You are on page 1of 19

Research Article

Journal of Elastomers & Plastics


2021, Vol. 53(7) 8 06­–824
Chemically modified ª The Author(s) 2021
Article reuse guidelines:

soybean oils as plasticizers sagepub.com/journals-permissions


DOI: 10.1177/0095244320988159

for silica-filled e-SBR/Br journals.sagepub.com/home/jep

compounds for tire tread


applications
Viviane Meyer Hammel Lovison1,2,
Maurı́cio Azevedo de Freitas1
and Maria Madalena de Camargo Forte2

Abstract
Silica-filled styrene butadiene rubber (SBR)/butadiene rubber (BR) compounds plasti-
cized with mineral oils are mainly used to produce green tire treads. Previous works
have demonstrated that the partial replacement of naphthenic oil (ONAF) by bio-based
oils can provide processing and performance improvements for rubber compounds,
along with environmental benefits. In this study, two modified soybean oils (esterified,
OEST or esterified and epoxidized, OEPX) were investigated with the aim of evaluating
the complete replacement of ONAF and determining whether the chemical properties
of the oils affect the performance of silica-filled E-SBR/BR compounds, using the com-
pound with ONAF as a reference. The physical properties, curing characteristics,
morphology, and dynamic mechanical behavior were evaluated. The use of the modified
soybean oils decreased the optimal cure time while increasing the crosslink density and
the abrasive wear resistance. Further, the compounds with both modified soybean oils
showed a good balance of mechanical properties. The modified soybean oils decreased
the glass transition temperature of the rubber compounds, thus acting as true plastici-
zers. At 0 C, the tan d value of E-SBR/BR/OEPX increased relative to that of E-SBR/BR/
ONAF, whereas at 60 C, the values of the compounds with both modified soybean oils
showed slight increases. The tan d values reveal that compared with E-SBR/BR/ONAF, E-
SBR/BR/OEPX has better wet grip and a similar rolling resistance, whereas E-SBR/BR/

1
Instituto SENAI de Inovação em Engenharia de Polı́meros (ISI), São Leopoldo, Brazil
2
Post-Graduation Program in Mining, Metallurgic and Materials Engineering (PPGE3M), School of Engineering,
Universidade Federal do Rio Grande do Sul (UFRGS), Porto Alegre, Brazil

Corresponding author:
Maria Madalena de Camargo Forte, Post-Graduation Program in Mining, Metallurgic and Materials Engineering
(PPGE3M), School of Engineering, Universidade Federal do Rio Grande do Sul (UFRGS), Porto Alegre, RS
90650-001, Brazil.
Email: mmcforte@ufrgs.br
Lovison
2 et al. 807
Journal of Elastomers & Plastics XX(X)

OEST has a higher rolling resistance. Thus, both modified soybean oils can fully replace
ONAF and appear to be extremely attractive plasticizers for use in silica-filled E-SBR/BR
compounds employed as green tire treads for passenger cars.

Keywords
Plasticizer, soybean oil, silica, green tire tread, dynamic properties

Introduction
Over the last few decades, many innovative solutions have been proposed in the auto-
motive industry to improve fuel efficiency and to use sustainable new materials for car
components. One of these innovations is green tire technology, in which silica is used as
the main filler of rubber compounds, along with a bifunctional silane as a coupling agent.
This eco-friendly approach is justified by improvements in the key properties of the tire
treads and in fuel economy.1 Various researches have been aimed at improving green tire
performance in consideration of the required compromise among rolling resistance, wet
traction, and abrasive wear resistance.2–6 Green tire tread formulations for passenger
vehicles are typically based on styrene butadiene rubber (SBR) blended with high-cis
poly(butadiene), along with processing oils, fillers (mainly silica), crosslinking agents,
and additives. Ingredient selection and processing conditions have a large impact on the
key properties of tire treads as well as their environmental fingerprint.7
Plasticizers or extender oils, traditionally used as processing oils in rubber formula-
tions, are usually highly aromatic petroleum oils that have good compatibility with most
synthetic elastomers.8,9 However, owing to increasing market requirements regarding
health, safety, and the environment, the tire industry has been seeking new alternatives.
Highly aromatic petroleum oils have been replaced by treated extracts (low aromatic
hydrocarbon content and low toxicity) or oils from renewable sources.10,11 Vegetable
oils are considered one of the most important and versatile classes of oils from renewable
sources owing to their potential for chemical modification, good thermal and chemical
stabilities, biodegradability, universal availability, and low cost.11,12 These oils have the
key attributes required for use as processing aids to replace petroleum-based oils.
In the last 15 years, several patents13–17 have been filed claiming the use of crude or
chemically modified vegetable oils as extender oils in some elastomer grades or as
processing oils in elastomeric compositions. Moutinho et al.13 evaluated silica-filled E-
SBR/BR (70/30) compositions with soybean-oil-extended E-SBR (crude or epoxidized)
and reported shorter and longer optimum curing times, respectively, compared with the
composition using commercial naphthenic oil (ONAF). Both E-SBR/BR/soybean oil
compositions showed higher elongation at break, lower moduli, and greater tear and
abrasion resistances. Sandstrom et al.16 compared the use of soybean and mineral oils in
silica-filled S-SBR/BR compositions for tire treads, with S-SBR obtained by a suspen-
sion process, and reported a higher tear resistance when using soybean oil. Hattori et al.17
reported that soybean oil could replace the aromatic oil plasticizer in natural rubber/SBR
808
Lovison et al. Journal of Elastomers & Plastics 53(7)
3

or epoxidized natural rubber/SBR compositions for tire production without compro-


mising the rolling resistance, abrasive wear resistance, and wet grip.
Various studies18–24 have reported the effect of different crude and chemically
modified vegetable oils on the plasticization of rubber compositions. Chandrasekara
et al.18 employed epoxidized sunflower oil as a plasticizer in carbon-black-filled natural
rubber and found that the oil could act as an accelerator to decrease the optimum cure
time. Kukreja et al.20 investigated the effect of castor oil on filler–rubber interactions in
natural rubber compounds reinforced with carbon black. At contents up to 3 phr, castor
oil acted as a coupling agent and plasticizer, whereas at higher contents, its main role was
plasticization, and the overall properties of the composition were not improved. Bezerra
et al.21 and Da Costa et al.22 observed the activation effects of linseed and peanut oils in
unfilled natural rubber compounds and reported that the vegetable oils can form ZnO/
sulfur/accelerator complexes. Zanchet et al.23 compared soybean oil and ONAF as
plasticizers for silica-filled natural rubber compounds with a bifunctional silane coupling
agent. A soybean oil content of 15 phr caused a decrease in the glass transition tem-
perature (Tg), silica particle agglomeration, and changes in the compound morphology.
Espósito and Marzocca24 evaluated a tire tread rubber consisting of a silica-filled S-SBR
compound plasticized with epoxidized soybean oil (ESO). The silanization conditions
and the sequence of silane/ESO addition during mixing were found to have a strong
influence on the final properties of the rubber compound.
Previous studies have demonstrated that the partial replacement of mineral oils, such
as ONAF, the most frequently used commercial plasticizer, by bio-based oils or soybean
oil can improve the processing and performance of natural and butadiene-based rubber
compounds, and can also provide health and environmental benefits.7,9,25 Furthermore,
the use of crosslinked soybean oil (factice) as a processing aid results in improved ozone
resistance, whereas epoxidized soybean oil used as a plasticizer enhances the low- and
high-temperature properties as compared to rubber compounds with mineral oils. The
choice of soybean oil, obtained from a renewable source by cultivation and harvesting of
seeds, as a green plasticizer is due in part to it being the second-most-produced oil among
vegetable oils,26 with sufficient availability to meet the demand of pneumatic industries.
Soybean oil is also preferentially cited in patent literature13–17,24 aimed at industrial use.
Furthermore, the good affinity between soybean oil and rubber compositions reinforced
with silica can be modified, because soybean oil can also interact with silica. In this
work, modified soybean oil was used as the sole plasticizer in an E-SBR/BR (70/30)
compound reinforced with silica for use in green tire treads, using a rubber compound
with only ONAF as a reference. Specifically, two chemically modified soybean oils, one
esterified (OEST) and the other esterified and epoxidized (OEPX), with different
functional groups and acid values were evaluated as plasticizers. The complete
replacement of ONAF with soybean oils was investigated with the aim of determining
whether the chemical properties of the soybean oils affect the performance of the rubber
compounds. The prepared rubber compositions only differed regarding the plasticizer
type. The cure characteristics, morphologies, and physical/mechanical properties of the
rubber compounds before and after vulcanization were investigated in detail. The
polymer–filler interactions were evaluated based on the bound rubber content.
Lovison
4 et al. 809
Journal of Elastomers & Plastics XX(X)

Requirements for good tire tread performance, such as wet grip (wet traction), rolling
resistance, and wear resistance, were assessed by investigating the viscoelastic properties
of the cured E-SBR/BR compounds.

Experimental
Materials
Poly(styrene-co-butadiene) rubber (E-SBR; styrene ¼ 23.5%, viscosity ¼ 51 MU; Buna
SE 1502) and high-cis poly(butadiene) (BR; 1,4-cis units ¼ 96.0%, viscosity ¼ 44 MU;
Buna CB 24) were obtained from Arlanxeo Brasil S/A (Brazil). ONAF (acid value ¼
0.02 mg KOH�g�1, specific density at 25� C ¼ 0.872 mg�cm�3) was obtained from quantiQ
Distribuidora Quı́mica Ltda (Brazil). OEST (acid value ¼ 1.68 mg KOH�g�1, specific
density at 25� C ¼ 0.9154 g�cm�3) and OEPX (acid value ¼ 0.71 mg KOH�g�1, specific
density at 25� C ¼ 0.955 g�cm�3, oxirane value ¼ 6.6 wt%) were obtained from SGS do
Brasil S/A (Brazil) and Nexoleum Bioderivados S.A. (Brazil), respectively. ONAF, OEST,
and OEPX were used as plasticizers (PLAST). High-dispersion silica (BET surface area ¼
208.2 m2�g�1; Zeosil Premium 200MP) was obtained from Rhodia Solvay S/A (Brazil).
Triethoxysilylpropyl tetrasulfide (TESPT), used as the silane coupling agent, was obtained
from Struktol Company of America (Brazil). Carbon black N339 was obtained from Cabot
(Brazil). Stearic acid, N-(1,3-dimethylbutyl)-N0 -phenyl-p-phenylenediamine (6PPD),
2,2,4-trimethyl-1,2-dihydroquinoline (TMQ), zinc oxide (99.9%), and masterbatches of
diphenylguanidine (DPG) (80%), N-cyclohexyl-2-benzothiazole sulfenamide (CBS)
(80%), and sulfur (80%) were kindly donated by Unique Rubber Technologies Ltda
(Brazil).

Preparation of E-SBR/BR/PLAST compounds


The rubber compounds (E-SBR/BR/PLAST) were prepared in an internal mixer (Haake
Polylab 02 RheoDrive 7, Thermo Scientific) equipped with tangential rotors at a 62%
filling factor according to the formulations shown in Table 1. The plasticizers (oils) were
chemically assessed via Fourier transform infrared (FT-IR) spectroscopy using a
Spectrum One spectrometer (PerkinElmer) in transmittance mode (32 scans, 4 cm�1
resolution). Thermogravimetric analysis (TGA; TA Q500 Thermobalance, TA Instru-
ments) of the plasticizers (*20 mg) was performed from 25 to 900� C at a heating rate of
10� C�min�1 under a nitrogen atmosphere. The mixing procedure was based on that
reported by Dierkes27 for rubber compounds filled with silica and consisted of three
steps, as summarized in Table 2. In step 1, the polymers, the silane coupling agent, silica
(3 � 1/3), plasticizer (2 � 1/2), and carbon black (pigment) were added. The coupling
agent and silica were added together to facilitate silica silanization and dispersion. In
step 2, the protective agents (TMQ and 6PPD), stearic acid, and zinc oxide were added.
The control criteria for ingredient addition were torque stabilization and time. The
mixing chamber was cooled for 40 min after each step. In step 3, the plasticized rubber
compound was transferred to a 2.5 L two-roll mill (Copé and Cia Ltda, Brazil) for
the addition of the curing agents at 50� C and subsequent lamination. The uncured
810
Lovison et al. Journal of Elastomers & Plastics 53(7)
5

Table 1. E-SBR/BR/PLAST compound formulations (amounts in phr).

E-SBR/ E-SBR/BR/ONAF E-SBR/BR/ E-SBR/BR/


Ingredient BR-blank (Reference) OEST OEPX

E-SBR 70 70 70 70
High-cis BR 30 30 30 30
Silica — 80 80 80
Carbon black — 7 7 7
Silane — 7.2 7.2 7.2
ONAF — 30 — —
OEST — — 30 —
OEPX — — — 30
Zinc oxide 3 3 3 3
Stearic acid 1 1 1 1
TMQ 2 2 2 2
6PPD 2 2 2 2
Sulfur 2.5 2.5 2.5 2.5
CBS 2.12 2.12 2.12 2.12
DPG 2.5 2.5 2.5 2.5

Table 2. Mixing procedure for E-SBR/BR/PLAST compounds.

Addition Time Rate


Mixture order (min) Ingredient (rpm) T (� C)

Step 1 i 0.5 Polymers 70 80


Mixture 1 ii 1.0 1=3 silica þ silane

iii 1.5 1=3 silica

1=
iv 2.0 3 silica þ ½ plasticizer

v 3.0 Carbon black þ ½ plasticizer 90 145–160


vi Discharge
Step 2 i 0.5 Mixture 1 70 80
Mixture 2 ii 2.0 Stearic acid þ 6PPD þ TMQ 90 145–160
iii 1.5 ZnO
iv Discharge
Step 3(a) i 5.0 Mixture 2 — 50
Rubber compound Sulfur þ DPG þ CBS

(a) In two-roll mill.

compositions were refrigerated at 4–5� C to prevent prevulcanization. For comparison


purposes, a compound without any oil and fillers (E-SBR/BR-blank) was also prepared.

Characterization of E-SBR/BR compounds


Cure characteristics. The cure characteristics of the E-SBR/BR compounds were assessed
using a rubber processing analyzer (RPA 2000, Alpha Technologies) at 160� C for 30 min
Lovison
6 et al. 811
Journal of Elastomers & Plastics XX(X)

with a frequency of 100 cycles�min�1 and an angle of 0.5� , according to the test pro-
cedure described in ASTM D 5289.

Vulcanization and specimen preparation. The E-SBR/BR compounds were vulcanized in a


compression press at 160� C for the time required to achieve 90% curing (t90), depending
on the sample geometry: t90 þ 20% and 100 kgf�cm2 for 2 mm sheets or t90 þ 50% and
30 kgf�cm2 for abrasive wear resistance specimens. The specimens were tailored
according to the standard of each test method.

Swelling. The crosslink density of each E-SBR/BR compound was estimated based on the
swelling degree (Q) in toluene. First, the vulcanized specimen (10 mm � 2 mm) was
washed with acetone in a Soxhlet extractor for 8 h to extract volatiles, according to
ASTM D 297 standard item 19. The weight of the dried nonswollen specimen (initial, mi)
was recorded and then the specimen was immersed in toluene at 23� C for 72 h (time to
attain swelling equilibrium) in the absence of light. After removing from toluene, the
specimens were quickly dried and weighed (final, mf). The swelling degree (Q) was
calculated as the ratio between the swollen and initial masses (Q ¼ [(mf � mi)/mi � 100])
and expressed as the average + standard deviation for three measurements.

Bound rubber. The bound rubber content (BdR) in the E-SBR/BR compounds was
measured according to the method of Gheller et al.28 Briefly, the samples without
curatives were Soxhlet extracted using toluene and the insoluble portion was subjected to
TGA (TA Q500 Thermobalance, TA Instruments). The calculation (%BdR ¼ (phr �
Dextracted)/residue) accounted for the filler quantity (phr), the weight loss between 250
and 700� C (Dextracted), and the residue at 700� C. The results were expressed as the
averages + standard deviations for three measurements.

Morphology. The morphologies of the E-SBR/BR compounds were assessed using atomic
force microscopy (AFM; XE7, Park Systems) in noncontact mode with 10 mm magni-
fication and a scan rate of 0.7 Hz at room temperature. The AFM images were recorded
perpendicular to the surface of the samples and both topography and phase images were
recorded.

Hardness (Shore A). The hardness of the E-SBR/BR compounds was measured on a digital
durometer (Shore A, Bareiss 64662 series) according to the ASTM D 2240 guidelines.
The test was performed within 1 s at different points on the samples and the results were
expressed as the averages + standard deviations of five measurements.

Tensile and tear tests. The tensile strength and tear resistance of the E-SBR/BR com-
pounds on type C specimens were determined using a universal testing machine (EMIC
DL 5000, Instron), according to the ASTM D 412 and D 624, respectively, with a 1 kN
load cell and a speed of 500 mm�min�1. The results were expressed as the averages +
standard deviations of five. The tensile strength (stress at break), elongation at break, and
moduli (stresses values) at 100% and 300% elongation were determined.
812
Lovison et al. Journal of Elastomers & Plastics 53(7)
7

DIN resistance to abrasion. The abrasive wear resistance of the E-SBR/BR compounds was
evaluated using an abrasion machine (Maqtest) according to DIN ISO 4649–method A.
The results were expressed as the averages + standard deviations of five measurements.

Dynamic mechanical analysis. The dynamic mechanical properties of the vulcanizates and
E-SBR/BR compounds, including the elastomeric blend E-SBR/BR-blank, were eval-
uated using a dynamic mechanical analyzer (DMA 25/50, Metravib) at a frequency of 10
Hz and a strain amplitude of 0.089% in tensile mode. The measurements were recorded
in the temperature sweep range of �70 to 80� C at a heating rate of 3� C�min�1 according
to ASTM D 4065 and D 5992. The storage modulus (E0 ) and tan d curves were recorded.
The tan d values at 0 and 60� C were used as predictors of the wet traction and rolling
resistance, respectively, of the compositions if applied for tire tread production. The Tg
value was evaluated as the temperature at the apex of the tan d curve. The results were
expressed as average values + standard deviations of two measurements.

Statistical analysis. One-way ANOVA (p < 0.05) and the Tukey’s range test were applied
to determine the significance of variance for the tan d values.

Results and discussion


The content of oil (e.g., plasticizer, extender oil, or processing aid) in rubber varies (*2–
10 phr) depending on the intended purpose. Because of the high silica content in the
formulation, a high plasticizer amount (30 phr) was necessary in this work to achieve the
required hardness for tire tread rubber (60–64 Shore A29). The SBR was obtained by an
emulsion process (E-SBR) and not oil-extended, to avoid the influence of any oils other
than the soybean oils of interest. The Tg value of the rubber compound without oil and
fillers (E-SBR/BR-blank) was evaluated to determine the effect of the soybean oils as
plasticizers on the vulcanized compounds. The plasticizer, which reduces the compound
viscosity, facilitates mixing, and increases the extrusion rate, is one of the major
ingredients in a rubber compound and should be compatible with the other chemical
ingredients. Although plasticizers can also affect the properties of cured rubber, such as
Tg, hardness, elongation, and modulus, it is desirable that the negative effects on the
finished product are insignificant. The performance of the modified soybean oils (OEST
and OEPX) was evaluated relative to that of ONAF based on the plasticizer character-
istics, cure parameters, swelling, morphology, polymer–filler interactions, stress–strain
and dynamic behavior, and tire tread performance. Samples of the vulcanized com-
pounds E-SBR/BR/PLAST that underwent acetone extraction (8 h) showed similar
plasticizer content (OEST: 9.1 wt%; OEPX: 10.0 wt%; ONAF: 10.5 wt%).

Plasticizer characteristics
The petrochemical oil ONAF, which is the most employed rubber plasticizer, is a
hydrocarbon containing saturated cyclic alkanes. In terms of polarity and saturation,
ONAF is intermediate between paraffinic and aromatic oils. The acid value is an important
Lovison
8 et al. 813
Journal of Elastomers & Plastics XX(X)

Figure 1. FT-IR spectra of naphthenic oil (ONAF), esterified soybean oil (OEST), and esterified
and epoxidized oil (OEPX).

indicator of the oxidation degree of an oil and provides a measure of the free fatty acids
present in a substance. ONAF is not strongly oxidized (acid value ¼ 0.02 mg KOH�g�1),
whereas the soybean oils OEST and OEPX (acid value ¼ 1.68 mg KOH�g�1 and 0.71 mg
KOH�g�1, respectively) are oxidized hydrocarbons. Moreover, OEPX has fewer free fatty
acids than OEST and has oxirane oxygens (6.6 wt%). Owing to their hydrocarbon
structures, all three oils exhibit similar FT-IR bands at 2800–2900 cm�1 corresponding to
asymmetric and symmetric C–H bond stretching in groups –C–H(CH3) and –C–H(CH2)
groups (Figure 1). Moreover, both the soybean oils show C¼O and C–O absorption bands
at 1744 and 1166 cm�1, which are related to ester groups. Thus, the main difference
between the soybean oils and ONAF is the presence of double bonds and polar groups.
While the FT-IR spectrum of OEST shows cis-n(C¼C) and cis-n(¼C–H) absorption
bands at 1654 and 3009 cm�1, respectively, that of OEPX shows a n(C–O–C) absorption
band at 843 cm�1. Thus, the main chemical difference between the soybean oils is the
presence of oxirane groups, which is plausible because double bonds are converted into
epoxy groups by epoxidation, as reported by Jebrane et al.30 These results indicate that the
modified soybean oils are polar, whereas the petrochemical oil is nonpolar. The interac-
tions of each plasticizer with the other ingredients in the formulation may be the same or
different, which may affect the morphology and properties of the rubber compound.
The thermal degradation of the soybean oils occurs via two events, whereas ONAF
degrades fully at 290� C (mass loss ¼ 100%) (Figure 2). The first and second maximum
rates of degradation in the TGA curves occur at 242� C (mass loss ¼ 32%) and 404� C
(mass loss ¼ 68%), respectively, for OEST and at 270� C (mass loss ¼ 78%) and 376� C
(mass loss ¼ 22%), respectively, for OEPX. Despite the events occurring at similar
temperatures, the mass losses for OEST and OEPX differ considerably. The thermal
degradation profile of an organic compound depends on the physical structure, the
molecule molar mass, the intermolecular forces, and chemical reactions that may occur
814
Lovison et al. Journal of Elastomers & Plastics 53(7)
9

Figure 2. TGA mass loss curves of naphthenic oil (ONAF), esterified soybean oil (OEST), and
esterified and epoxidized oil (OEPX).

while increasing the temperature. Thus, the different degradation profiles confirm that
these hydrocarbon oils are chemically different. Both the soybean oils are thermally
stable up to 200� C, as no mass loss was detected below this temperature, making them
suitable for use at rubber processing temperatures (< 200� C), similar to ONAF.

Rheological parameters and swelling


The E-SBR/BR/PLAST vulcanization process can be followed by the rheological curves
shown in Figure 3, and the characteristic rheological parameters are summarized in
Table 3. E-SBR/BR/ONAF and E-SBR/BR/OEST show similar curing curves. Although
the curing curve of E-SBR/BR/OEPX exhibits higher torque values, the minimum torque
was similar for all the E-SBR/BR/PLAST compounds. The maximum torque (MH,
S0 max) is associated with the elastic modulus of the vulcanized compound. Further, the
difference between the maximum and minimum torques (MH � ML ¼ DM) is related to
the crosslink density, which is inversely related to the rubber swelling (in a suitable
solvent).31 The slopes of the rheological curves indicate that the modified soybean oils
accelerate the vulcanization/curing reaction. E-SBR/BR/ONAF and E-SBR/BR/OEST
have similar MH and DM values but significantly different optimum curing times (t90;
9.9 and 7.3 min, respectively), indicating that the soybean oil has an advantage when
compared with the petrochemical oil. E-SBR/BR/OEPX has the highest MH and DM
values and the lowest t90 (6.6 min). The lower t90 indicates more favorable curing
kinetics, which may be due to an accelerating effect resulting from chemical interactions
between the oxirane (–C–O–C) and silanol (Si–OH) groups of OEPX and the filler,
respectively. Such interactions prevent the adsorption of accelerator molecules onto the
surfaces of silica particles, as described by Kim et al.32 for silica-filled epoxidized SBR.
Lovison
10 et al. 815
Journal of Elastomers & Plastics XX(X)

Figure 3. Rheological curves for E-SBR/BR/PLAST compounds.

Table 3. Rheological parameters of curing and swelling.


E-SBR/BR/PLAST compound

Rheological parameter ONAF OEST OEPX

t90 at 160� C, min 9.9 7.3 6.6


ML (S0 min), dN�m 2.5 2.9 2.5
MH (S0 max), dN�m 18 17.6 21.4
DM (MH � ML), dN�m 15.5 14.7 18.9
Swelling, % (average + std. deviation) 191 + 0.15 215 + 0.81 183 + 0.39

In the case of E-SBR/BR/OEST, the presence of double bonds leads to a high con-
sumption of the curing agent, causing not only a higher t90 but also a lower crosslink
density, as the MH value is lower than that of E-SBR/BR/OEPX.
Considering that all the E-SBR/BR/PLAST compounds have same composition with
the exception of the plasticizer type, the swelling behavior can be used to predict the
crosslink density of the elastomers. As shown in Table 3, E-SBR/BR/OEPX exhibits the
lowest swelling degree, followed by E-SBR/BR/ONAF and E-SBR/BR/OEST. Thus, it
can be inferred that E-SBR/BR/OEPX has a higher crosslink density, probably owing to
the adsorption of less accelerator onto the surfaces of silica particles, enabling a greater
acceleration of the elastomer crosslink reaction. The observation of the highest crosslink
density for E-SBR/BR/OEPX is consistent with its higher MH and DM values.

Morphology and polymer–filler interactions


Figure 4 shows the AFM 3D topographical images and corresponding phase images for
E-SBR/BR/ONAF, E-SBR/BR/OEST, and E-SBR/BR/OEPX. During sample surface
816
Lovison et al. Journal of Elastomers & Plastics 53(7)
11

Figure 4. 3D topographical (1) and phase images (2) of the E-SBR/BR/PLAST compounds:
(a) ONAF, (b) OEST, and (c) OEPX.

scanning, the phase oscillates according to the interaction between the tip and the spe-
cimen surface, allowing differentiation between sample domains. The darker and lighter
phases in the AFM 3D topographical images are related to the more ductile and more
rigid components of the sample, respectively, as reported elsewhere,33–35 with the lighter
phase being composed of silica (less tip penetration). The morphologies of all the
Lovison
12 et al. 817
Journal of Elastomers & Plastics XX(X)

compounds consist of silica particles (lighter phase) dispersed in the nonpolar E-SBR/
BR elastomer (darker phase), but the extent of dispersion differs. The dispersion of silica
depends on the action of the silane coupling agent (TESPT) and a high shear during
component mixing, as reported by Perez and Lopez.36 E-SBR/BR/ONAF and the E-
SBR/BR/OEST exhibit better dispersion, whereas E-SBR/BR/OEPX contains some
silica agglomerates. This particle agglomeration is possibly due to the higher polarity of
the plasticizer OEPX, which facilitates chemical interactions between the epoxy groups
of OEPX and the hydroxyl groups of silica (80 phr). To avoid interference during the
silica–silane chemical reaction, silane was the first component added to the elastomers
(Table 2) followed by silica, although the reaction between these components occurs at
145 C. An OEPX–silica interaction, related to the nucleophilic attack of the reactive
sites of the silica surface on the oxirane groups,24 may decrease the availability of silanol
moieties for coupling with the silane coupling agent, which is used to improve silica
dispersion in the elastomeric matrix. This behavior may cause undesirable filler–filler
interactions and silica particle agglomeration.
The phase images of the samples show good silica dispersion in all the elastomeric
matrixes. The lighter phase image for E-SBR/BR/OEST (Figure 4b2) reveals a higher
homogeneity along the sample surface, i.e. lower dark/lighter contrast, than for the other
samples. Robertson et al.,37 who evaluated the dark/light contrast in silica-filled SBR
compounds with a silane coupling agent, reported that less contrast corresponds to better
filler dispersion in the elastomeric matrix. Thus, it is suggested that OEST favors silica
(silica–silane) dispersion in the E-SBR/BR matrix rather than direct chemical interac-
tions with silica particles, whereas such interactions occur with OEPX, as indicated by
the enlarged lighter phases in the 3D and phase images. Polymer–filler interactions were
evaluated based on the bound rubber contents, which were 53.0% (+0.6) for E-SBR/BR/
ONAF, 50.1% (+1.2) for E-SBR/BR/OEST, and 38.1% (+0.2) for E-SBR/BR/OEPX.
The low bound rubber content for E-SBR/BR/OEPX corroborates the occurrence of
interactions between the oxirane groups of OEPX and the hydroxyl groups of silica on
the filler surface, while avoiding chemical reactions between the hydroxyl groups and
the silane (silanization). In contrast, the bound rubber contents of E-SBR/BR/OEST
(50%) and E-SBR/BR/ONAF (53%) were remarkably similar. This observation indicates
that OEST favors the dispersion of silane–silica particles in the elastomeric matrix, what
is supported by the topographical 3D and phase images.

Stress–strain behavior
The mechanical properties (tensile strength, elongation at break, and moduli at 100 and
300%), tear strength, abrasive wear resistance, and hardness of the E-SBR/BR/PLAST
compounds are shown in Table 4. For all the compositions, the property values were of
the same order of magnitude with slight differences. The oil type did not significantly
affect the tensile strength (*19 MPa), 100% and 300% moduli (*2.3 and *9.3 MPa,
respectively), and hardness (*63 Shore A). The E-SBR/BR/PLAST compound with
ONAF exhibited a lower elongation a break value (450%) than the modified soybean oils
(*570%) but there was no significant difference between the values for the E-SBR/BR
818
Lovison et al. Journal of Elastomers & Plastics 53(7)
13

Table 4. Physical and mechanical properties of E-SBR/BR/PLAST compounds.

E-SBR/BR/PLAST

Property (average + std. deviation) ONAF OEST OEPX

Tensile strength, MPa 17 + 2 19 + 1 20 + 0.5


Elongation at break, % 444 + 48 585 + 60 559 + 20
Modulus @ 100%, MPa 2.4 + 0.4 2.4 + 0.4 2.2 + 0.1
Modulus @ 300%, MPa 9.9 + 1.2 8.5 + 0.8 9.3 + 0.1
Hardness, Shore A 62 + 0.1 63 + 0.7 64 + 0.4
Tear strength, N�mm�1 42 + 0.4 55 + 4 53 + 3
Relative volume loss, mm3 (Abrasive wear) 105 + 4 86 + 1 100 + 1.0

compounds with modified soybean oils. However, some differences were observed in the
tear strength and abrasive wear resistance. The tear strength (force required to cause a
rupture) increased by approximately 30% when ONAF (42 N�mm�1) was replaced by the
modified soybean oils (*55 N�mm�1). This highest tear strength for the E-SBR/BR/
PLAST compounds with modified soybean oils could be associated with the higher
elongation at break values. Similar physical/mechanical behavior has also been reported
by Moutinho et al.13 for silica-filled E-SBR/BR rubber compounds when E-SBR was
extended in vegetable oil instead of ONAF. In terms of abrasive wear, E-SBR/BR/OEST
exhibited a significantly lower volume loss (22%) than the other two rubber compounds.
The higher wear resistance (lower relative volume loss) could be related to the lower Tg
of E-SBR/BR/OEST (see below). These results show that the use of the modified soy-
bean oils instead of ONAF did not change the main properties of the silica-filled E-SBR/
BR compound. Therefore, depending on the intended application of the rubber com-
pound, ONAF can be replaced by modified soybean oils without property loss.

Dynamic properties and tire tread performance


Figure 5 shows the storage modulus (E0 ) and tan d or loss factor curves in the range of �70
to 80� C for the E-SBR/BR/PLAST compounds. For the elastomeric blend without oil and
filler (E-SBR/BR-blank), only the tan d curve was considered for the determination of Tg.
The E0 and tan d values at 0 and 60� C as well as Tg for the E-SBR/BR/PLAST compounds
are listed in the Table 5. A sharper decrease in E0 was observed for E-SBR/BR/ONAF and
E-SBR/BR/OEST, whereas E-SBR/BR/OEPX showed a smoother decrease in stiffness,
which could be associated with its higher crosslink density and a filler network that
restricts energy dissipation via macromolecular segment motion. The glass transition
region of the E-SBR/BR elastomer blends occurs between �70 and �10� C (with Tg
defined by the tan d peak), in which E0 decreases drastically from 109 to 2–4 � 107 MPa.
At temperatures below Tg, a high compound viscosity and a small free volume fraction
hinder macromolecular segment movement, resulting in a rigid solid with a high elastic
modulus and low energy dissipation.38 In the range of �10 to 80� C, all the compounds
exhibit similar behavior; this range corresponds to the rubbery plateau the elastomer
Lovison
14 et al. 819
Journal of Elastomers & Plastics XX(X)

Figure 5. Storage modulus (E0 ) (a) and tan d (b) curves of the E-SBR/BR/PLAST compounds.

Table 5. E0 , tan d, and Tg values of the E-SBR/BR/PLAST compounds.

E-SBR/BR/PLAST

Properties ONAF OEST OEPX

E0 at 0� C, MPa 28.9 32.2 39.2


E0 at 60� C, MPa 13.3 12.8 13.1
Glass transition temperature (Tg)a, � C �50 + 0.6 �56 + 0.6 �53 + 1.8
Tan d at 0� C (wet grip)b 0.149 + 0.0003 0.148 + 0.001 0.164 + 0.0006
Tan d at 60� C (rolling resistance)c 0.131 + 0.003 0.148 + 0.002 0.139 + 0.003
a
Tg (E-SBR/BR-blank) ¼ �42 + 2.3� C.
b
The differences among the samples for tan d 0� C were analyzed using ANOVA (p ¼ 0.0007).
c
The differences among the samples for tan d 60� C were analyzed using ANOVA (p ¼ 0.022302).

application region. In this temperature range, the variation in E0 is low owing to macro-
molecular crosslinking, with a slight decrease occurring with increasing temperature.
All the tan d curves of the E-SBR/BR/PLAST compounds (Figure 5b) are shifted to
lower temperatures compared with that of E-SBR/BR-blank owing to the effects of the
820
Lovison et al. Journal of Elastomers & Plastics 53(7)
15

plasticizers, and the apex of these curves are taken as Tg. ONAF, OEPX, and OEST at the
same oil content (30 phr) decreased the Tg of E-SBR/BR-blank (42� C) by approximately
8, 11, and 14� C, respectively. As plasticizers, the oil molecules interact chemically with
the polymer chains, reducing the intermolecular force density, thus increasing the free
volume fraction, as described by free volume theory,39,40 and reducing Tg.
The lower Tg of E-SBR/BR/OEST is in accordance with its higher abrasive wear
resistance, which is a generally accepted correlation.25 In addition to shifting the tan d
curve peaks, the oils also decreased the tan d peak height. The tan d peak height is
associated with the hysteresis of a compound, i.e. the balance between losing and storing
energy (tan d ¼ E00 /E0 ). In a rubber compound, many parameters, including polymer
microstructure, blend composition, crosslink density, the presence of plasticizers, the
amount of reinforcing filler, and coupling agents, can affect the tan d response in the
glass transition region.41 The tan d peak height decreased in the order E-SBR/BR/ONAF
> E-SBR/BR/OEST > E-SBR/BR/OEPX, which corresponded the density of elastomer
crosslinking, as previously reported.42,43 The observation of the lower hysteresis for E-
SBR/BR/OEPX could be a result of a higher crosslink density, causing the storage
modulus (E0 ) to be predominant over the loss modulus (E00 ).
As tire tread requirements, the wet grip (associated with car driving safety) and rolling
resistance (associated with fuel consumption) were inferred from the tan d values at 0
and 60� C, respectively. This correlation is derived from the temperature–frequency
superposition, taking into account the high frequency nature of tire sliding (wet grip,
104–106 Hz) and low frequency nature of tire rolling (rolling resistance, 102–104
Hz).41,44 A rubber compound with a higher tan d value at 0� C and a lower tan d value at
60� C is expected to exhibited better tire tread performance.5 The tan d value at 0� C was
not affected by the use of OEST instead of ONAF in the rubber compound, whereas this
value increased by 13% when the epoxidized oil (OEPX) was used in the compound
instead of the naphthenic oil (reference). Thus, both vegetal oils can replace the naph-
thenic oil without change (OEST) or with improving (OEPX) of the wet grip property.
Therefore, tire treads fabricated with E-SBR/BR/OEPX may provide better wet grip
performance than those fabricated with E-SBR/BR/OEST or E-SBR/BR/ONAF. The tan
d value at 60� C increased for the rubber compound with OEST (13%) compared with
that of E-SBR/BR/ONAF but was the same for that with OEPX. Thus, tire treads pro-
duced using E-SBR/BR/OEST is expected to provide a higher rolling resistance than
those produced using the other compounds. Consequently, the fuel consumption of a car
with this tire tread would be higher. However, this increase in fuel consumption could be
compensated for by a longer tire tread lifetime, as E-SBR/BR/OEST also exhibited
higher abrasive wear resistance and a lower Tg.
The radar graphic shown in Figure 6 allows a clear comparison among the key
properties of the silica-filled E-SBR/BR/PLAST compounds. In this graphic, the ref-
erence compound E-SBR/BR/ONAF represents 1.0 and enhanced properties are located
closer to the edge. Thus, it is possible to replace ONAF as a plasticizer with either
modified soybean oil while maintaining a good balance among most of the properties.
E-SBR/BR/OEST and E-SBR/BR/OEPX showed enhanced tensile strength, elongation
at break, tear strength, and abrasive wear resistance, whereas the modulus at 100% value
Lovison
16 et al. 821
Journal of Elastomers & Plastics XX(X)

Figure 6. Radar graphic for the main properties of the E-SBR/BR/PLAST compounds

was the same for the former and slightly lower for latter compared with that for E-SBR/
BR/ONAF. In terms of the dynamic properties, E-SBR/BR/OEPX exhibited the best
performance, with a higher tan d value at 0 C and the same value as the reference
compound at 60 C. In contrast, E-SBR/BR/OEST had the same tan d value as the ref-
erence compound at 0 C and a higher value at 60 C. Advantageously, the vulcanization
time was also lower for rubber compounds with the modified soybean oils as plasticizers.

Conclusions
The replacement of ONAF by chemically modified soybean oils (OEST and OEPX)
affected several aspects of the silica-filled E-SBR/BR/PLAST compounds. Both the
modified soybean oils acted as rubber vulcanization activators, leading to faster curing and
higher crosslink densities. All the compounds exhibited good silica dispersion, despite the
occurrence of some silica agglomeration in the compounds with modified soybean oils,
which did not influence the overall compound properties. A good balance of mechanical
properties was observed, with a notable increase in abrasive wear resistance obtained for
the compounds with modified soybean oils. The use of OEPX led to a rubber compound
with better wet grip but a similar rolling resistance to the compound with ONAF, indi-
cating that OEPX is a good option for replacing ONAF. In contrast, the rolling resistance
was higher for the compound with OEST than for that with ONAF. Despite this increase in
rolling resistance, E-SBR/BR/OEST showed lower abrasive wear, which could provide a
longer tire tread lifetime and therefore compensate for the increased fuel consumption;
moreover, this compound has the potential for winter tire production owing to its lower Tg.
822
Lovison et al. Journal of Elastomers & Plastics 53(7)
17

Thus, both modified soybean oils can completely replace ONAF in silica-filled E-SBR/BR
rubber compounds and are interesting options for use as sustainable plasticizers in green
tire treads for passenger cars. According to our study, it is predicted that silica-filled E-
SBR/BR rubber plasticized solely with modified soybean oils can not only meet the
requirements of process sustainability but also decrease the time required for the manu-
facture of tire treads owing to the higher cure rates of these compounds. Bio-oil-based
plasticizers have gained considerable attention in the rubber industry, and modified soy-
bean oil will become commercially more competitive than petrochemical oils as con-
sumption and production increase. The effects of modified soybean oil on the crosslink
density of elastomer mixtures (especially E-SBR/BR) with silica still require further study
based on rheological measurements (elastic modulus) using a rubber processing analyzer
and 1H-NMR relaxation measurements.

Acknowledgements
The authors acknowledge SENAI-RS for supporting the development of this research and Unique
Rubber Technologies for providing chemical ingredients. The authors are also grateful to Mr
Jordão Gheller Jr (PhD) and Mrs Karin Brito (MSci) for useful discussions. We would like to
thank Editage (www.editage.com) for English language editing.

Declaration of conflicting interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship,
and/or publication of this article.

Funding
The author(s) received no financial support for the research, authorship, and/or publication of this
article.

ORCID iD
Maria Madalena de Camargo Forte https://orcid.org/0000-0001-6677-0698

References
1. Qin X, Wang J, Han B, et al. Novel design of eco-friendly super elastomer materials with
optimized hard segments micro-structure: toward next-generation high-performance tires.
Front Chem 2018; 6: 240.
2. Lee CK, Seo JG, Kim HJ, et al. Novel green composites from styrene butadiene rubber and
palm oil derivatives for high performance tires. J Appl Polym Sci 2019; 136: 47672.
3. Qin X, Han B, Lu J, et al. Rational design of advanced elastomer nanocomposites towards
extremely energy-saving tires based on macromolecular assembly strategy. Nano Energy
2018; 48: 180–188.
4. Kim MC, Adhikari J, Kim JK, et al. Preparation of novel bio-elastomers with enhanced
interaction with silica filler for low resistance and improved wet grip. J Clean Prod 2019; 208:
1622–1630.
5. Veiga VDA, Rossignol TM, Crespo JS, et al. Tire tread compounds with reduced rolling
resistance and improved wet grip. J Appl Polym Sci 2017; 134: 45334.
Lovison
18 et al. 823
Journal of Elastomers & Plastics XX(X)

6. Srivastava VK, Basak GC, Maiti M, et al. Synthesis and utilization of epoxidized poly-
butadiene rubber as an alternate compatibilizer in green-tire composites. Int J Ind Chem 2017;
8: 411–424.
7. Dasgupta S, Agrawal SL, Bandyopadhyay S, et al. Eco-friendly processing oils: a new tool to
achieve the improved mileage in tyre tread. Polym Test 2009; 28: 251–263.
8. Bocqué M, Voirin C, Lapinte V, et al. Petro-based and bio-based plasticizers: chemical
structures to plasticizing properties. J Polym Sci Part A: Polym Chem 2016; 54: 11–33.
9. Dasgupta S, Agrawal SL, Bandyopadhyay S, et al. Characterization of eco-friendly processing
aids for rubber compound. Polym Test 2007; 26: 489–500.
10. Li J, Isayev AI, Ren X, et al. Toward replacement of petroleum oils by modified soybean oils
in elastomers. Rubber Chem Technol 2016; 89: 608–630.
11. Santos JCO, Santos IMG, Conceição MM, et al. Thermoanalytical, kinetic and rheological
parameters of commercial edible vegetable oils. J Therm Anal Calorim 2004; 75: 419–428.
12. Meier MAR, Metzger JO and Schubert US. Plant oil renewable resources as green alterna-
tives in polymer science. Chem Soc Rev 2007; 36: 1788–1802.
13. Moutinho MTM, Dos Santos MR, and Hardy D. Oil extended functionalized
styrene-butadiene copolymer. Patent EP2986647B1, Germany, 2014.
14. Bastioli C, Capuzzi L, Magistrali P, et al. Vegetable oil derivatives as extender oils for
elastomer compositions. Patent EP2963084B1, Italy, 2011.
15. Papakonstantopoulos GJ, Hahn BR, and Rodewald S. Styrene/butadiene rubber extended with
low unsaturated soybean oil and tire with component. Patent US2018/0148567A1, USA, 2018.
16. Sandstrom PH, Rodewald S, and Ramanathan A. Tire with rubber tread containing combi-
nation of resin and vegetable oil, particularly soybean oil. Patent US2014/0135437A1, USA,
2014.
17. Hattori T, Sakaki T, Ichikawa N, et al. Rubber composition and pneumatic tire using the same.
Patent US2007/0123636A1, USA, 2007.
18. Chandrasekara G, Mahanama MK, Edirisinghe DG, et al. Epoxidized vegetable oils as pro-
cessing aids and activators in carbon-black filled natural rubber compounds. J Natl Sci Found
Sri Lanka 2011; 39: 243–250.
19. Jayewardhana WGD, Perera GM, Edirisinghe DG, et al. Study on natural oils as alternative
processing aids and activators in carbon black filled natural rubber. J Natl Sci Found Sri Lanka
2009; 37: 187–193.
20. Kukreja TR, Chauhan RC, Choe S, et al. Effect of the doses and nature of vegetable oil on
carbon black/rubber interactions: studies on castor oil and other vegetable oils. J Appl Polym
Sci 2003; 87: 1574–1578.
21. Bezerra A, Santos ACS, Da Costa HM, et al. Effect of linseed oil and peanut oil upon natural
rubber (NR) vulcanization. Part II: detailed model. Polı́m Ciênc Tecnol 2013; 23: 493–500.
22. Da Costa HM, Ramos VD, Campbell BC, et al. Thermal analysis of the sulfur vulcanization:
Part III. The role of linseed oil and peanut oil as activators. J Therm Anal Calorim 2017; 129:
755–766.
23. Zanchet A, Garcia PS, Nunes RCR, et al. Sustainable natural rubber compounds: naphthenic
oil exchange for another alternative from renewable source. Int Ref J Eng Sci 2016; 5: 10–19.
24. Espósito LH and Marzocca AJ.Silica-filled S-SBR with epoxidized soybean oil: influence of
the mixing process on rheological and mechanical properties of the compound. J Appl Polym
Sci 2020; 137: 48504.
25. European Union: Directive 2005/69/EC of the European Parliament and of the Council of 16
November 2005. Official Journal of the European Union (9.12.2005) L323, Vol. 18, 2005, p. 51.
824
Lovison et al. Journal of Elastomers & Plastics 53(7)
19

26. https://www.statista.com/statistics/263933/production-of-vegetable-oils-worldwide-since-
2000/ (accessed 7 December 2020).
27. Dierkes WK. Economic mixing of silica-rubber compounds: interaction between the chemistry
of the silica-silane reaction and the physics of mixing. PhD Thesis, University of Twente,
Netherlands, 2005.
28. Gheller J Jr, Ellwanger MV, and Oliveira V. Polymer–filler interactions in a tire compound
reinforced with silica. J Elastom Plast 2015; 48: 217–226.
29. Ciullo PA and Hewitt N. The rubber formulary. 1st edn. Norwich: William Andrew Pub-
lishing, 1999, p. 764.
30. Jebrane M, Cai S, Sandström C, et al. The reactivity of linseed and soybean oil with different
epoxidation degree towards vinyl acetate and impact of the resulting copolymer on the wood
durability. eXPRESS Polym Lett 2017; 11: 383–395.
31. Bragaglia M, Lamastra FR, Cherubini V, et al. 3D printing of polybutadiene rubber cured
by photo-induced thiol-ene chemistry: a proof of concept. eXPRESS Polym Lett 2020; 14:
576–582.
32. Kim K, Seo B, Lee JY, et al. Reduced filler flocculation in the silica-filled styrene–butadiene–
glycidyl methacrylate terpolymer. Compos Interfaces 2015; 22: 137–149.
33. De Souza FDB and Scuracchio CH. The use of atomic force microscopy as an important
technique to analyze the dispersion of nanometric fillers and morphology in nanocomposites
and polymer blends based on elastomers. Polı́m Ciênc Tecnol 2014; 24: 661–672.
34. Jeon IH, Kim H and Kim SG. Characterization of rubber micro-morphology by atomic force
microscopy (AFM). Rubber Chem Technol 2003; 76: 1–11.
35. Vanlandingham MR, McKnight SH, Palmese GR, et al. Nanoscale indentation of polymer
systems using the atomic force microscope. J Adhes 1997; 64: 31–59.
36. Perez LD and Lopez BL. Thermal characterization of SBR/NBR blends reinforced with a
mesoporous silica. J Appl Polym Sci 2012; 125: E327–E333.
37. Robertson CG, Lin CJ, Bogoslovov RB, et al. Flocculation, reinforcement, and glass transition
effects in silica-filled styrene-butadiene rubber. Rubber Chem Technol 2011; 84: 507–519.
38. George KM, Varkey JK, George B, et al. Physical and dynamic mechanical properties of silica
filled nitrile rubber modified with epoxidised natural rubber. KGK Kaut Gummi Kunst 2006;
59: 544–549.
39. Fox TG Jr and Flory PJ.Second-order transition temperatures and related properties of
polystyrene. I. Influence of molecular weight. J Appl Phys 1950; 21: 581–591.
40. Williams ML, Landel RF and Ferry JD. The temperature dependence of relaxation mechanisms
in amorphous polymers and other glass-forming liquids. J Am Chem Soc 1955; 77: 3701–3707.
41. Warasitthinon N and Robertson CG. Interpretation of the tand peak height for particle-filled
rubber and polymer nanocomposites with relevance to tire tread performance balance. Rubber
Chem Technol 2018; 91: 577–594.
42. Bandzierz K, Reuvekamp L, Dryzek J, et al. Influence of network structure on glass transition
temperature of elastomers. Materials 2016; 9: 607.
43. Park J, Eslick J, Ye Q, et al. The influence of chemical structure on the properties in
methacrylate-based dentin adhesives. Dent Mater 2011; 27: 1086–1093.
44. Akutagawa K. Technology for reducing tire rolling resistance. Tribol Online 2017; 12:
99–102.

You might also like