You are on page 1of 17

Tribology Letters (2018) 66:76

https://doi.org/10.1007/s11249-018-1027-9

ORIGINAL PAPER

Biobased Polyalphaolefin Base Oil: Chemical, Physical, and Tribological


Properties
Girma Biresaw1

Received: 1 February 2018 / Accepted: 7 May 2018 / Published online: 14 May 2018
© This is a U.S. Government work and not under copyright protection in the US; foreign copyright protection may apply 2018

Abstract
The properties of a biobased polyalphaolefin with a viscosity of 40 cSt at 100 °C (BPAO-40) were investigated relative to a
commercial petroleum-based PAO of similar viscosity at 100 °C (PAO-40). BPAO-40 was synthesized by oligomerization
of a mixture of alpha olefins, with and without terminal methyl esters. These olefins were obtained from vegetable oils via a
biorefinery process. In contrast to BPAO-40, commercial PAO-40 is synthesized only from non-functionalized alpha olefins.
Thus, BPAO-40 is not only biobased, but also has a unique chemical structure, which makes it a functionalized PAO. The
effect of chemical structure (presence or lack of methyl ester functionalization) on chemical, physical, and tribological prop-
erties of these two base oils was investigated. The investigation showed that, relative to the commercial non-functionalized
PAO-40, the functionalized BPAO-40 displayed the following properties: higher density at 40–100 °C, lower number aver-
age molecular weight, higher polydispersity index, higher viscosity index, lower oxidation stability (pressurized differential
scanning calorimetry), higher total acid number, higher free fatty acid, lower four-ball anti-wear coefficient of friction (COF)
and lower wear scar diameter (WSD), higher elastohydrodynamic (EHD) lubricant film thickness under boundary conditions
(low speeds and high temperature), lower EHD traction coefficient at 40 and 100 °C, similar pressure–viscosity coefficient,
lower COF, lower WSD, and higher relative film thickness on a high-frequency reciprocating rig tribometer under boundary
conditions (low speeds).

Keywords Biobased polyalphaolefin (BPAO) · Biorefinery · Coefficient of friction (COF) · Elastohydrodynamic (EHD)
film thickness · EHD traction coefficient · Free fatty acid (FFA) · Functionalized PAO · Number average molecular weight
(Mn) · Oligomerization · Oxidation stability · Polydispersity index (PDI) · Pressure–viscosity coefficient (PVC) · Total acid
number (TAN) · Viscosity · Viscosity index (VI) · Wear scar diameter (WSD)

1 Introduction from each category could be present in the formulation.


Also, in both cases, the component that is present in a larger
Lubricants and greases comprise two categories of compo- portion in the formulations (60–98%) is always the base oils
nents in their formulations, namely, base oils and additives [1–3]. Thus, most efforts at developing biobased lubricant
[1, 2]. In both lubricants and greases, one or more materials and grease formulations are primarily focused on replac-
ing petroleum base oils with biobased base oil. The suc-
cess of such substitution will provide lubricant formulations
Mention of trade names or commercial products in this publication with high biocontent along with the many benefits of using
is solely for the purpose of providing specific information and
renewable, biodegradable, and environmentally friendly
does not imply recommendation or endorsement by the U.S.
Department of Agriculture. USDA is an equal opportunity ingredients [4].
provider and employer. According to the American Petroleum Institute, base oils
used in lubricant formulation are divided into five groups,
* Girma Biresaw Groups I–V [5]. Groups I–III are petroleum distillates of
girma.biresaw@ars.usda.gov
specified sulfur content and/or viscosity index (VI). Group
1
Bio‑Oils Research Unit, National Center for Agricultural IV base oils are polyalphaolefins (PAOs). Group V base
Utilization Research, Agricultural Research Service, United oils are those that do not belong to the first four groups and
States Department of Agriculture, 1815 N. University Street, include vegetable oils, esters, silicones, phosphate esters.
Peoria, IL 61604, USA

13
Vol.:(0123456789)

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


76 Page 2 of 16 Tribology Letters (2018) 66:76

Earlier work on biobased lubricant development explored This product is commercialized under the trade name
using commodity vegetable oils (e.g., soybean oil, canola NovaSpec [25]. Petroleum distillate base oils (Groups I–III
oil) to replace the petroleum-based Groups I–IV base oils and naphthenics) are very cost competitive and account for
used in commercial lubricant formulations. The success of 99% of the lubricant base oil market share [3]. They are
this effort required overcoming some inherent weaknesses also very highly compatible with the farnesene-based Group
of vegetable oils (e.g., poor oxidation stability and cold flow III base oils (due to their structural similarity). Blending
properties) that impeded its application in lubricant formula- of petroleum distillate base oils with farnesene-based base
tion. Various approaches were successfully used in this effort oils (with renewable carbon) will provide a less expensive
including the use of appropriate additives, partially blend- route for producing biobased lubricant formulations with the
ing PAOs, partially blending high oleic vegetable oils which required biocontent.
have superior oxidation stability to commodity vegetable In the work describe here, we discuss a biobased base
oils, and various combinations of these methods [6–9]. oil that belongs to Group IV, i.e., PAO, produced by the
One of the drawbacks of vegetable oils is that, with some “Elevance” process [26–28]. In this process, a biorefinery
minor exceptions, they are available only in a very narrow is used to convert vegetable oils into functionalized (methyl
viscosity range of 30–40 cSt at 40 °C [10], which limits the ester) and non-functionalized alpha olefins with renewable
viscosity range of biobased lubricants that can be formulated carbons. The mixture of the functionalized and non-func-
with it. This problem was solved with the use of heat-bodied tionalized alpha olefins are then oligomerized to the desired
oils [11, 12]. These are biobased base oils produced by con- viscosity using the appropriate reaction conditions (catalyst,
trolled heating (time and temperature) of vegetable oil under temperature, etc). The resulting product is a biobased PAO
nitrogen. Heat-bodied oils of viscosity 75–1330 cSt at 40 °C (BPAO), partially functionalized due to the presence of ter-
have been produced from commodity soybean oil by vary- minal methyl ester groups on some of its branches. One such
ing the heating protocol and have been successfully used to product, BPAO-40 (commercially known as Aria WTP-40),
formulate high viscosity biobased lubricants, including gear with a viscosity of 40 cSt at 100 °C, was investigated for its
oils [13–15]. chemical, physical, and tribological properties relative to a
Vegetable oils and their derivatives (fatty acids, esters, commercial PAO of similar viscosity at 100 °C (PAO-40).
epoxides, etc.) have been used to synthesize a variety of The details of this investigation are given next.
biobased base oils [5]. Most of these products are at very
early stages of development and far from commercializa-
tion. The exception is estolides, which are synthesized from 2 Experimental
vegetable oils and/or fatty acids [16–18]. Estolides possess
excellent cold flow properties and have been used to formu- 2.1 Materials
late and test a variety of lubricants. Engine oil lubricants
formulated with estolides were reported to perform as well Samples of the functionalized BPAO-40, known under the
as or better than commercial petroleum-based products [19]. trade name Aria WTP-40, and non-functionalized petro-
Efforts at commercializing estolides under the trademark leum-based PAO (PAO-40), commercially marketed under
Biosynthetic™ base oil are underway [20]. the trade name SpectraSyn™ 40 by ExxonMobil (Hou-
Biobased base oils with branched hydrocarbons struc- ston, TX), were supplied by Elevance Renewables, LLC
tures, similar to petroleum distillates (Groups I–III), have (Woodridge, IL), and used as supplied. All other chemicals
been successfully produced using two different routes. The (solvents and reagents) were also used as supplied.
first route involves conversion of vegetable oil, either by the
“Total” process (Total S. A., France: catalytic deoxygena- 2.2 Nuclear Magnetic Resonance (NMR)
tion and isomerization) [21], or by the “Advonex” process
1
(hydrolysis, Kolbe electrolysis, separation, hydroisomeri- H NMR spectra of samples in C­ DCl3 solvent were obtained
zation) [22, 23]. Currently, both of these processes are at on a Bruker Avance 500 NMR spectrometer (Bruker Corpo-
pre-commercial stages of development. The second route ration, Billerica, MA) operating at 500.11 MHz and using
involves the production of farnesene by fermentation of a 5 mm BBO probe. Chemical shifts are reported in parts
sugar, which is then converted to Group III base oil with per million (ppm) from tetramethylsilane calculated from
renewable carbon [24]. The term “renewable carbon” is used the lock signal.
here to indicate that the sugars fermented for the produc-
tion of these farnesene-based Group III base oils are derived 2.3 Gel Permeation Chromatography (GPC)
from renewable agricultural crops such as corn, rice, potato,
sugarcane. On the other hand, regular Group III base oils are GPC profiles were obtained on a HPLC system that includes
distillates of crude petroleum with non-renewable carbon. a 1515 isocratic HPLC pump, 717 plus automated injector,

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Tribology Letters (2018) 66:76 Page 3 of 16 76

column heater, and Breeze software (Waters Corporation, min. The instrument was fitted with a computer and appro-
Milford, MA). Columns used were PLgel 3 µm MIXED- priate software to allow for data acquisition and analysis.
E, 300 × 7.5 mm and a PLgel 5 µm Guard, 50 × 7.5 mm, Duplicate runs were conducted, and average values of onset
part numbers PL1110-6300 and PL1110-1520, respectively (OT) and peak (PT) oxidation temperatures are reported.
(Polymer Laboratories, Varian, Inc., Amherst, MA). Sig- A detailed description of the instrument and the procedure
nals generated from an Optilab rEX refractive index detector have been given before [33].
were processed using ASTRA V macromolecular characteri-
zation software (Wyatt Technology Corporation, Santa Bar- 2.7 Four‑Ball (4‑Ball) Anti‑wear (AW)
bara, CA). THF was used as the mobile phase at a flow rate
of 1 mL/min and columns were maintained at 40 °C. The 4-Ball AW tests were conducted according to ASTM D4172-
liquid phase samples were brought into solution with THF 94 [34] on a model KTR-30L 4-ball tribometer equipped
stabilized with butylated hydroxytoluene (Fisher Scientific, with TriboDATA software (Koehler Instrument Company,
Suwanee, GA) at a known concentration near 0.004 g/mL. Bohemia, NY). Details about the instrument (hardware and
A Waters autosampler was used to make 100 µL injections software) and test procedure has been given previously [35].
from a 1 mL sample vial. Linear polystyrene standards of Coefficient of friction (COF) from each test was calculated
Mn = 580–100,000 Da, Mw/Mn = 1 (Polymer Laboratories, from the corresponding torque and load data in accordance
Santa Clara, CA), were used for the calibration of molecular with ASTM D5183 [36]. Wear scar diameters (WSDs), along
weights of all polymers. Chromatogram data exported using and across the wear direction of the three bottom balls, were
Astra V software were used to calculate molecular weight measured at the end of each test using a wear scar measure-
using Microsoft Excel (Redmond, WA) software. ment system comprising hardware and ScarView software
(Koehler Instrument Company, Inc., Bohemia, NY). Each
2.4 Total Acid Number (TAN) test lubricant was used in at least two measurements, and
average COF and WSD are reported.
TAN was determined according to the official AOCS Method
Te 2a-64 [29], on the 751 GPD Titrino (Metrohm Ltd., Heri- 2.8 High‑Frequency Reciprocating Rig (HFRR)
sau, Switzerland), with ethanol substituted for methanol to Tribotest
increase solubility of the samples during titration. In general,
1.0 g of sample was dissolved in 80 mL of ethanol:water Tests were conducted on a model D519 HFRR tribometer
(9:1) under vigorous stirring with a titration dose rate of (PCS Instruments, London, England) between a reciprocat-
0.05 mL/min. All TAN experiments were conducted in trip- ing steel ball sliding against a stationary steel disc, under
licate, and average and standard deviation values of TAN the set test conditions of applied load, lubricant tempera-
and free fatty acid (FFA) are reported. ture, stroke length, and frequency. All tests were conducted
in an enclosed chamber maintained at ambient temperature
2.5 Density, Viscosity, and Viscosity Index (VI) and 60% relative humidity. The instrument measures and
displays the following data as a function of time: lubricant
Density and dynamic viscosity were measured on a Stabin- temperature, COF, and relative film thickness (%). At the
ger SVM3000/G2 viscometer (Anton Paar GmbH, Graz, end of each test, wear scars (along and perpendicular to the
Austria) according to ASTM D 7042 [30]. The data were direction of the wear tracks) on the ball and disc are meas-
then used to calculate kinematic viscosity at the correspond- ured on a metallurgical microscope (model N334PCS, PCS
ing temperatures using the procedures described in the Instruments, London, England), fitted with an adjustable
method. VI was estimated from kinematic viscosity values stage micrometer designed to fit the ball holder. Average
at 40 and 100 °C following the procedure outlined in ASTM WSD values for the ball are reported; the width and length
D 2270-93 [31]. of the wear on the disc are measured and reported sepa-
rately. Details about the various components of the D519,
2.6 Pressurized Differential Scanning Calorimetry its specifications, test procedures, and the specification of
(PDSC) the steel ball and disc test specimen used in the test, can be
found elsewhere [37].
PDSC tests were conducted on a Q20P pressurized differ-
ential scanning calorimeter (TA Instruments—Waters LLC, 2.9 Elastohydrodynamic (EHD) Film Thickness
New Castle, DE) following the procedure outlined in ASTM and Traction
D6186 [32]. All tests were conducted with the cell pressur-
ized with pure oxygen to 500 ± 25 psig in dynamic mode, Measurements were conducted using the film thickness
i.e., with positive oxygen flowing at a rate of 100 ± 10 mL/ or traction configuration of the EHL Ultra Thin Film

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


76 Page 4 of 16 Tribology Letters (2018) 66:76

Measurement System (PCS Instruments, London, England). 1-octene, 1-dodecene) as well as mixtures of linear alpha
Film thickness was measured between a glass disk and steel olefins are also used. Conventionally, the linear alpha ole-
ball by the method of optical interferometry. In this configu- fins, catalyst, and oligomerization conditions are selected
ration, the speed of the disk (ud) is controlled while the ball, to produce PAO products of a specific kinematic viscosity
which is in intimate contact with the disk, rotates freely at at 100 °C, after a hydrogenation step, which is necessary
the same speed as the disk (ud = ub), and film thickness is to remove any residual unsaturation in the product. Thus,
measured under pure rolling conditions. Traction is meas- for example, PAO-6 and PAO-40 products will have kin-
ured between independently driven steel disk and steel ball. ematic viscosities of around 6 and 40 cSt, respectively.
This allows for independently controlling the speeds of the The BPAOs differ from the PAOs in that they are syn-
disk and ball to obtain the desired entrainment speed (u) and thesized from a mixture of functionalized and non-func-
slide-to-roll ratio (SRR), defined as follows: tionalized linear alpha olefins [27]. The non-functionalized
( ) linear alpha olefin components are similar to those used
u (m∕s) = ub + ud ∕2, (1) in the PAO synthesis discussed above, and could be of
( ) ( ) ( ) petroleum or renewable non-petroleum origin. The func-
SRR (%) = 100 ||ub − ud ||∕u = 200||ub − ud || ∕ ub + ud .
tionalized linear alpha olefins comprise a functional group
(2) (e.g., methyl ester) at the other end of the chain, and could
Under pure rolling conditions, the disk and ball have be of petroleum or renewable origin.
identical speeds resulting in 0% SRR. The BPAO investigated in this work was synthesized
Description of the instrument and its specifications, as from a mixture of functionalized and non-functionalized
well as the specifications of the glass disk, steel disk, and linear alpha olefins, both of which were obtained from
steel balls used in film thickness and traction measurements, renewable feedstock, namely, vegetable oils [27]. A sche-
has been given before [38, 39]. The application of interfer- matic depicting the biorefinery process for converting veg-
ometry for measuring film thickness is described elsewhere etable oils into alpha olefins and other products is given
[40]. Traction measurement on the Ultra Thin Film Measure- in Fig. 1 [27, 28]. The first step in this process involves
ment System has been discussed before [39]. Film thickness a patented metathesis reaction of the vegetable oils with
and traction measurement procedures, including instrument ethylene as the co-reactor. The product from this step is
set up/calibration and data acquisition/analysis, are given distilled to separate the 1-decene and other non-function-
elsewhere [38, 39]. alized linear alpha olefin. The residue from the distillation
is then subjected to a transesterification step to produce
9-decenoic acid methyl ester and other functionalized lin-
3 Results and Discussion ear alpha olefins.
The functionalized and non-functionalized linear alpha
3.1 Synthesis and Structure olefins are then used in the oligomerization reactions in
the presence of the appropriate catalyst to produce the
Petroleum-based PAOs are commercially manufactured BPAOs. A schematic comparing the synthesis of BPAOs
by the oligomerization of linear alpha olefins in the pres- and PAOs and the corresponding base oil structures is
ence of one or more catalysts [41]. Most PAOs are synthe- given in Fig. 2.
sized from 1-decene, but other linear alpha olefins (e.g.,

Fig. 1  Elevance biorefinery


process for synthesis of func-
tionalized and non-functional-
ized linear alpha olefins from
vegetable oils [27]

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Tribology Letters (2018) 66:76 Page 5 of 16 76

Fig. 2  Schematics for the


synthesis of biobased (BPAO-
40) and commercial petroleum-
based (PAO-40) polyalphaole-
fins

Table 1  Density of biobased (BPAO-40) and commercial petroleum- Table 2  Viscosity (ASTM D7042 [30]) and viscosity index (ASTM
based (PAO-40) polyalphaolefins investigated in this work D-2270 31]) of biobased (BPAO-40) and commercial petroleum-
based (PAO-40) polyalphaolefins
T (°C) Density (g/mL)a
⟨kVis⟩ ± stdev ­(mm2/s) ⟨dVis⟩ ± stdev (mPa s)
BPAO-40 PAO-40
BPAO-40 PAO-40 BPAO-40 PAO-40
40 0.8697 ± 0.0006 0.8333 ± 0.0002
75 0.8487 ± 0.0005 0.8132 ± 0.0002 40 °C 324.6 ± 3.4 397.6 ± 0.5 282.3 ± 2.8 331.3 ± 0.4
100 0.8338 ± 0.0003 0.7988 ± 0.0002 75 °C 77.8 ± 0.6 86.4 ± 0.1 66.1 ± 0.5 70.2 ± 0.0
100 °C 37.3 ± 0.3 39.3 ± 0.0 31.1 ± 0.2 31.4 ± 0.0
All data from this work unless noted VI 164 149
a
ASTM D-7402 [30]
All data from this work unless noted

3.2 Physical and Chemical Properties


similar kinematic viscosity at 100 °C. However, as shown in
In discussions henceforth, the functionalized biobased and Table 2, the change in viscosity with change in temperature
the non-functionalized petroleum-based PAOs will be des- from 40 to 100 °C was steeper for PAO-40 than for BPAO-
ignated as BPAO-40 and PAO-40, respectively. 40. The consequence of this is a lower VI for PAO-40 com-
Table 1 compares the density of these two materials at 40, pared to BPAO-40 (Table 2).
75, and 100 °C measured in accordance with ASTM D-7042 The two oils were evaluated for their oxidation stability
[30]. The data show the BPAO-40 oil to be denser at all using a PDSC method as described in ASTM D-6186 [32].
temperatures. This can be attributed to the presence of some Measured onset (OT) and peak (PT) oxidation temperatures
oxygen atoms in the BPAO-40, whereas the PAO-40 is all for the two oils are compared in Table 3. The data given
hydrocarbon. The presence of oxygen can result in increased in Table 3 indicate a slightly higher OT and PT oxidation
density due to its slightly higher molecular weight than car- stability for the commercial and non-functionalized PAO-
bon. In addition, the presence of oxygen in the form of ester 40 relative to the functionalized BPAO-40. This difference
functional groups can produce dipolar interactions that could might be due to the presence of traces of unsaturation in
result in lower molecular volume and, hence, higher density. the BPAO-40 sample due to incomplete hydrogenation of
The viscosity indices of the functionalized BPAO-40 and residual double bonds after the oligomerization process.
PAO-40 polyalphaolefins at 40, 75, and 100 °C were meas- This explanation is supported by the 1H NMR analysis data
ured according to ASTM D-7042 [30] and are compared of the samples, which showed absorbances related to unsatu-
in Table 2. The data show that PAO-40 has a much higher ration in the range 5.7–5.9 ppm for BPAO-40 but none for
viscosity than BPAO-40 at 40 °C, which correlates well with PAO-40 (Fig. 4).
the higher number average molecular weight value (⟨Mn⟩) of TAN and FFA analysis, using the AOCS official method
PAO-40 compared to BPAO-40, measured using GPC [42] [29], showed very low values (TAN < 0.3 mg KOH/g;
(Fig. 3). The GPC traces in Fig. 3 also show PAO-40 with FFA < 0.2%) for both oils (Table 4). Still, the values for the
narrower molecular weight distribution, i.e., a lower poly- BPAO-40 base oil were higher than those for the non-func-
dispersity index (PDI) than that for BPAO-40. The viscosity tionalized PAO-40 polyalphaolefin. This result might be an
data in Table 2 show that the gap in viscosity between the indication of potential hydrolysis of the methyl esters into
two oils narrows with increasing temperature, from about fatty acids in the BPAO-40 polyalphaolefin, possibly due to
73 cSt at 40 °C to about 2 cSt at 100 °C. This is consist- the presence of traces of moisture in the oil. The data given
ent with the fact that the two oils were synthesized to have in Table 4 also show that the TAN values measured in this

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


76 Page 6 of 16 Tribology Letters (2018) 66:76

Fig. 3  GPC trace for BPAO-40


and PAO-40
-6
15x10

10

Intensity
BPAO-40, <Mn>=1730 Da; PDI=2.0
PAO-40, <Mn>=1820 Da; PDI=1.3
5

0
4 5 6 7 8 9 2 3 4 5 6 7 8 9 2
3 4
10 10
Mol. wt. (g/mol)

Table 3  PDSC onset (OT) and peak (PT) oxidation temperatures of 3.3 Elastohydrodynamic (EHD) Properties
biobased (BPAO-40) and commercial petroleum-based (PAO-40)
polyalphaolefins
3.3.1 Film Thickness
BPAO-40 PAO-40

OT (°C) 185.8 ± 0.6 195.0 ± 0.8


Film thickness was measured between a steel ball and a
glass disk by the method of optical interferometry [40].
PT (°C) 205.1 ± 0.8 209.7 ± 0.9
Measurements were conducted under pure rolling condi-
All data from this work. ASTM D-6186 [32] tions, i.e., at a SRR of 0% defined as follows:
( ) ( ) ( )
SRR (%) = 100 ||ub − ud ||∕u = 200||ub − ud || ∕ ub + ud ,
work for the PAO-40 were higher than the values reported (2)
in the manufacturer’s Product Datasheet [43]. This might be where ub and ud are the speeds of the ball and disk, respec-
an indication of potential oxidation of the PAO-40 during tively, in m/s, u = (ub + ud)/2, is entrainment speed in m/s.
storage and/or transportation.

Fig. 4  Comparison of 1H NMR


absorptions in the region for
olefinic protons of BPAO-40
and PAO-40

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Tribology Letters (2018) 66:76 Page 7 of 16 76

Table 4  Total acid number and free fatty acid of biobased (BPAO-40) Figure 6 compares the measured film thickness of the
and commercial petroleum-based (PAO-40) polyalphaolefins two oils at 100 °C with that predicted for BPAO-40 by
BPAO-40 PAO-40 the Hamrock–Dowson equation. The two oils have almost
Average ± Stdev Average ± Stdev identical viscosity at 100 °C (Table 2), and, as a result, dis-
played almost identical film thickness at all entrainment
Total acid number 0.246 ± 0.007 0.154 ± 0.005 (lit.a < 0.10)
speeds above about 0.05 m/s. The measured data for the
(mg KOH/g)
two oils were also close to that predicted for BPAO-40 by
Free fatty acid (%) 0.124 ± 0.004 0.077 ± 0.002
the Hamrock–Dowson equation above 0.05 m/s. However,
All data from this work unless noted. Measured using AOCS Official below 0.05 m/s entrainment speed, which corresponds to
Method [29] boundary lubrication conditions, the functional PAO BPAO-
a
Data from [43] measured using ASTM D-974 method [54] 40 produced thicker films than that predicted by the Ham-
rock–Dowson equation (Fig. 5). On the other hand, in the
Under pure rolling conditions, the disk and ball have identi- same low entrainment speed regions (< 0.05 m/s), the non-
cal speeds resulting in 0% SRR. functional PAO-40 produced thinner lubricant films than
Film thickness measurements were conducted as a func- those predicted by the Hamrock–Dowson equation (Fig. 5).
tion of entrainment speed [u in Eq. (1)] and various com- This difference in the film forming properties of the func-
binations of temperatures (40, 75, and 100 °C) and loads tional (BPAO-40) and non-functional (PAO-40) polyal-
(10, 20, 30, 40 N). phaolefins in the boundary regime can be attributed to the
The film thickness of the functionalized BPAO-40, as a difference in their chemical structures. As mentioned earlier,
function of entrainment speed, is compared to those of the BPAO-40 comprises functional ester groups in its structures,
non-functionalized PAO-40 shown in Fig. 5. Data at 40 °C which allows it to adsorb onto friction surfaces and generate
and 10 or 40 N load, and 75 °C and 40 N load are com- thin boundary lubricant films. On the other hand, the non-
pared. The data for both oils show film thickness increas- functional PAO-40 has no such functional groups and cannot
ing with increasing entrainment at all temperatures and generate significant boundary films.
loads. At 40 °C, the film thickness values for BPAO-40
were much lower than those for PAO-40 at all entrainment 3.3.2 Traction
speeds. The difference in film thickness between the two
oils was larger at 10 N load compared to that at 40 N load Traction experiments were conducted between a steel
(Fig. 5a, b). At 75 °C and 40 N load, the two oils displayed disk and a steel ball at various combinations of tempera-
similar film thickness properties at all entrainment speeds ture (40 or 100 °C), load (10, 20, 30, 40 N), entrainment
except at very low speeds corresponding to boundary con- speed (0–3.0 m/s), and SRR (0–100%). The disk and ball
ditions (Fig. 5c). are equipped with independent drive motors to allow for
The above discussed observations of the data given precise setting of entrainment speeds and SRR. Two types
in Fig. 5 are consistent with the predictions of the Ham- of traction experiments were conducted. In the first type,
rock–Dowson equation which is given as follows [44]: traction coefficient was measured as a function of entrain-
ment speed (u) at a constant SRR of 50%. In the second type
h = k(u𝜂)0.67 𝛼 0.53 , (3) of experiments, the traction coefficient was measured as a
where h is the lubricant film thickness in nm, k is a constant function of SRR at a constant entrainment speed of 1.0 m/s.
whose value is a function of the film thickness measurement Figure 7 shows the traction coefficient of the functional-
instrument parameters and test geometry, u is entrainment ized (BPAO-40) and non-functionalized (PAO-40) polyal-
speed in m/s, η is dynamic viscosity in cP at the film thick- phaolefins, at 40 and 100 °C, as a function of entrainment
ness measurement temperature and ambient atmospheric speed and a constant SRR of 50%. At both temperatures,
pressure, and α is pressure–viscosity coefficient (PVC) at the traction coefficient data for both oils displayed simi-
the film thickness measurement temperature in ­GPa− 1. Thus, lar profiles comprising three distinct regions: region I
as predicted by the Hamrock–Dowson equation, the film with high traction coefficient values (> 0.015) that varied
thickness of both oils displayed a dramatic increase with slightly with varying entrainment speeds in the range of 0 to
increasing entrainment speed at all temperatures and loads. about 2.0 m/s, region III with low traction coefficient values
The higher film thickness of PAO-40 over BPAO-40 at 40 °C (< 0.003) that were more or less constant with increasing
shown in Fig. 5a is consistent with its higher viscosity at entrainment speed above about 2.7 m/s, region II, which is a
this temperature (Table 2). However, the viscosity differ- transition between regions I and III, with the traction coeffi-
ence between the oils narrows at 75 °C (Table 2), resulting cient decreasing sharply with increasing entrainment speed.
in a very similar film thickness for the two oils at almost all The two oils displayed identical traction coefficient values in
entrainment speeds (Fig. 5c). region III but very different values in region I.

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


76 Page 8 of 16 Tribology Letters (2018) 66:76

Fig. 5  EHD film thickness of


1000 (a)
BPAO-40 and PAO-40 at 40
8
and 75 °C 7
6
5

Film Thickness:nm
4

3
film thickness at 40°C,10N
PAO-40
2
BPAO-40

100
8
7
6
5
4
2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 2 3
0.01 0.1 1
Entrainment Speed (m/s)

1000 (b)
8
7
6
5
Film Thickness:nm

2 film thickness at 40°C, 40N


PAO-40
BPAO-40
100
8
7
6
5
4
2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 2 3
0.01 0.1 1
Entrainment Spees (m/s)

8
6
(c)
4

2
Film Thickness: nm

100
8 film thickness at 75°C, 40N
6 PAO-40
4 BPAO-40

10
8
6

2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 2 3 4
0.01 0.1 1
Entrainment Speed (m/s)

Examination of the data in region I in Fig. 7a, b shows of temperature on oil viscosity, which is one of the major
that traction coefficient is a function of both temperature and oil properties affecting traction coefficient. Higher tempera-
oil chemistry. Comparing the effect of temperature shows ture reduces viscosity which in turn reduces traction coef-
that both oils displayed lower traction coefficient at 100 °C ficient. However, at both temperatures, the functionalized
compared to that at 40 °C. This was attributed to the effect BPAO-40 displayed a lower traction coefficient than the

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Tribology Letters (2018) 66:76 Page 9 of 16 76

Fig. 6  Measured versus 4

predicted EHD film thickness 3


of BPAO-40 and PAO-40 at meas vs H-D pred film thickness at 100°C, 40N
100 °C 2 meas_h_BPAO-40 meas_h_PAO-40
0.67
pred_h_BPAO-40 =const (u)
100

Film Thickness (h, nm)


7
6
5
4

10
7
6
5
4
2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 2 3 4
0.01 0.1 1
Entrainment Speed (u, m/s)

Fig. 7  EHD traction coefficient


for BPAO-40 and PAO-40 as a -3
(a)
50x10
function of entrainment speed at Traction coeff: 40°C;40N;50% SRR
50% SRR and 40 N load PAO-40
40 BPAO-40
Traction Coefficient

30

20

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

Entrainment speed (m/s)

-3
30x10
(b)
25 traction coeff: 100°C;40N;50%SRR
PAO-40
BPAO-40
Traction Coefficient

20

15

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

Entrainment speed (m/s)

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


76 Page 10 of 16 Tribology Letters (2018) 66:76

non-functionalized PAO-40. This difference can be attrib- viscosity and, hence will have nearly equal µh contribution
uted to the difference in the chemical structure of the two to the measured value (µ). On the other hand, because
oils and perhaps, to a lesser extent, their viscosities. Thus, the functionalized PAO BPAO-40 can adsorb on friction
the BPAO-40 oil with functional group produced lower trac- surfaces and lower boundary friction, it will have a much
tion coefficients than the PAO-40 non-functionalized oil. lower µb contribution to the traction coefficient than the
Similar results have been reported before from the compari- non-functionalized PAO-40. This analysis is consistent
son of the traction coefficients of vegetable oils, which have with the observed significantly lower traction coefficient
multiple ester functional groups in their structures, relative of BPAO-40 than that of PAO-40 shown in region I of
to commercial PAOs of similar viscosity but without func- Fig. 7b.
tional groups [39]. Traction coefficient data for the two oils as a function of
The temperature and chemistry effects discussed above SRR at 1.0 m/s entrainment speed are given in Fig. 8. The
indicate that the traction coefficient in region I (µ) can be results display two common features with the entrainment
considered to be a composite value of contributions from speed versus traction coefficient data (Fig. 7) discussed
hydrodynamic (µh) and boundary (µb) traction coefficients above. First is that the traction coefficients are inversely
as shown below: related to temperature, where the values at 100 °C are
lower than those at 40 °C (compare Fig. 8a vs. b). Second,
𝜇 = 𝜇h + 𝜇b . (4) the traction coefficients of the functionalized BPAO-40
The hydrodynamic component is affected by the viscos- polyalphaolefin are lower than the values for the non-func-
ity of the oils, and, at 100 °C, the two oils have similar tionalized PAO-40 at all temperatures and SRR.

Fig. 8  EHD traction coefficient


of WTP-40 and PAO-40 as (a)
-3
40x10
a function of SRR, at 1 m/s
entrainment speed and 40 N
load
Traction Coefficient

30
traction coeff: 40°C;40N;1m/s
PAO-40
BPAO-40
20

10

0
0 20 40 60 80 100 120

Slide/Roll Ratio (%)

-3
35x10
(b)
30

25
Traction Coefficient

20
traction coef: 100°C, 40N;1m/s
15 PAO-40
BPAO-40
10

0
0 20 40 60 80 100 120

Slide/Roll ratio (%)

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Tribology Letters (2018) 66:76 Page 11 of 16 76

3.3.3 Pressure–Viscosity Coefficient (uη) values, i.e., at (uη)t = (uη)r. If these conditions are met,
Eq. (5) can be rearranged and simplified to calculate the
As shown in Eq. (2), the two lubricant properties that affect PVC of the test lubricant as follows:
its film thickness (h) are its viscosity (η) and its PVC (α).
( )[1∕0.53]
PVC values of lubricants vary in the range 5–60 GPa− 1 𝛼t = 𝛼r ht ∕hr . (6)
[45], and the higher the PVC of a lubricant the higher its
film thickness. The PVC of lubricants are estimated using a In applying Eq. (6) in this work, PAO-6, which is a PAO
variety of methods that can be grouped into direct and indi- sold under the trade name Durasyn 166 (INEOS Oligomers,
rect methods. The direct method which is also the preferred LLC), was used as the reference oil. PAO-6 has a viscosity
method involves the analysis of pressure vs. viscosity data of 6 cSt at 100 °C and a PVC (αr) of 15.8 GPa− 1 indepen-
on the lubricants of interest [46, 47]. However, this method dently determined by the direct method [53]. Film thickness
is very hard to apply for most researchers because of the data for the test (BPAO-40, PAO-40) and reference (PAO-6)
scarcity of access to instruments capable of generating the oils, measured on the same instrument at 40 °C and 20 N
pressure vs. viscosity data. The alternative indirect methods load, were used in the analysis. Figure 9 shows the film
involve analysis of either lubricant film thickness data [48, thickness data used in the analysis as a function of (uη). The
49] or physical property data [50–52]. points in Fig. 9 are experimental values and the lines are
In this work, the PVC values of the functionalized predicted using a simplified form of the Hamrock–Dowson
(BPAO-40) and the non-functionalized (PAO-40) polyal- equation [Eq. (3)] as follows:
phaolefins were estimated from film thickness data. Applica- h = a(u𝜂)0.67 , (7)
tion of the film thickness method of PVC estimation requires
where a is a constant.
combining Eq. (3) for the test (ht) and reference (hr) lubri-
Application of the film thickness data shown in Fig. 9
cants as follows:
for the test and reference lubricants along with the value of
( ) [ ( )0.53 ] [ ( )0.53 ] αr (15.8 GPa− 1) in Eq. (6) gave 8.9 and 11.1 GPa− 1, for the
ht ∕hr = k(u𝜂)0.67 𝛼t ∕ k(u𝜂)0.67 𝛼r . (5)
t r
PVC of the functionalized BPAO-40 and the non-functional-
In order to apply Eq. (5) to estimate the PVC of a new or ized PAO-40 polyalphaolefins, respectively. These values are
test lubricant requires (a) a reference lubricant whose PVC consistent with expectations based on the structures of these
(αr) was experimentally determined using a direct method, oils [45]. PAOs belong to intermediate PVC oils (PVC of
(b) the film thickness data used in the analysis for both the 10–20 GPa− 1), which include most Groups I–V oils. Inter-
test (ht) and reference (hr) samples have been measured on mediate PVC oils have distinct structural differences from
the same instrument under identical conditions of tempera- low PVC (< 10 GPa− 1) oils such as liquid crystals, which
ture, load, entrainment speeds, and (c) the film thickness have layered structures that shear and are unable to support
data of the test (ht) and reference (hr) samples to be ana- high loads. Intermediate PVC oils are also different from
lyzed must have been generated under identical range of high PVC (> 20 GPa− 1) oils such as traction oils, whose

Fig. 9  Measured (hm) and pre- 300


dicted (hp) EHD film thickness 0.67
meas vs pred [const(uη) ] film thickness
as a function of (uη) used for pao-6: h_m h_p
estimation of the pressure–vis- 250 pao-40: h_m h_p
cosity coefficient [αt in Eq. (6)] bpao-40: h_m h_p
of BPAO-40 and PAO-40 by
film thickness (h), nm

the Hamrock–Dowson method. 200


PAO-6 is the reference fluid
whose pressure–viscosity coeffi-
cient [αr in Eq. (6)] has already 150
been independently determined
to be 15.8 GPa− 1 [53]
100

50

0
0 10 20 30 40 50

uη, mN/m

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


76 Page 12 of 16 Tribology Letters (2018) 66:76

structure allows them to support high load due to entangle- polyalphaolefins is compared in Fig. 10a–c. Each point in
ments and/or strong intermolecular interactions [45]. Fig. 10a–c corresponds to the average and standard deviation
of multiple tests at each average speed, which is obtained by
3.4 Friction and Wear Properties using varying combinations of stroke length and frequency.
The three sets of data (COF, WSD, hr) shown in Fig. 10
3.4.1 High‑Frequency Reciprocating Rig (HFRR) show the following common features:
Evaluations
(a) At high average speeds (ua > 0.12 m/s), the two oils dis-
The functionalized BPAO-40 and non-functionalized PAO- played similar COF, WSD, and hr values that remained
40 polyalphaolefins were evaluated for their friction, wear, constant with increasing ua. This speed region corre-
and relative film thickness properties as a function of aver- sponds to the hydrodynamic regime, where lubricant
age speed, ua, in m/s, on a HFRR. Average speed on the properties are functions of viscosity. Since the two
HFRR is defined as follows: oils have similar viscosity values at this temperature
(Table 2), they are expected to display similar values
ua (m∕s) = (2s)𝜐, (8) of COF, WSD, and hr, as was observed.
where s is stroke length in m, and υ is frequency in Hz. (b) At very low average speeds (< 0.08 m/s for PAO-
Experiments were conducted on a wide range of ua values 40, < 0.03 m/s for BPAO-40), the two oils displayed
(0.4 × 10− 3–400 × 10− 3 m/s) using various combinations of different COF, WSD, and hr values that were almost
stroke length (s) and frequency (υ) settings available on the independent of ua. Since this speed region corresponds
instrument. All experiments were conducted at 1 kgf (9.8 N) to the boundary lubrication regime, the properties of
load, 75 °C for 1 h. the oils are a function of their chemical structures. In
COF and relative film thickness (hr, %) data were meas- this region, the functionalized BPAO-40 polyalphaole-
ured as a function of time, displayed, and saved in real time fin is expected to adsorb on friction surfaces and pro-
at a rate of 60/min. The real time data and corresponding duce lower COF, lower WSD, and higher hr than the
averages are saved and displayed at the end of each test. non-functionalized PAO-40 polyalphaolefin. The data
Also, at the end of each test, the WSD on the ball and disc shown in Fig. 10 are consistent with this expectation.
are measured manually using a specialized microscope, and (c) In between the above two regions of average speed, the
the wear scar values, along and across the wear direction, two oils displayed different COF, WSD, and hr values
are recorded and combined with the corresponding COF and that varied with increasing ua. In this region, both oils
film thickness data of the specific run. displayed relative film thickness values that increased
The relative film thickness data on the HFRR are obtained with increasing average speed, with the values for the
from electric contact resistance data between the steel disc BPAO-40 being consistently much higher than those
and steel ball measured during the test. According to this for the PAO-40. Similarly, COF and WSD values, espe-
method, measured relative film thickness values vary from cially for the PAO-40, sharply decreased with increas-
0 to 100%, depending on the value of the measured electric ing average speed.
contact resistance. In the absence of lubricant film, the ball (d) The COF and relative film thickness data from the
and disk are in direct contact, there will be zero electric con- HFRR measurements shown in Fig. 10 are consistent
tact resistance between the two, and a relative film thickness with film thickness and traction data discussed earlier
of 0% is measured. In the presence of thick non-conductive from EHL experiments. Thus, the increasing relative
lubricant film, the ball and disk are completely separated, film thickness with increasing average speed shown in
maximum electric contact resistance occurs, and a relative Fig. 10c correlates well with the entrainment speed (u)
film thickness of 100% is measured. Relative film thick- versus film thickness (h) data shown in Figs. 5 and 6.
ness values between these two extremes are obtained when Similarly, the decreasing COF with increasing aver-
partial contact between the ball and disk occurs, resulting age speed shown in Fig. 10a correlates well with the
in intermediate electric contact resistance values. As will decrease of traction coefficient with increasing entrain-
be shown below, the relative film thickness data obtained ment speed displayed in Fig. 7.
on the HFRR are capable of differentiating film thickness
between the major lubrication regimes (boundary, hydro- 3.4.2 Four‑Ball Anti‑wear Evaluations
dynamic, mixed), even though it is too crude to distinguish
film thickness within each regime. The 4-ball AW evaluations were conducted according to
The effect of average HFRR speed (ua) at 75 °C on COF, ASTM D-4172 and ASTM D5183 [34, 36] under the fol-
steel ball WSD, and relative film thickness (hr) of the func- lowing test conditions: 75 °C, 1200 rpm, 392 N load, and
tionalized (BPAO-40) and non-functionalized (PAO-40) for 1 h. COF data, collected at a rate of 60/min, are averaged

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Tribology Letters (2018) 66:76 Page 13 of 16 76

Fig. 10  HFRR coefficient of


friction, relative film thickness,
(a)
and wear scar diameter as a 0.4
function of average speed, at COF @ 75°C
75 °C and 1 kgf, for BPAO-40 PAO-40

<Coefficient of Friction>
and PAO-40 0.3 BPAO-40

0.2

0.1

0.0
0.0 0.1 0.2 0.3 0.4
average speed, m/s

800
(b)
WSD @ 75°C
600 PAO-40
<wear scar diameter>, µm

BPAO-40

400

200

0
0.0 0.1 0.2 0.3 0.4
average speed, m/s

(c)
100

80
<relative film thickness>, %

60 relative film thickness @ 75°C


BPAO-40
PAO-40
40

20

0.0 0.1 0.2 0.3 0.4


average speed, m/s

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


76 Page 14 of 16 Tribology Letters (2018) 66:76

and reported. The average WSD values from measurements was more viscous than BPAO-40, and the viscosity gap
along and across the wear direction of each of the three balls between the oils increased with decreasing temperature,
from each test are reported. Average and standard deviation from about 2 cSt at 100 °C to about 9 cSt at 75 °C, and to
values from duplicate runs on the functionalized BPAO-40 about 73 cSt at 40 °C.
and non-functionalized PAO-40 polyalphaolefins are sum- Physical property investigation showed that, relative to
marized in Table 5. The data show higher COF and WSD PAO-40, the functionalized BPAO-40 showed
values for the non-functionalized PAO-40 polyalphaolefin
compared to the values for functionalized BPAO-40. Since • higher density in the temperature range of 40–100 °C,
the AW test is conducted in the mixed film regime, both • higher VI,
hydrodynamic (viscosity) and boundary (polar chemical • lower oxidation stability, i.e., lower PDSC onset and
structure) properties of the lubricants will have a critical peak oxidation temperatures, which was attributed to
influence on the observed COF and WSD. At 75 °C, the incomplete hydrogenation of residual double bonds in
non-functionalized PAO-40 is more viscous (Table 2) and, BPAO-40, and
hence, will generate higher hydrodynamic friction (traction) • higher TAN, attributed to potential hydrolysis of methyl
than the functionalized BPAO-40. Unlike the functionalized esters by traces of moisture in the oil, especially during
BPAO-40, the non-functionalized PAO-40 cannot adsorb storage and/or transportation in high humidity condi-
onto friction surfaces and reduce friction and wear under tions.
boundary conditions. Thus, it will generate higher bound-
ary COF and wear in the boundary regime. Thus, based on In 4-ball AW tests, BPAO-40 displayed lower COF and
viscosity and chemical structure considerations, the non- lower WSD than PAO-40, which was attributed to the abil-
functionalized PAO-40 is predicted to produce a higher COF ity of the functionalized BPAO-40 to adsorb onto friction
and a higher WSD in the mixed film regime than the func- surfaces and lower boundary friction and wear.
tionalized BPAO-40, as is observed in Table 5. In EHD film thickness investigations, relative to PAO-
40, the functionalized BPAO-40 displayed:

4 Summary/Conclusion • lower EHD film thickness at 40 °C, 10 N load, and all


speeds, attributed to its lower viscosity at this tempera-
The properties of a BPAO-40, with a viscosity of 40 cSt at ture;
100 °C, were investigated relative to a commercial petro- • lower EHD film thickness at 40 °C, 40 N load, and
leum-based PAO of similar viscosity at 100 °C (PAO-40). ≥ 0.03 m/s speeds, attributed to its lower viscosity at
BPAO-40 is synthesized by oligomerization of a mixture this temperature;
of alpha olefins, with and without terminal methyl esters, • similar EHD film thickness at 40 °C, 40 N load, and
derived from vegetable oils via a biorefinery process. By < 0.03 m/s speeds, attributed to its functional property
contrast, commercial PAOs are synthesized only from non- to adsorb onto surfaces and form boundary films, even
functionalized alpha olefins. Thus, BPAO-40 is not only though it has a much lower viscosity at this tempera-
biobased, but also has a unique chemical structure, which ture;
makes it a functionalized PAO. The effect of chemical struc- • similar EHD film thickness at 75–100 °C, all loads (10–
ture (presence or lack of functionalization) on chemical, 40 N), and ≥ 0.03 m/s speeds, attributed to the similar
physical, and tribological properties of these two PAO base viscosity of the oils in this temperature range; and
oils was investigated. • higher EHD film thickness at 75–100 °C, all loads (10–
The two PAO oils were synthesized to have similar 40 N), and < 0.03 m/s speeds, attributed to its functional
kinematic viscosity of around 40 cSt at 100 °C, which property to adsorb onto surfaces and form a boundary
was found to be true. However, below 100 °C, PAO-40 film, even though both oils have similar viscosity at this
temperature.
Table 5  4-Ball anti-wear coefficient of friction (COF, ASTM D4172
[34], ASTM D5183 [36]) and wear scar diameter (WSD, ASTM The two oils displayed similar PVCs derived from EHD
D4172 [34]) of biobased (BPAO-40) and commercial petroleum- film thickness data, which was expected since both oils have
based (PAO-40) polyalphaolefins structures consistent with intermediate PVC oils that are nei-
BPAO-40 PAO-40 ther layered and shear (low PVC oils), nor entangled and/or
display strong intermolecular interactions (high PVC oils).
COF 0.073 ± 0.016 0.091 ± 0.020
In EHD traction, investigations as a function of entrain-
WSD (mm) 0.734 ± 0.029 0.815 ± 0.080
ment speed, 50% SRR, 40 N load, 40–100 °C temperature
All data from this work range, and < 2.25 m/s entrainment speed:

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Tribology Letters (2018) 66:76 Page 15 of 16 76

• Both oils displayed a lower traction coefficient at 100 °C used in this study. The author is grateful for the technical help from Ms.
compared to that at 40 °C, which was attributed to a Linda Cao, Ms. Amber Durham, Mr. Dan Knetzer, Mr. Kevin Steidley,
and Dr. Karl Vermillion.
lower viscosity of the oils at 100 °C
• Functionalized BPAO-40 displayed a lower traction coef-
ficient than PAO-40 at both temperatures, which was
attributed to its functionality producing lower boundary
friction, and, as a result, lower overall traction coeffi-
References
cient. 1. Rudnick, L.R. (ed.): Lubricant Additives, 3rd edn. CRC Press,
Boca Raton (2017)
In EHD traction investigations as a function of SRR, 2. Rudnick, L.R., Shubkin, R.L. (eds.): Synthetic Lubricants and
1 m/s entrainment speed, 40 N load: High-Performance Functional Fluids, 2nd edn. Marcel Dekker,
Inc., New York (1999)
3. Benton, J.: Changing trade winds for base oil. Lubes-n-Greases
• Both oils displayed lower a traction coefficient at 100 °C 24, 20–25 (2018)
relative to that at 40 °C, which was attributed to the lower 4. Sharma, B.K., Biresaw, G. (eds.): Environmentally Friendly and
viscosity at 100 °C Biobased Lubricants. CRC Press, Taylor and Francis Group, Boca
Raton (2016)
• Functionalized BPAO-40 displayed lower traction coef- 5. Brown, S.F.: Base oil groups: manufacture, properties and perfor-
ficient than PAO-40 at both temperatures, attributed to mance. TLT 71, 32–35 (2015)
its functionality producing lower boundary friction and, 6. Schneider, M.P.: Plant-oil-based lubricants and hydraulic fluids.
as a result, lower overall traction coefficient, at all SRR J. Sci. Food Agric. 86, 1769–1780 (2006)
7. Erhan, S.Z.: Oxidative stability of mid-oleic soybean oil: synergis-
values > 3%. tic effect of antioxidant–antiwear additives. In: Proceedings USB
Lube TAP (2006)
HFRR investigations at 75 °C and 10 N load, as a func- 8. Honary, L.A.T.: Biobased greases and lubricants: from research
tion of average speed, gave COF, WSD, and relative film to commercialization. In: Proceedings USB Lube TAP (2007)
9. Hope, K., Garmier, B.: High performance engine oils with veg-
thickness values that were in line with the EHD traction and etable oil and PAO blends. In: STLE 63rd Annual Meeting and
film thickness results. Thus, relative to the non-functional- Exhibition, Cleveland, OH, 18–22 May, Program Guide, p. 180
ized PAO-40, the functionalized BPAO-40 displayed: (2008)
10. Lawate, S.S., Lal, K., Huang, C.: Vegetable oils—structure and
performance. In: Booser, E.R. (ed.) Tribology Data Handbook,
• lower COF at lower speeds (< 0.12 m/s), but similar COF pp. 103–116. CRC Press, New York (1997)
at higher speeds (> 0.12 m/s), which is in line with the 11. Gast, L.E., Schneider, W.J., Forest, C.A., Cowan, J.C.: Composi-
trend observed in the traction versus entrainment speed tion of methyl esters from heatbodied linseed oils. J. Am. Oil
data; Chem. Soc. 40, 287 (1963)
12. Adhvaryu, A., Erhan, S.Z., Perez, J.M.: Tribological studies of
• higher relative film thickness at lower speeds thermally and chemically modified vegetable oils for use as envi-
(< 0.12 m/s), but similar relative film thickness at higher ronmentally friendly lubricants. Wear 257, 359 (2004)
speeds (> 0.12 m/s), which is in line with the EHL film 13. Arca, M.: Oxidative properties and thermal polymerization of soy-
thickness data at high temperatures; and bean oil and application in gear lubricants. M.S. Thesis, Pennsyl-
vania State University, University Park, PA (2011)
• lower WSD at lower speeds (< 0.12 m/s), but similar 14. Arca, M., Sharma, B.K., Perez, J.M., Doll, K.M.: Gear oil formu-
WSD at higher speeds (> 0.12 m/s), which is in line with lation designed to meet bio-preferred criteria as well as give high
the COF results summarized above. performance. Int. J. Sustain. Eng. 6, 326 (2013)
15. Erhan, S.Z., Adhvaryu, A., Sharma, B.K.: Chemically function-
alized vegetable oils. In: Rudnick, L.S. (ed.) Synthetics, Min-
Overall, relative to the commercial non-functionalized eral Oils, and Bio-based Lubricants Chemistry and Technology,
PAO-40, the functionalized BPAO-40 displayed the follow- pp. 361–387. CRC Press, Boca Raton (2006)
ing properties: higher density at 40–100 °C, higher VI, lower 16. Cermak, S., Isbell, T.: Estolides: the next biobased functional
number average molecular weight (Mn), higher PDI, lower fluid. INFORM 15, 515–517 (2004)
17. Isbell, T., Cermak, S.: Synthesis of triglyceride estolides from
oxidation stability (PDSC OT and PT), higher TAN, higher Lesquerella and castor oils. J. Am. Oil Chem. Soc. 79, 1227–1233
FFA, lower 4-ball AW COF and WSD, higher EHD lubricant (2002)
film thickness under boundary conditions (low speeds and 18. Cermak, S., Kendra, B., Isbell, T.: Synthesis and physical proper-
high temperature), lower EHD traction coefficient at 40 and ties of estolides from Lesquerella and castor fatty acid esters. Ind.
Crops Prod. 23, 54–64 (2006)
100 °C, similar PVC, lower COF, lower WSD, and higher 19. Bredsguard, J.W., Thompson, T.D., Cermak, S.C., Isbell, T.A.:
relative film thickness on HFRR tribometer under boundary Estolides: bioderived synthetic base oils. In: Sharma, B.K., Bire-
conditions (low speeds). saw, G. (eds.) Environmentally Friendly and Biobased Lubricants,
pp. 35–49. CRC Press, Boca Raton (2016)
Acknowledgements The author thanks Elevance Renewables, LLC, 20. Biosynthetic Technologies: Biosynthetic Base Oil (2017). http://
for providing free samples of the biobased and petroleum-based PAOs biosy​nthet​ic.com/biosy​nthet​ic-base-oil/. Accessed 21 Dec 2017

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


76 Page 16 of 16 Tribology Letters (2018) 66:76

21. Joassard, A., Kupiec, D.: New base oils for the aluminum industry: 37. Biresaw, G., Bantchev, G., Murray, R.: Investigation of biobased
moving along the vegetable road. In: STLE 71st Annual Meeting and petroleum base oils in the entire spectrum of lubrication
and Exhibition. Program Guide and Schedule, p. 115 (2016) regimes. J. Am. Oil Chem. Soc. 94, 1197–1208 (2017)
22. Candelaria, K.: Advonex Aims for Biobased Lube Production. 38. Biresaw, G.: Elastohydrodynamic properties of seed oils. J. Am.
Lube Report, 5 October 2016 (2017). http://pubs.lubes​ngrea​ses. Oil Chem. Soc. 83, 559–566 (2006)
com/lubere​ port/​ 16_40/bio-synthe​ tics/​ -11119-​ 1.html. Accessed 23 39. Biresaw, G., Bantchev, G.: Elastohydrodynamic (EHD) traction
Jan 2017 properties of seed oils. Tribol. Trans. 53, 573–583 (2010)
23. Joshi, C.H., Gibson, G.T.T., Malvich, D., Horner, M.G.: High 40. Johnston, G.J., Wayte, R., Spikes, H.A.: The measurement and
productivity Kolbe reaction process for transformation of fatty study of very thin lubricant films in concentrated contacts. Tribol.
acids derived from plant oil and animal fat. US Patent 8,961,775 Trans. 34, 187 (1991)
B2, 24 Feb 2015 41. Rudnick, L.S., Shubkin, R.L.: Poly(α-olefins). In: Rudnick, L.S.,
24. Brown, J., Hahn, H., Vettel, P., Wells, J.: Farnesene-derived Shubkin, R.L. (eds.) Synthetic Lubricants and High-Performance
base oils. In: Sharma, B.K., Biresaw, G. (eds.) Environmentally Functional Fluids, 2nd edn, pp. 3–52. Marcel Dekker, Inc., New
Friendly and Biobased Lubricants, pp. 3–34. CRC Press, Boca York (1999)
Raton (2016) 42. Noor, M.A.M., Sendijarevic, V., Hoong, S.S., Sendijarevic, I.,
25. Novvi: NovaSpec™ Base Oils. http://novvi​.com/novas​pec-base- Ismail, T.N.M.T., Hanzah, N.A., Noor, N.M., Palam, K.D.P.,
oils/. Accessed 21 Dec 2017 Ghazali, R., Hassan, H.A.: Molecular weight determination of
26. Moon, M.: Novel ester-functionalized HV PAO base stock for palm olein by gel permeation chromatography using polyether
advanced lubricant formulations. TLT 73, 72–74 (2017) polyols calibration. J. Am. Oil Chem. Soc. 93, 721–730 (2016)
27. Havelka, K.O., Gerhardt, G.E.: From biorefinery to performance 43. Anon: Product Datasheet, SpectraSyn 40, pp. 1–2. ExxonMobil
technology: transforming renewable olefinic building blocks into (2009)
lubricants and other high-value products. ACS Symp. Ser. 1192, 44. Hamrock, B.T., Dowson, D.: Ball Bearing Lubrication: The Elas-
201–222 (2015) tohydrodynamics of Elliptical Surfaces. Wiley, New York (1981)
28. Elevance Renewable Sciences: Technology. https:​ //elevan​ ce.com/ 45. Biresaw, G., Bantchev, G.B.: Pressure viscosity coefficient of veg-
techn​ology​/. Accessed 22 Dec 2017 etable oils. Tribol. Lett. 49, 501 (2013)
29. A.O.C.S. Official: Method Te 2a-64, Acid Value, p. 1 (1997) 46. Jones, W.R., Johnson, R.L., Winer, W.O., Sanborn, D.M.: Pres-
30. ASTM D7042-11a: Standard test method for dynamic viscosity sure viscosity measurements for several lubricants to 5.5 × 108
and density of liquids by Stabinger viscometer (and the calcula- Newtons per square meter (8 × 104 PSI) and 149 °C (300 °F).
tion of kinematic viscosity). In: Annual Book of ASTM Standards ASLE Trans. 18, 249 (1975)
05.04, pp. 186–193. American Society for Testing and Materials, 47. Bair, S.: The high pressure rheology of some simple model hydro-
West Conshohocken (2012) carbons. Proc. Inst Mech. Eng. J 216, 139–149 (2002)
31. ASTM D2270-93: Standard practice for calculating viscosity 48. Cooper, D., Moore, A.J.: Application of the ultra-thin elastohydro-
index from kinematic viscosity at 40 and 100 °C. In: Annual Book dynamic oil film thickness technique to the study of automotive
of ASTM Standards, 05.01, pp. 849–854. American Society for engine oils. Wear 175, 93 (1994)
Testing and Materials, West Conshohocken (2012) 49. Gunsel, S., Korcek, S., Smeeth, M., Spikes, H.A.: The elastohy-
32. ASTM D6186-08: Standard test method for oxidation induction drodynamic friction and film forming properties of lubricant base
time of lubricating oils by pressure differential scanning calorim- oils. Tribol. Trans. 42, 559 (1999)
etry (PDSC). In: Annual Book of ASTM Standards, 05.03, pp. 50. So, B.Y.C., Klaus, E.E.: Viscosity–pressure correlation of liquids.
52–56. American Society for Testing and Materials, West Con- ASLE Trans. 23, 409 (1980)
shohocken (2012) 51. Wu, C.S., Klaus, E.E., Duda, J.L.: Development of a method for
33. Biresaw, G., Laszlo, J.A., Evans, K.O., Compton, D.L., Bantchev, the prediction of pressure–viscosity coefficients of lubricating oils
G.B.: Synthesis and tribological investigation of lipoyl glycerides. based on free-volume theory. Trans. ASME J. Tribol. 111, 121
J. Agric. Food Chem. 62, 2233–2243 (2014) (1989)
34. ASTM D4172-94. Standard test method for wear preventive char- 52. Johnston, W.G.: A method to calculate the pressure–viscosity
acteristics of lubricating fluid (four-ball method). In: Annual Book coefficient from bulk properties of lubricants. ASLE Trans. 24,
of ASTM Standards, 05.02, pp. 752–756. American Society for 232 (1981)
Testing and Materials, West Conshohocken (2002) 53. Bair, S.: Personal communication (2013)
35. Biresaw, G., Bantchev, G.B.: Tribological properties of biobased 54. ASTM D974: Standard test method for acid and base number by
ester phosphonates. J. Am. Oil Chem. Soc. 90, 891–902 (2013) color-indicator titration. Annual Book of ASTM Standards, 05.01,
36. ASTM D5183-95: Standard test method for determination of the pp. 382–388. American Society for Testing and Materials, West
coefficient of friction of lubricants using the Four-Ball wear test Conshohocken (2012)
machine. In: Annual Book of ASTM Standards, 05.03, pp. 165–
169. American Society for Testing and Materials, West Consho-
hocken (2002)

13

Content courtesy of Springer Nature, terms of use apply. Rights reserved.


Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:

1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at

onlineservice@springernature.com

You might also like