2003 - Synthesis and Physical Properties of Estolide-Based Functional Fluids PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Industrial Crops and Products 18 (2003) 183 /196

www.elsevier.com/locate/indcrop

Synthesis and physical properties of estolide-based functional


fluids
Steven C. Cermak *, Terry A. Isbell
New Crops and Processing Technology Research, National Center for Agricultural Utilization Research, USDA, Agricultural Research
Service, 1815 N. University St., Peoria, IL 61604, USA

Received 19 November 2002; accepted 12 May 2003

Abstract

Biodegradable, vegetable oil-based lubricants must have better low temperature properties as well as comparable cost
to petroleum oil before they can become widely acceptable in the marketplace. These include low pour point
temperature, low viscosity and viscosity indices that do not change over a wide temperature range. Our objective was to
synthesize estolides from various sources with improved lubricating physical properties. Oleic acid and lauric acid were
treated with varying equivalents of perchloric acid at 60 8C to produce complex estolides. Yields ranged between 45 and
75% after purification by Kugelrohr distillation. The estolide number (EN), the average number of fatty acid units
added to a base fatty acid, varied with reaction conditions. The saturate-capped, oleic estolides were esterified with 2-
ethylhexanol to obtain high yields of the corresponding ester. Coconut /oleic 2-ethylhexyl estolide esters were produced
by varying the ratio of oleic and coconut fatty acids with 0.05 equivalents of HClO4 to give estolides with excellent cold
temperature properties. The amount of oligomerization (EN) played an important role on viscosity; viscosity increased
with higher oligomerization. The free acid estolides were generally several hundred centistokes (cSt) more viscous than
their corresponding esters. The viscosity index ranged from 141 to 170 for the free acid estolides, whereas the complex
estolide 2-ethylhexyl esters had slightly higher viscosity indices, which ranged from 159 to 232. These new coco /oleic
estolide esters displayed superior low temperature properties (/36 8C), were of reasonable cost, and were more suitable
as a base stock for biodegradable lubricants and functional fluids than current vegetable oil-based commercial
materials.
Published by Elsevier B.V.

Keywords: Cloud point; Coconut; Estolides; Lauric; Oleic; Pour point; Viscosity


Names are necessary to report factually on available data; 1. Introduction
however, the USDA neither guarantees nor warrants the
standard of the product, and the use of the name by USDA Extensive research have been conducted on
implies no approval of the product to the exclusion of others biodegradable, renewable functional fluids. Many
that may also be suitable.
groups have tried to use the cheap commodity
* Corresponding author. Tel.: /1-309-681-6233; fax: /1-
309-681-6524. prices of soybeans to force this traditional US
E-mail address: cermaksc@ncaur.usda.gov (S.C. Cermak). farmland oil into a functional fluid market (Tilton,
0926-6690/03/$ - see front matter. Published by Elsevier B.V.
doi:10.1016/S0926-6690(03)00061-X
184 S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196

2002). Several factors have contributed to this temperature properties (Fig. 2). Yields varied
movement: namely, vegetable oil-based lubricants between 45 and 65% after Kugelrohr distillation.
and derivatives have excellent lubricity and biode- These new saturated complex estolides and the
gradability properties (Mang, 1994, 1997; Legrand simple estolide esters showed very different physi-
and Durr, 1998; Canter, 2001). Two major pro- cal properties. By varying the capping material on
blems exist with vegetable oils as functional fluids the estolide, the crystal lattice structure of the
related to low resistance to thermal oxidative material was disrupted at low temperatures, which
stability (Becker and Knorr, 1996) and poor low led to estolides with excellent low temperature
temperature performance (Asadauskas and Erhan, properties.
1999; Zehler, 2001). However, these properties These new, complex estolides have certain
sometimes can be improved with additives, but physical characteristics that could help eliminate
only at the expense of biodegradability, toxicity, common problems associated with vegetable oils
and price. as functional fluids (Cermak and Isbell, 2002). One
Estolides and estolide esters from meadowfoam of the concerns with vegetable oils is their oxida-
fatty acids (Isbell and Kleiman, 1996), oleic acid tion stability. Simple homo-oleic estolides, when
(Isbell et al., 2001), castor oil, or any source of formulated with a small amount of oxidative
hydroxy fatty acids (Zoleski and Gaetani, 1984) stability package, show better oxidative stability
have been explored and show promise as cos- than both petroleum and vegetable oil-based fluids
metics, coatings, and biodegradable lubricants. (Cermak and Isbell, 2001a), but improvements can
Estolides are formed when the carboxylic acid still be made. The oxidative stability of an oil can
functionality of one fatty acid links to the site of be improved by several ways. Akoh (1994) in-
unsaturation of another fatty acid to form oligo- creased the oxidative stability index of refined
meric esters (Fig. 1). The estolide number (EN) is soybean oil from 9.4 h at 110 8C to 15.3 h by
defined as the average number of fatty acids added partial hydrogenation. A similar approach was
to the base fatty acid (Fig. 1, EN /n/1). taken with the oleic estolides, where hydrogena-
Complex estolides have been synthesized by tion with 2% w/w of 10% palladium on activated
Cermak and Isbell (2001a,b) where oleic acid and carbon as catalyst gave completely saturated
various saturated fatty acids, butyric through estolides (Isbell et al., 2001). The saturated oleic
stearic, were treated with 0.4 equivalents of per- estolides are expected to be more oxidatively stable
chloric acid at either 45 or 55 8C to produce a new than the unsaturated estolides, assuming the same
class of saturated estolides with superior low trend displayed by soybean oil holds true.

Fig. 1. Reaction scheme for the formation of oleic estolides.


S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196 185

Fig. 2. Reaction scheme for the formation of complex estolides and estolide 2-ethylhexyl esters.

Other low cost and relatively available sources a complex estolide mixture that would have even
of saturated fatty acids is necessary where coconut greater low temperature properties than the simple
oil could serve this need. Coconut fatty acids are oleic estolide.
relatively cheap and readily available (Tilton, Estolides and estolide esters have compared
2002). The fatty acid profile of coconut is primar- favorably to commercially available industrial
ily saturated fatty acids ranging from C8 to C18 products such as petroleum-based hydraulic fluids,
with the majority being C12 and C14, Table 1. soy-based fluids and petroleum oils. In this paper,
Cermak and Isbell (2002) reported that the C12 we report the easy and cost effective synthesis of
and C14 fatty acids made excellent low tempera- complex estolides and corresponding esters along
ture estolides. Coconut oil would make an ex- with their physical properties, such as pour and
cellent fatty acid alternative source due to it’s fatty cloud points, viscosity, and other low temperature
acid distribution (Table 1). It can be used to create properties. These new coco /oleic estolides, which
should have superior biodegradability and lubri-
Table 1 cating properties than petroleum products, were
Fatty acid composition of coco fatty acids compared with commercially available materials
based on their low temperature properties.
Fatty acid composition GC (% of total)

C 8:0 5.1
C 10:0 4.9
2. Materials and methods
C 12:0 48.3
C 14:0 21.7
C 16:0 10.6 2.1. Materials
C 18:0 7.0
C 18:1 2.4
Oleic acid (90%) and N ,O -bis(trimethylsilyl)a-
Determined by GC (SP-2380, 30 m /0.25 mm i.d.). cetamide (derivatization grade) were purchased
186 S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196

from Aldrich Chemical Co. (Milwaukee, WI). hold 5 min at 265 8C; injector and detector
Ethyl acetate and hexanes (for extractions), ace- temperatures set at 250 8C. Retention times for
tone (for high performance liquid chromatogra- eluted peaks with ECL values in parentheses were:
phy, HPLC), perchloric acid (70%), concentrated methyl oleate 9.73 min (18.36), hydroxy methyl
sulfuric acid (98%), methanol and 2-ethylhexyl oleate 14.10 (27.81), g-stearolactone 15.31 min
alcohol, were purchased from Fisher Scientific Co. (30.39), and d-stearolactone 15.72 min (31.59)
(Fair lawn, NJ). Potassium hydroxide was ob- (Cermak and Isbell, 2002).
tained from J.T. Baker Chemical Co. (Phillips-
burg, NJ). Filter paper was obtained from 2.2.2. GC /mass spectrometry
Whatman (Clifton, NJ). Acetonitrile and acetic GC /mass spectrometry (GC /MS) was per-
acid (both for HPLC), charcoal, sodium hydro- formed on a Hewlett-Packard 5890A GC with a
genphosphate, and sodium dihydrogenphosphate 30 m/0.20 mm i.d. SPB-1 column (Supelco) and
were obtained from EM Science (Gibbstown, NJ). a Hewlett-Packard 5970 mass selective detector.
Ethanol was purchased from AAPER Alcohol and GC conditions: helium head pressure 103.4 kPa at
Chemical Company (Shelbyville, KY). Pyridine 170 8C set for constant flow with varying pressure;
was purchased from Mallinckrodt (Paris, KY). split ratio 50:1; injector temperature set at 250 8C;
The fatty acid methyl ester (FAME) standard transfer line temperature set at 250 8C; pro-
mixtures were obtained from Alltech Associates, grammed ramp from 170 to 270 8C at 3 8C/min.
Inc. (Deerfield, IL). Solvents for chromatography MS conditions: mass range 50 /550 amu; electron
and extraction were HPLC grade or equivalent, multiplier 200 V relative.
and were used without further purification. Petro-
leum oil: Mobil† 10W-30 and synthetic oil: 2.2.3. High performance liquid chromatography
Castrol Synthetic† 10W-30 were obtained from Reverse phase HPLC, analyses were performed
Wal-Mart Department Store† (Peoria, IL). Soy- on a Thermo Separations Spectra System AS1000
based oil: Biosoy† was obtained as a free sample autosampler/injector (Fremont, CA) with a P2000
from University of Northern Iowa (Cedar Falls, binary gradient pump from Thermo Separation
IA). Hydraulic fluid: Traveler Universal Hydraulic Products (Fremont, CA) coupled to a Alltech
Fluid† was obtained from Tractor Supply ELSD 500 evaporative light scattering detector
Company† (Peoria, IL). (Alltech Associates). A C-8 reverse phase analysis
used to separate reaction mixtures was carried out
2.2. Equipment and procedures with a Dynamax column (250 mm /4.5 mm, 5-m
particle size) from Rainin Instrument Co. (Wo-
2.2.1. Gas chromatography burn, MA).
Gas chromatography (GC) was performed with Two methods for reverse phase analysis were
a Hewlett-Packard 5890 Series II gas chromato- used to separate the reaction mixtures. Method A
graph (Palo Alto, CA) equipped with a flame- (16 min run time) was used to follow the reaction.
ionization detector and an autosampler/injector. It provided information on the overall progress of
Analyses were conducted on a SP-2380 30 m / the reaction. Method B (35 min run time) pro-
0.25 mm i.d. column (Supelco, Bellefonte, PA). duced a more detailed separation of the reaction
Saturated C8 /C30 FAMEs provided standards for mixture, in particular a separation of estolides,
calculating equivalent chain length (ECL) values, lactones, fatty acids, and hydroxy fatty acids.
which were used to make fatty acid and by- Operating parameters for method A were: flow
product assignments. rate of 1 ml/min; 0/4 min 80% acetonitrile, 20%
Operating parameters for SP-2380 analysis acetone; 6 to 10 min 100% acetone; 11/16 min
were: column flow 3.3 ml/min with helium head 80% acetonitrile, 20% acetone. The ELSD drift
pressure of 103.4 kPa; split ratio 22:1; pro- tube was set at 55 8C with the nebulizer set at 137.9
grammed ramp 75 /165 8C at 15 8C/min, 165/ kPa N2, providing a flow rate of 2.0 standard liters
185 8C at 7 8C/min, 185/265 8C at 15 8C/min, per minute (SLPM). Retention times for eluted
S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196 187

peaks: estolides, 9.8 /13.0 min; methyl oleate, 6.3 PA). Viscosity measurements were made in a
min; oleic acid, 5.1 min; lactones, 4.8 min; and Temp-Trol (Precision Scientific, Chicago, IL)
hydroxy acids 4.1 min. viscometer bath set at 40.0 and 100.0 8C. Viscosity
Operating parameters for method B were: flow and viscosity index were calculated using ASTM
rate of 1 ml/min; 0 /2 min 60% acetonitrile 40% methods D445-97 (ASTM, 1997) and ASTM
acetone; 20 /25 min 100% acetone; 30/35 min 60% D2270-93 (ASTM, 1998), respectively. Duplicate
acetonitrile 40% acetone. The ELSD drift tube was measurements were made and the average value
set at 55 8C with the nebulizer set at 137.9 kPa N2, was reported.
providing a flow rate of 2.0 SLPM. Retention
times for eluted peaks: estolides, 8.2 /25.6 min;
2.2.6. Pour point
methyl oleate, 5.5 min; oleic acid, 4.8 min; g-
Pour points were measured by ASTM method
lactones, 4.5 min; d-lactones, 4.1 min; and hydroxy
D97-96a (ASTM, 1996) to an accuracy of 9/3 8C.
acids 3.8 min.
The pour points were determined by placing a test
Normal phase HPLC analyses were performed
jar with 50 ml of the sample into a cylinder
using a Spectra-Physics 8800 ternary pump (San
submerged in a cooling medium. The sample
Jose, CA) with a Spectra System AS3000 auto-
temperature was measured in 3 8C increments at
sampler/injector from Thermo Separation Pro-
the top of the sample until the material stopped
ducts coupled to a Varex ELSD III light
pouring. The sample no longer poured when the
scattering detector (Alltech Associates). A silica
material in the test jar did not flow when held in a
normal phase analysis was carried out with a
horizontal position for 5 s. The temperature of the
Dynamax column (250 mm /4.6 mm, 8 mm)
cooling medium was chosen based on the expected
from Rainin Instrument Co. Components were
pour point of the material. Samples with pour
eluted isocratically from the column with a 80:20
points that ranged from (/9 to /6, /6 to /24,
mixture hexanes/acetone at a flow rate of 1 ml/min
and /24 to /42 8C) were placed in baths of
with the ELSD drift tube set at 45 8C and
temperature (/18, /33, and /51 8C), respec-
nebulizer set at 69.0 kPa N2, flow rate 1.50 SLPM.
tively. The pour point was defined as the coldest
Normal phase HPLC was used to separate
temperature at which the sample still poured. All
esterification reaction mixtures. Retention times
pour points were run in duplicate and average
for eluted peaks: 2-ethylhexyl estolide ester, 2.6 /
values were reported.
2.8 min; and estolide, 3.5 /3.7 min depending on
the capping fatty acid.
2.2.7. Cloud point
2.2.4. Gardner color Cloud points were determined by ASTM
Gardner color was measured on a Lovibond 3- method D2500-99 (ASTM, 1999) to an accuracy
Field Comparator from Tintometer Ltd (Salis- of 9/1 8C. The cloud points were determined by
bury, England) using AOCS method Td 1a-64 placing a test jar with 50 ml of the sample into a
(Firestone, 1994b). Gardner color of both the cylinder submerged into a cooling medium. The
residue and distillate materials was measured sample temperature was measured in 1 8C incre-
throughout the distillation. The / and / notation ments at the bottom of the sample until any
was employed to designate samples that did not cloudiness was observed at the bottom of the test
match one particular color or, in the case of the jar. The temperature of the cooling medium was
residue, an 18/ represented a color darker than chosen based on the expected cloud point of the
the upper limit of 18. material. Samples with cloud points that ranged
from (room temperature to 10, 9 to /6, and /6 to
2.2.5. Viscosity /24, /24 to /42 8C) were placed in baths of
Viscosity was determined with calibrated temperature (0, /18, /33, and /51 8C), respec-
Cannon /Fenske viscometer tubes purchased tively. All cloud points were run in duplicate and
from Cannon Instrument Co. (State College, average values were reported.
188 S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196

2.2.8. Acid value connected to a recirculating constant temperature


The acid values were measured on a 751 GPD bath maintained at 459/0.1 8C. All reactions
Titrino from Metrohm Ltd (Herisau, Switzerland). described were mixed with an overhead stir motor
Acid values were determined by the AOCS using a glass shaft and a Teflon blade. The
Method Te 2a-64 (Firestone, 1994a) with ethanol reactions were conducted at atmospheric pressure
substituted for methanol to increase the solubility in a sealed flask. Reactions were performed under
of the estolide ester during the titration. All acid the general conditions described previously while
values were run in duplicate and average values varying the equivalents of perchloric acid as
were reported. reported in Table 2. In most cases, oleic acid
(100.0 g, 354.0 mmol) and saturated fatty acids,
2.2.9. Iodine values and EN i.e. lauric, (35.5 g 177.0 mmol) were combined
Iodine values (IV) were calculated from GC together and heated to the desired temperature as
data using AOCS Method Cd 1c-85 (Firestone, outlined in Table 2. Once the temperature was
1994c). ENs were determined by GC from the SP- reached, perchloric acid was added and the flask
2380 column analysis described by Isbell and was stoppered. Product distribution was moni-
Kleiman, (1994). tored by HPLC, GC and/or GC/MS. Completed
reactions were quenched by the addition of 0.5 M
2.2.10. GC analysis of hydroxy fatty acids Na2HPO4 (212.4 mmol, 424.8 ml) to the reaction
Analytical estolide samples for GC were pre- vessel. The reactor was disconnected from the
pared by heating a 10 mg sample of estolide or circulating bath and the solution was allowed to
estolide 2-ethylhexyl ester in 0.5 ml of 0.5 M KOH/ cool with stirring for 30 min. The material was
MeOH to reflux on a heating block for 60 min in a transferred to a separatory funnel followed by the
sealed vial. After cooling to room temperature, 2 addition of 200 ml of a 2:1 ethyl acetate:hexanes
ml of 1 M H2SO4/MeOH was added to the vial, the solution. The pH of the organic layer was adjusted
vial was resealed and heated to reflux on a heating to 5.3 /6.0 with the aid of a pH 5 buffer
block for 30 min. The solution was transferred to a (NaH2PO4, 519 g in 4 l H2O, 2 /50 ml) followed
separatory funnel with water (1 ml) and washed by saturated NaCl (2 /50 ml). The organic layer
with hexanes (2 /2 ml), dried over sodium sulfate, was dried over sodium sulfate and filtered. All
gravity filtered, placed in a GC vial with hexanes, reactions were concentrated in vacuo then Kugel-
sealed, and injected into the GC and/or GC/MS. rohr-distilled at 160 /90 8C at 0.0131 /0.067 kPa to
remove any lactones, saturated and unsaturated
2.2.11. TMS derivatization of hydroxy fatty esters fatty acids.
Hydroxy FAMEs (10 mg) (prepared as de-
scribed) were dissolved in 0.1 ml of pyridine and
0.05 ml N ,O -bis(trimethylsilyl)acetamide. This 2.2.12.2. Estolide 2-ethylhexyl ester. The distilled,
solution was then placed in a sealed vial for 10 free acid estolides were combined with a 0.5 M
min at 100 8C. After this time hexanes (1 ml) were BF3/2-ethylhexyl alcohol solution (3 /estolide wt,
added, and the resulting solution was filtered w/v) in a 500 ml round bottom flask. The reactions
though a silica plug. The filtered sample was were conducted at 60 8C with magnetic stirring
placed in a sealed GC vial with hexanes and and were monitored hourly by normal phase
injected onto the GC and/or GC/MS. HPLC (Table 3). Esterification reactions were
run until /99% complete then were transferred
2.2.12. Synthesis to a separatory funnel and were washed with
saturated NaCl (2 /75 ml). The pH of the organic
2.2.12.1. Free-acid estolide. Acid-catalyzed con- layer was adjusted to 5.3 /6.0 with pH 5 buffer
densation reactions were conducted without sol- (NaH2PO4, 519 g in 4 l H2O, 2/50 ml). The oil
vent in a 500 ml, baffled, jacketed reactor with a was dried over sodium sulfate and filtered. All
three-neck reaction kettle cover. The reactor was reactions were concentrated in vacuo, then Ku-
S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196 189

Gardner color
gelrohr-distilled at 100/120 8C at 0.013 /0.067 kPa
to remove any excess 2-ethylhexyl alcohol.

NR
2.2.12.3. Estolide 2-ethylhexyl ester (one step).

12
8
8
8
Acid-catalyzed condensation reactions were con-
Viscosity index

ducted without solvent in a 500 ml, baffled,

Reactions were run for 24 h with overhead stirring in a 2:1 molar ratio, oleic:lauric fatty acids at 45 8C followed by distillation at 180 /200 8C.
jacketed reactor with a three-neck reaction kettle
cover. The reaction was connected to a recirculat-
ing, constant temperature bath maintained at 9/
NR
143
143
144
143

0.1 8C of the set point. All reactions described were


Vis@100 8C (cSt)

mixed with an overhead stir motor using a glass


shaft and a Teflon blade as reported in Table 5 and
Table 6. In most cases, oleic acid (100.0 g, 354.0
mmol) and saturated fatty acids, lauric, (35.5 g,
31.0
26.2
22.8
22.4
NR

177.0 mmol) were combined together and heated


to 60 8C under house vacuum, as in Table 6. Once
Vis@40 8C (cSt)

the desired temperature was reached, perchloric


acid (0.05 eq., 26.5 mmol, 2.3 ml, Table 5) was
added and the flask was placed under vacuum and
297.9
235.7
192.5
189.8

stirred for 24 h. After 24 h, 2-ethylhexyl alcohol


NR
Acid-catalyzed condensation reaction with oleic and lauric acid with varying amounts of HClO4

(59.6 g, 457.6 mmol, 67.5 ml) was added to the


vessel, vacuum was restored, and the mixture was
Cloud point (8C)

stirred for 2 additional h. The completed reactions


were quenched by the addition of KOH (22.3
mmol, 1.25 g, 1.2 equivalent based on HClO4) in
/26
/28
/22
/24
NR

90% ethanol/water (10 ml) solution. The reactor


was disconnected from the circulating bath and the
Pour point (8C)

solution was allowed to cool with stirring for 30


min. The material was filtered through a Buckner
Yield based on mass of pure estolide obtained via distillation.

funnel with Whatman #1 filter paper. The organic


layer was dried over sodium sulfate and filtered.
/21
/27
/24
/24
NR

All reactions were concentrated in vacuo then


Kugelrohr-distilled at 160/190 8C at 0.013 /0.067
GC (ENb)

kPa to remove any lactones, saturated and un-


2.23
1.85
1.65
1.53
NR

saturated fatty acids, and any excess 2-ethylhexyl


alcohol.
Estolides (%a)

2.2.12.4. Decolorization of 2-ethylhexyl estolide


esters. Gardner color was measured on a Lovibond
3-Field Comparator from Tintometer Ltd. Gard-
93
57
54
56
0

ner color of the 2-ethylhexyl estolide esters was


Estolide Equiv. HClO4

measured before and after decolorization. To


Estolide number.

remove color, the estolides were placed in a 500-


ml Erlenmeyer flask with 2% w/w of activated
0.05
0.01
0.4
0.2
0.1

carbon (Table 3) then stirred with a magnetic


stirrer for 4 h at room temperature. Upon
Table 2

completion, hexanes (100 ml) were added and the


b
a

mixture was filtered through a silica plug. The


D
A

C
B

E
190 S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196

Table 3
Esterification of lauric /oleic estolides with 2-ethylhexyl alcohol

Estolide GC GC Capped Gardner Decolorized Gardner color Acid value


estera (ENb) iodine value (%c) color Gardner color improvement (mg/g)

A-2EH 2.26 12.0 61.5 9 5 4 1.01


B-2EH 1.69 18.5 60.9 10 5 5 1.27
C-2EH 1.48 20.4 59.9 10 5 5 1.35
D-2EH 1.47 21.8 63.3 12 10 2 2.33

Esterification reactions were run with magnetic stirring and 0.5 M BF3 at 60 8C.
a
Estolide 2-ethylhexyl ester.
b
Estolide number.
c
Ratio of estolide capped with saturated fatty acids determined by GC (SP-2380, 30 m/0.25 mm i.d.).

organic layer was dried over sodium sulfate and (t), 25.3 (t), 25.2 (t), 24.9 (t), 24.7 (t), 24.6 (t), 22.6
gravity filtered. All reactions were concentrated in (t) and 13.9 ppm (q, /C H3).
vacuo to remove excess hexanes.
2.2.13.2. 1H NMR and 13C of estolide 2-ethylhexyl
2.2.13. Nuclear magnetic resonance ester OC-EH-DD (Table 7). 1H NMR: d 5.37 /
1
H and 13C NMR spectra were obtained on a 5.34 (m, 0.3H, /CH /CH /), 4.87 /4.81 (m, 1.0H,
Bruker ARX-400 (Karlsruhe, Germany) with a 5 /CH /OC /O /), 3.96 (d, J/5.7 Hz, 1.5H, /
mm dual proton/carbon probe (400 MHz 1H/ OCH2 /CH(CH2 /)CH2 /), 2.29 /2.24 (m, 4.1H, /
100.61 MHz 13C) using CDCl3 as a solvent in all CH2(C /O) /O /CH2 /, /CH2(C /O) /O /CH /),
experiments. The assignment of protons was not to 1.96 /1.24 (m, 55.7H), and 0.89 /0.85 ppm (m,
the whole number. The representative nuclear 10.7H, /CH3). 13C NMR: d 174.0 (s, C /O), 173.5
magnetic resonance (NMR) contained a com- (s, C /O), 130.0 (d, /C H/C H /, very small
pound that had an average EN of 1.99 for the signals, only a small amount of alkene present),
free estolide (I-2EH) and 1.94 for the estolide ester 73.9 (d, /C H/O /C /O), 66.5 (t, /O /C H2 /
(OC-EH-DD), which made whole number assign- CH /), 38.6 (d, /CH2 /C H(CH2 /)/CH2 /), 34.3
ment impossible. The data reported for the num- (t), 34.0 (t), 31.8 (t), 30.3 (t), 29.6 (t), 29.5 (t), 29.5
ber of protons in the NMR reflected the actual (t), 29.4 (t), 29.4 (t), 29.3 (t), 29.2 (t), 29.2 (t), 29.1
numbers. The numbers could be multiplied by a (t), 29.0 (t), 28.8 (t), 25.2 (t), 24.9 (t), 23.7 (t), 22.8
factor to obtain whole numbers that correspond to (t), 22.5 (t), 14.0 (q, /C H3), 13.9 (q, /C H3), and
a whole number EN. 10.9 ppm (q, /C H3).

2.2.13.1. 1H and 13C NMR of free acid estolide D


(Table 2). 1H NMR: d 5.37 /5.33 (m, 0.5H, / 3. Results and discussion
CH /CH /), 4.86 /4.83 (m, 1.7H, /CH /OC /O /
CH2 /), 2.32 (t, J /7.4 Hz, 2H, /CH2(C /O) / A series of reactions (Fig. 2) that explored the
OH), 2.28 /2.23 (m, 3.4H, /CH2(C /O) /O / formation of new complex estolides when the
CH /), 1.96 /1.20 (m, 74.3H), and 0.88 /0.84 ppm amount of HClO4 was varied and all other
(m, 8.3H, /CH3). 13C NMR: d 179.7 (s, HO /C / variables were held constant are listed in Table 2.
O), 173.6 (s, /CH /O /(C /O) /CH2 /), 130.5 (d, / Vacuum distillation removed any excess fatty acids
C H /C H/, very small signal, only a small amount and by-products providing neat estolide samples.
of alkene present), 74.1 (d, /C H /O /C /O), 34.7 The equivalents of HClO4, percent yields, ENs,
(t), 34.2 (t), 34.0 (t), 32.5 (t), 31.8 (t), 31.8 (t), 31.5 pour and cloud points, viscosity, viscosity index,
(t), 29.8 (t), 29.7 (t), 29.6 (t), 29.6 (t), 29.5 (t), 29.5 and color are reported in Table 2. These new
(t), 29.4 (t), 29.3 (t), 29.2 (t), 29.1 (t), 29.0 (t), 27.2 estolides have an oleic acid backbone with a
S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196 191

terminal saturated fatty acid. Estolides are formed evaluated and general trends, such as the EN, were
from the carbocationic homo-oligomization of noted as with the free acid estolides. The expected
unsaturated fatty acids (Isbell et al., 1994) result- increase in Gardner color (Table 3) caused by the
ing from the addition of a fatty acid carboxyl estolide undergoing an additional treatment with
adding across the olefin. This condensation can acid. This was easily resolved by the decolorization
continue, resulting in oligomeric compounds of the estolides. The Gardner colors improved in
where the average extent of oligomerization is all cases, resulting in colors similar to commer-
defined as the EN (EN /n/1, Fig. 2) (Isbell and cially available material. The estolides esters were
Kleiman, 1994). When saturated fatty acids are also evaluated in terms of structure. The estolides
added to the reaction mixture, the oligomerization ranged from 63 to 49% capped with a saturated
terminates upon addition of the saturated fatty fatty acid (Table 3).
acid to the olefin because the saturate provides no Effects on the other physical properties of these
additional reaction site to further the oligomeriza- estolide 2-ethylhexyl esters were noted as the
tion. Consequently, the estolide is stopped at this amount of HClO4 varied (Table 4). These physical
point from further growth; thus we term the property differences were not readily noticeable in
estolide as being ‘capped’ (Cermak and Isbell, the free acid estolides. As the amount of HClO4
2002). decreased, a dramatic decrease in the pour and
Yield decreases as the amount of HClO4 is cloud points of the lauric /oleic estolide 2-ethyl-
decreased from 0.4 to 0.01 equivalents. At 0.01 hexyl esters occurs (Table 4). The best pour point
equivalents of HClO4, no estolide was detected by improvements were observed between Estolide D,
HPLC. There is a darkening of the free acid (Table 2, PP /24 8C) and the corresponding
estolide as well as a decrease in the EN as the Estolide ester D-2EH (Table 4, PP /41 8C). These
amount of HClO4 is decreased. The physical complex estolides also showed improvements in
properties listed in Table 2 are similar to each viscosities and viscosity indexes.
other in terms of functional fluids. A new series of reactions (Fig. 2) that explored
The alcohol portion of the ester functionality the complex estolides formed when the amount of
was determined by Isbell et al. (2001) to play a HClO4 was varied as the reaction was stirred
significant role in pour point reductions as under vacuum is listed in Table 5. The free
branched chain alcohols dramatically lower the estolides were converted to their corresponding
pour point. Thus, these estolides were converted esters in situ under vacuum, to improve the
into their corresponding 2-ethylhexyl estolide progress of the reaction. One of the by-products
esters by the addition of 0.5 M 2-ethylhexanol/ of the esterification reaction is water. By conduct-
BF3 at 60 8C for 2 /4 h for enhanced pour point ing the reaction under vacuum and at a higher
capability. Samples were vacuum distilled to temperature 60 8C, the water was removed to help
remove any excess 2-ethylhexanol, providing neat drive the reaction to completion. The final pro-
estolide 2-ethylhexyl ester samples. The physical ducts underwent vacuum distillation to remove
properties of the estolide 2-ethylhexyl esters were any excess fatty acids, by-products, and 2-ethyl-

Table 4
Physical properties of lauric /oleic estolide 2-ethylhexyl esters

Estolide estera Pour point (8C) Cloud point (8C) Vis@40 8C (cSt) Vis@100 8C (cSt) Viscosity index

A-2EH /33 /30 89.5 14.5 169


B-2EH /27 /32 69.6 12.6 183
C-2EH /36 /33 52.2 9.8 176
D-2EH /41 /40 52.2 10.0 182

Esterification reactions were run with magnetic stirring and 0.5 M BF3 at 60 8C.
a
Estolide 2-ethylhexyl ester.
192 S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196

Reactions were run for 24 h with overhead stirring under vacuum, 2:1 molar ratio oleic:lauric fatty acids with 0.05 equivalents of perchloric acid, then 1.2 equivalents of
hexyl alcohol, providing neat estolide ester sam-

12/
12/
10/
11/
Gardner
ples.

14
color
When the reaction was conducted under va-
cuum, the yields of the estolide conversions were
increased. In the case of J-2EH (Table 5), the
reaction produced estolide, whereas when the
Acid value

6.11
1.46
0.72
1.60
40.8
(mg/g)

reaction was not under vacuum, no product was


formed. There were other problems in obtaining
completely esterified material with the 0.01 equiva-
lent HClO4. Under normal reaction times, the
Viscosity

product showed a very high acid value for the


174
179
181
184
159
index

distilled product. A longer reaction time may


improve the low conversion rate conditions for
the reaction. J-2EH was changed to allow the
esterification reaction to proceed four times the
Vis@100 8C

10.9
11.2
11.1
11.5
12.4

normal time, which made only a minimal change


(cSt)
Acid-catalyzed condensation reaction with oleic and lauric acid with varying amounts of HClO4 under vacuum

in the overall conversion. Due to low amounts of


Ratio of estolide capped with saturated fatty acids determined by GC (SP-2380, 30 m/0.25 mm i.d.).

acid, this conversion could not be carried out


completely.
The previous data (Tables 2/5) suggested that
Vis@40 8C

the optimal conditions, in terms of properties,


60.5
61.1
59.8
61.5
76.6
(cSt)

yields, and cost, for the synthesis of a complex


estolide is 0.05 equivalents of HClO4 at 60 8C for
24 h followed by the direct conversion to the
corresponding esters. Based on these results, a
cheap and readily available source of saturated
Cloud point

/19
/37
/35
/1
/2

fatty acids is important. Coconut fatty acids are an


(8C)

excellent source based on cost (Tilton, 2002) and


2-ethylhexanol at 60 8C for 2 h followed by distillation at 180 /200 8C.

fatty acid profile (Table 1). The coconut fatty acids


contain the same fatty acids that Cermak and
Isbell (2002) used to synthesize complex estolides,
Pour point

/12
/24
/36
/33
/9

as individual saturated fatty acids, and have


(8C)

demonstrated their excellent cold temperature


properties. Thus, the mixture of fatty acids might
have an even greater effect on the cold temperature
properties due to the increased disruption of the
72.5
63.2
57.5
50.1
30.0
Capped
(%b)

crystal lattice structure.


Esterification reaction went for 8 h.

The physical properties of new complex esto-


lides (Fig. 3) synthesized by varying the amounts
2.24
2.15
2.03
1.99
1.36
(ENa)

of saturated fatty acids are listed in Table 6. The


GC

material represented in Table 6 has not had any of


the excess 2-ethylhexyl monomer ester removed.
Estolide number.
0.05
0.01
HClO4
Equiv.

0.4
0.2
0.1

This material represented in Table 6 is the most


cost-effective material because there is no addi-
tional costs from distillation or decolorization.
Estolide OC-EH-D is the best performing estolide
Estolide
Table 5

G-2EH
H-2EH

J-2EHc
F-2EH

I-2EH

and compares with the simple oleic estolide in


b
a

terms of its cold temperature properties (Isbell et


S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196 193

Table 6
Acid-catalyzed condensation reaction with oleic and coconut acids with varying amounts of fatty acids

Estolidea Oleic to coconut Pour point Cloud point Vis@40 8C Vis@100 8C Viscosity Gardner
ratio (8C) (8C) (cSt) (cSt) index color

OC-EH-A 1:3 18 23 28.8 6.1 165 11


OC-EH-B 1:1 /21 /13 21.3 5.1 173 10
OC-EH-C 1:2 /18 16 21.3 5.0 175 10
OC-EH-D 2:1 /33 /26 52.2 9.6 169 11
OC-EH-E 3:1 /24 /21 58.8 13.2 232 11

Reactions were run for 24 h with overhead stirring under vacuum, oleic:coconut fatty acids with 0.05 equivalents of perchloric acid,
then 1.2 equivalents of 2-ethylhexanol at 60 8C for 2 h followed by distillation at 90 /110 8C.
a
(Coco /oleic)-(2-ethylhexyl ester)-(sample #).

al., 2001). However, it should have increased too high for cold weather climates (Erhan and
oxidative stability over the homo-oligomeric esto- Asadauskas, 2000). All of the commercial pro-
lides. When larger ratios of coconut to oleic fatty ducts listed had higher cloud points than the
acids are used, cold temperature performance is estolide 2-ethylhexyl esters. A high cloud point
reduced, cloud and pour points increased, because could lead to filter plugging and poor pumpability
of an increase in saturated fatty esters. in cold weather applications. The high cloud point
This potential savings from these cost effective of commercially available oils demonstrates a need
materials were limited by the need of a cold for a better cold weather performing oil.
weather performer. After the removal of the coco The proton NMR for the free acid estolides in
and oleic monomers through a second distillation, Table 2, specifically estolide ester D, shows some
there was a dramatic improvement in the cold key features of a typical estolide. The ester methine
temperature properties (Table 7). The best im- signal at 4.85 ppm is indicative of an estolide
provements were observed with the cloud points linkage. Another distinctive feature is the a-
where improvements of 41 8C were noted. These methylene proton shift (2.32 ppm) adjacent to
results were not surprising because the literature the acid and the a-methylene proton shift (2.25
notes that the cold temperature properties of fatty ppm) adjacent to the ester. Integration of these
acid esters are poor, particularly the saturated signals provides a ratio for the number of ester
fatty acid esters. The OC-EH-DD gave the best bonds to acid functionalities. This ratio of a-ester/
cold temperature properties, although, the yield a-acid can be used as another means to determine
was modest. the EN complementary to the GC method. The
The physical properties of various commercial NMR analysis indicates some presence of alkene
materials were compared with the better perform- in the estolide by the appearance of an alkene
ing estolide 2-ethylhexyl esters (Table 8). The two signal at 5.37 /5.33 ppm. The alkene signal in-
estolide 2-ethylhexyl esters selected for comparison dicates that some of the estolide is capped with
with the commercial products were based on price unsaturated material, i.e. oleic acid. The alkene
and best cold weather performance (Tables 6 and signal in the proton NMR supports the IV
7, OC-EH-D and OC-EH-DD). These estolide 2- determined by GC, as the intensity of the NMR
ethylhexyl esters were completely unformulated, signals is comparable to the reported IV.
unlike the commercial products, which contained The carbon NMR spectrum contains the ex-
up to 40% additives designed to improve cold pected estolide signals. Two different carbonyl
temperature properties. All the commercial pro- signals are present at 179.7 ppm (acid) and 173.6
ducts in Table 8 have cold weather-functional pour ppm (estolide ester). The other distinctive signal is
points except the soy-based oil. Soy-based pro- the methine carbon at 74.1 ppm, which is common
ducts have been demonstrated to have pour points to estolides. These major peaks in the carbon
194 S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196

Oleic to coconut ratio GC (ENb) GC iodine value Capped (%) Pour point (8C) Cloud point (8C) Vis@40 8C (cSt) Viscosity index Gardner color
NMR are also confirmed by a DEPT experiment.
The alkene carbons are only slightly noticeable, as
these signals are about the same as the signal-to-
17 noise ratio.
12
13
12
12
The proton NMR for the estolide esters in Table
7, specifically estolide ester OC-EH-DD, gives
some predictable signal changes in the proton
NMR as compared with the free acid estolides.
138
175
164
170
232

The a-carbonyl methylene protons have similar


shifts, resulting in a multiplet from 2.29 /2.24 ppm.
As before, the alkene signal is noticeable at 5.37 /
5.34 ppm, confirming the IV determined by GC.
The carbon NMR signals are indicative of the
149.5
58.4
61.1
92.8
86.3

estolide 2-ethylhexyl ester and confirmed by a


DEPT experiment.
Estolides D and OC-EH-DD (Tables 2 and 7)
were saponified in 0.5 M KOH/MeOH then
esterified with 1 M H2SO4/MeOH to give the
/18
/25
/22
/33
/32

corresponding hydroxy and non-hydroxylated


fatty esters. The isolated mixtures of fatty acid
esters were then silylated and analyzed by GC /MS
(Erhan and Kleiman, 1992). The main mass
spectral features were m/e 371 (M/ /15, 1.2%),
/21
/24
/27
/33
/33

73 (TMS/, 100%), and a Gaussian fragment


representing cleavage at the C /C bond adjacent
to the silyloxy positions (masses 173 /315). The
Physical properties of coco /oleic estolide 2-ethylhexyl esters after distillation

fragments and abundances were very similar to


82.0
49.5
58.0
35.6
40.9

previously reported complex estolide data (Cer-


Samples were distilled until all the monomer was removed by LC.

mak and Isbell, 2001b). The estolide position was


distributed from positions 5 /13 with the original
D9 and 10 positions having the largest abundances
(Coco /oleic) /(2-ethylhexyl ester) /(sample # distilled).

in the mass spectrum, which also was very similar


7.33
7.33
5.81

to complex estolide data of Cermak and Isbell


12.8
14.2

(2001b). This method demonstrated that during


the estolide reaction a migration of the double
bond occurs.
1.49
1.91
1.46
1.94
1.96

4. Conclusions

Estolides have been synthesized with low


amounts of HClO4 to produce lauric/oleic esto-
Estolide number.

lides that have great cold temperature properties.


1:3
1:1
1:2
2:1
3:1

The esterification was incorporated into an in situ


second step, and thus, the estolide and estolide
OC-EH-DD
OC-EH-AD

OC-EH-CD
OC-EH-BD

OC-EH-ED

ester reactions have been modified to produce


Estolidea
Table 7

functional fluids of different grades at a very


b
a

reasonable cost.
S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196 195

Fig. 3. Oleic estolide 2-ethylhexyl ester and coconut /oleic estolide 2-ethylhexyl ester.

Table 8
Comparison of low temperature properties and viscosity index of coconut /oleic estolide 2-ethylhexyl esters to that of commercial
lubricants

Lubricanta Pour point (8C) Cloud point (8C) Vis@40 8C (cSt) Viscosity index

Commercial petroleum oil /27 2 66.0 152


Commercial synthetic oil /21 /10 60.5 174
Commercial soy-based oil /18 1 49.6 220
Commercial hydraulic fluid /33 1 56.6 146
OC-EH-Db (Table 6) /33 /26 55.2 162
OC-EH-DDb (Table 7) /33 /33 92.8 170
a
Commercial, fully-formulated material from local vendors.
b
Unformulated.

Estolides from oleic and coconut fatty acids and these estolides. Thus, as the amount of HClO4 is
their 2-ethylhexyl esters have excellent low tem- reduced, the degree of oligomerization decreases.
perature properties. The coco/oleic estolides These new saturated, coconut fatty acid-capped,
synthesized at 60 8C and a 2:1 oleic to coconut oleic estolides and corresponding 2-ethylhexyl
fatty acids ratio produced materials that had esters have outperformed commercial products in
especially good low temperature properties with cold temperature properties even though the
pour point of /33 8C. The reaction produced a estolides remained unformulated.
functional fluid with great properties at a reason-
able cost. This material can be further refined to
produce a high-value estolide with even superior
cold temperature properties. Acknowledgements
Viscosity is also controlled by the amount of
oligomerization, with higher oligomerization pro- Jason E. Adkins, Kendra B. Brandon, Dawn N.
viding higher viscosity. Many applications are Symonds, and Melissa L. Winchell assisted in the
available for materials with the viscosity range of synthesis of these estolides and the derivation of
196 S.C. Cermak, T.A. Isbell / Industrial Crops and Products 18 (2003) 183 /196

the estolides for further analysis. David Weisleder Erhan, S.M., Kleiman, R., 1992. Meadowfoam and rapeseed oil
provided the NMR spectra. as accelerators in factice production. Inform 3, 482.
Erhan, S.Z., Asadauskas, S., 2000. Lubricant base stock from
vegetable oils. Ind. Crops Prod. 11, 277 /282.
Firestone, D. (Ed.), 1994a. Official and Tentative Methods of
References the American Oil Chemists’ Society, fourth ed. Acid Value,
AOCS Official Method Te 2a-64. AOCS, Champaign, IL.
Akoh, C.C., 1994. Oxidative stability of fat substitutes and Firestone, D. (Ed.), 1994b. Official and Tentative Methods of
vegetable oils by the oxidative stability index method. J. the American Oil Chemists’ Society, fourth ed. Color
Am. Oil Chem. Soc. 71, 211 /216. Gardner 1963 (Glass Standards), AOCS Official Method
American Society for Testing Materials, 1996. Standard Test Td 1a-64. AOCS, Champaign, IL.
Method for Pour Point of Petroleum Products, ASTM (D Firestone, D. (Ed.), 1994c. Official and Tentative Methods of
97 /96a). ASTM, West Conshohocken, PA, pp. 1 /8. the American Oil Chemists’ Society, fourth ed. Calculated
American Society for Testing Materials, 1997. Standard Test Iodine Value, AOCS Official Method Cd 1c-85. AOCS,
Method for Kinematic Viscosity of Transparent and Champaign, IL.
Opaque Liquids (the Calculation of Dynamic Viscosity), Isbell, T.A., Kleiman, R., 1994. Characterization of estolides
ASTM (D 445 /97). ASTM, West Conshohocken, PA, pp. produced from the acid-catalyzed condensation of oleic
1 /9. acid. J. Am. Oil Chem. Soc. 71, 379 /383.
American Society for Testing Materials, 1998. Standard Isbell, T.A., Kleiman, R., 1996. Mineral acid-catalyzed con-
Practice for Calculating Viscosity Index from Kinematic densation of meadowfoam fatty acids into estolides. J. Am.
Viscosity at 40 and 100 8C, ASTM (D 2270 /93). ASTM,
Oil Chem. Soc. 73, 1097 /1107.
West Conshohocken, PA, pp. 1 /6.
Isbell, T.A., Kleiman, R., Plattner, B.A., 1994. Acid-catalyzed
American Society for Testing Materials, 1999. Standard Test
condensation of oleic acid into estolides and polyestolides.
Method for Cloud Point of Petroleum Products, ASTM (D
J. Am. Oil Chem. Soc. 71, 169 /174.
2500 /99). ASTM, West Conshohocken, PA, pp. 1 /3.
Isbell, T.A., Edgcomb, M.R., Lowery, B.A., 2001. Physical
Asadauskas, S., Erhan, S.Z., 1999. Depression of pour points of
properties of estolides and their ester derivatives. Ind. Crops
vegetable oils by blending with diluents used for biodegrad-
Prod. 13, 11 /20.
able lubricants. J. Am. Oil Chem. Soc. 76, 313 /316.
Legrand, J., Durr, K., 1998. High-performing lubricants based
Becker, R., Knorr, A., 1996. An evaluation of antioxidants for
vegetable oils at elevated temperatures. Lubr. Sci. 8, 95 / on renewable resources. J. Agro. Food Ind. Hi-tech. 9, 16 /
117. 18.
Canter, N., 2001. It isn’t easy being green. Lubricants World Mang, T., 1994. Biodegradable lubricants and their future
September, 16 /21. markets. Lipid Technol. November/December, 139 /143.
Cermak, S.C., Isbell, T.A., 2001. Synthesis and physical Mang, T., 1997. Lubricants. In: Gunstone, F.D., Padley, F.B.
properties of cuphea oleic estolides. The Association for (Eds.), Lipids Technologies and Applications. Marcel
the Advancement of Industrial Crops Meeting, November Decker, New York, pp. 737 /758.
10 /13, Atlanta, GA, p. 34. Tilton, H. (Ed.), 2002. Late Spot Prices. Chemical Market
Cermak, S.C., Isbell, T.A., 2001b. Synthesis of estolides from Reporter, October, pp. 14, 15.
oleic and saturated fatty acids. J. Am. Oil Chem. Soc. 78, Zehler, G.R., 2001. Performance tiering of biodegradable
557 /565. hydraulic fluids. Lubricants World September, 22 /26.
Cermak, S.C., Isbell, T.A., 2002. Physical properties of Zoleski, B.H., Gaetani, F.J., 1984. Low Foaming Railroad
saturated estolides and their 2-ethylhexyl esters. Ind. Crops Diesel Engine Lubricating Oil Composition. US Patent
Prod. 16, 119 /127. 4,428,850.

You might also like