of elastomers in a complex downhole environment (Xiaobin 2016). The test enables
an understanding of the effect of temperature on the sealing force and stiffness of elastomeric material in an
aggressive fluid environment, and helps in predicting their reliability at application conditions. At ageing
conditions, overtime, due to slow chemical and/or mechanical degradation of rubber, material's stiffness
increases, and the sealing force decreases.

Figure 10—Results of CSR testing in two different fluids

The test demonstrated that even an un-optimized in-house compound is capable of developing and
maintaining high sealing force under accelerated aging conditions. This is an important property that
correlates with reliability in downhole conditions. It is worth noting that the initial drop (0 days) in sealing
force is typical for elastomers after immersion in downhole fluids, it is the behavior after the initial drop
which is usually considered by researches. Nano-enhanced compound was able to maintain the sealing force
during accelerated aging for at least three weeks, which was the duration of the test.

Conclusions
This study shows how nanotechnology can be used to improve performance of oilfield elastomers. A
detailed investigation was performed of how composition of nano-enhanced elastomers, as well as mixing
conditions, which determine quality of particle dispersion, affect tensile properties of elastomers at room
temperature and at 325° F. Nanotechnology provided a path to greatly improve thermal properties and wear
resistance of oilfield elastomers. Furthermore, un-optimized in-house developed elastomers demonstrated
decent performance in tests that resemble real-life oilfield conditions.
Another important issue which was addressed is regarding the translation of property enhancement,
provided by nanoparticles, from lab to the field. A scale up study demonstrated the feasibility of
manufacturing nano-enhanced elastomers, at least on the scale level of ~1 kg rubber batches.
OTC-27609-MS 11

Acknowledgements
The authors would like to thank the management of Baker Hughes for support and permission to publish
this paper. Authors also would like to thank Dennis Matuszak, David Livingston, Donna Colbert, Mikey
Benes, John Thurston, Don Quach, and Wayne Furlan (all are employees of Baker Hughes) for their support
and technical expertise.

Nomenclature
phr per hundred parts of rubber
CNTs carbon nanotubes
psi pound per square inch
RGD rapid gas decompression
HTHP high temperature high pressure
NBR nitrile butadiene rubber
HNBR hydrogenated nitrile butadiene rubber
ACN acrylonitrile
RPM rotations per minute
E25 stress at 25% strain (modulus), also designated in the literature as M25
CSR compressive stress relaxation
ISO International Organization for Standardization

References
Bhattacharyya, S., Sinturel, C., Bahloul, O., Saboungi, M.-L., Thomas, S., Salvetat, J.-P. 2008. Improving reinforcement
of natural rubber by networking of activated carbon nanotubes, Carbon 46 (7): 1037–1045. http://dx.doi.org/10.1016/
j.carbon.2008.03.011.
Bokobza, L. 2007. Multiwall carbon nanotube elastomeric composites: A review. Polymer 48 (17): 4907–4920. http://
dx.doi.org/10.1016/j.polymer.2007.06.046.
Campion, R.P., Thomson, B., and Harris, J.A. 2005. Elastomers for fluid containment in offshore oil and gas production:
Guidelines and review. Health and Safety Executive. HSE Books. Sudbury.
Dolog, R., and Khabashesku, V. 2016. Effect of Nanofillers on Thermal and Mechanical Properties of Oilfield Elastomers.
Presented at AIChE 2016 Annual Meeting.
Frogley, M.D., Ravich, D., Wagner, D.H. 2003. Mechanical properties of carbon nanoparticle-reinforced elastomers,
Composites Science and Technology, 63 (11): 1647–1654. http://dx.doi.org/10.1016/S0266-3538(03)00066-6.
Perez, L.D., Zuluaga, M.A., Kyu, T., Mark, J.E., Lopez, B.L. 2009. Preparation, Characterization, and Physical Properties
of Multiwall Carbon Nanotube/Elastomer Composites. Polymer Engineering and Science; Newtown 49. 5 (May 2009):
866–874.
Szeluga, U., Kumanek, B., Trzebicka B. 2015. Synergy in hybrid polymer/nanocarbon composites. Composites: Part A
73 (2015) 204–23. http://dx.doi.org/10.1016/j.compositesa.2015.02.021.
Xiaobin, H., Furlan, W., Quach, D., Thurston, J. 2016. Study of Compressive Stress Relaxation (CSR) to Evaluate the
Sealing Performance for Downhole Tools at High Temperature. Presented at 2016 Energy Polymer Group Winter
Technical Meeting.

You might also like