Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

This article was downloaded by: 10.3.98.

104
On: 22 Jun 2021
Access details: subscription number
Publisher: CRC Press
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: 5 Howick Place, London SW1P 1WG, UK

Synthetics, Mineral Oils, and Bio-Based Lubricants


Chemistry and Technology, Third Edition
Leslie R. Rudnick

Esters

Publication details
https://www.routledgehandbooks.com/doi/10.1201/9781315158150-3
Stephen Boyde
Published online on: 13 Feb 2020

How to cite :- Stephen Boyde. 13 Feb 2020, Esters from: Synthetics, Mineral Oils, and Bio-Based
Lubricants, Chemistry and Technology, Third Edition CRC Press
Accessed on: 22 Jun 2021
https://www.routledgehandbooks.com/doi/10.1201/9781315158150-3

PLEASE SCROLL DOWN FOR DOCUMENT

Full terms and conditions of use: https://www.routledgehandbooks.com/legal-notices/terms

This Document PDF may be used for research, teaching and private study purposes. Any substantial or systematic reproductions,
re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents will be complete or
accurate or up to date. The publisher shall not be liable for an loss, actions, claims, proceedings, demand or costs or damages
whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.
3 Esters
Stephen Boyde

CONTENTS
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

3.1 Introduction and Historical Development.......................................................................................................................... 46


3.2 Chemistry........................................................................................................................................................................... 47
3.2.1 Ester Structure........................................................................................................................................................ 47
3.2.2 Raw Materials......................................................................................................................................................... 47
3.2.2.1 General..................................................................................................................................................... 47
3.2.2.2 Renewability............................................................................................................................................ 48
3.2.2.3 Acids........................................................................................................................................................ 48
3.2.2.4 Alcohols................................................................................................................................................... 49
3.2.3 Ester Types.............................................................................................................................................................. 50
3.2.3.1 Monoesters............................................................................................................................................... 50
3.2.3.2 Diesters.................................................................................................................................................... 50
3.2.3.3 Neopentyl Polyol Esters........................................................................................................................... 51
3.2.3.4 Complex Esters........................................................................................................................................ 52
3.2.3.5 Other Esters............................................................................................................................................. 53
3.2.4 Synthesis and Production........................................................................................................................................ 53
3.2.4.1 Manufacturing Process............................................................................................................................ 53
3.2.4.2 Specifications........................................................................................................................................... 54
3.3 Properties and Performance Characteristics...................................................................................................................... 54
3.3.1 Structure–Property Relationships.......................................................................................................................... 54
3.3.1.1 Viscosity.................................................................................................................................................. 55
3.3.1.2 Volatility.................................................................................................................................................. 55
3.3.1.3 Flow Properties (Viscosity Index and Pour Point).................................................................................. 56
3.3.1.4 Lubricity................................................................................................................................................... 56
3.3.1.5 Polarity and Material Compatibility........................................................................................................ 58
3.3.2 Stability................................................................................................................................................................... 59
3.3.2.1 Thermal and Oxidative Stability............................................................................................................. 59
3.3.2.2 Hydrolytic Stability.................................................................................................................................. 61
3.3.2.3 Shear Stability.......................................................................................................................................... 62
3.3.3 Safety, Health, and Environmental Properties....................................................................................................... 62
3.3.3.1 Biodegradability....................................................................................................................................... 62
3.3.3.2 Toxicity and Ecotoxicity.......................................................................................................................... 63
3.3.3.3 Recycling and Reuse................................................................................................................................ 63
3.3.3.4 Renewability............................................................................................................................................ 63
3.4 Application Areas............................................................................................................................................................... 63
3.4.1 Engine Oils............................................................................................................................................................. 63
3.4.2 Automotive Gear..................................................................................................................................................... 64
3.4.3 Two Stroke.............................................................................................................................................................. 65
3.4.4 Aviation Turbine Oils............................................................................................................................................. 65
3.4.5 Hydraulic Fluids..................................................................................................................................................... 66
3.4.5.1 Biodegradable Hydraulic Fluids.............................................................................................................. 66
3.4.5.2 Fire Resistant........................................................................................................................................... 67
3.4.6 Air Compressor...................................................................................................................................................... 68
3.4.7 Refrigeration Compressor....................................................................................................................................... 68
3.4.8 High-Temperature Chain Oils................................................................................................................................ 69
3.4.9 Metalworking Fluids............................................................................................................................................... 69
3.4.10 Greases................................................................................................................................................................... 70
3.4.11 Drilling Muds......................................................................................................................................................... 70
3.4.12 Transformer/Dielectric Fluids................................................................................................................................ 70

45
46 Synthetics, Mineral Oils, and Bio-Based Lubricants

3.5 Marketing and Economics.................................................................................................................................................. 71


3.5.1 Manufacturers......................................................................................................................................................... 71
3.5.2 Markets................................................................................................................................................................... 72
3.6 Outlook............................................................................................................................................................................... 72
Acknowledgments........................................................................................................................................................................ 73
References.................................................................................................................................................................................... 73

3.1 INTRODUCTION AND range of synthetic basefluid (SBF) chemistries were investi-


HISTORICAL DEVELOPMENT gated [7]. Additional impetus was provided by the German
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

experience in Russia during the winter of 1941–1942, where


The ester functional group provides a versatile means of assem- mineral oil-based lubricants gelled in the extreme cold,
bling complex organic structures from simple acid and alcohol immobilizing motorized transport. Diester basefluids, as well
building blocks. Ester products are ubiquitous in both synthetic as synthetic hydrocarbons based on Fischer–Tropsch chem-
and natural product chemistry. This chapter focuses on synthetic istry, were found to have much better low-temperature flow
organic ester basefluids, that is, viscous, water-immiscible liq- properties, and these became the main focus of German syn-
uids, containing only C, H, and O, manufactured by esterifica- thetic lubricant technology development in the latter years of
tion processes, and used as major components of formulations the war [8].
intended to reduce friction and wear in machinery. Approximately 3500 different ester structures were synthe-
Other chapters in this book cover inorganic esters contain- sized and characterized during this program. Diesters based
ing phosphorus (Neutral Phosphate Esters), carbonate esters on adipic or methyladipic acids and C8–C12 oxoalcohols were
(Dialkyl Carbonates), naturally occurring esters produced selected for scale-up and were manufactured by I.G. Farben
by physical processing of plant or animal tissues (Natural at production rates of up to 100 barrels per day (i.e., an annu-
Oils as Lubricants), and materials manufactured by addition alized rate of approximately 6.5 ktes pa). The products were
polymerization of olefinic ester monomers (Polymer Esters) mainly used as minor components (10%–25%) of blends with
or unsaturated acids (Estolides). The use of amphiphilic esters mineral oils or synthetic hydrocarbons, in high-performance
(polyalkylene glycol esters and partial esters) at lower concen- applications including aircraft engine and hydraulic oils and
trations as surfactants in aqueous emulsion formulations is also low-temperature railroad oils [8].
covered (Metalworking Fluids) and their use as lubricity addi- Prior to the war, workers at the Deutsche Fettsäure-Werke
tives in nonaqueous formulations is reviewed elsewhere [1]. GmbH in Witten (DFW) had also synthesized esters by reac-
The use of esters as lubricants is as old as human technol- tion of the neopentyl polyols trimethylolethane or pentae-
ogy because natural fats and oils, which are predominantly rythritol (PE) with short chain (C6 –C12) natural fatty acids
triesters of glycerol with linear fatty acids, have been used and shown that these polyol esters had superior properties to
to lubricate sliding contacts from the chariots of ancient the castor oil triglycerides then used in aviation lubricants.
Egypt [2] to the steam engines of the Industrial Revolution Although not developed as extensively as diester fluids by the
[3]. However, modern synthetic ester basefluids date from the German war effort, DFW recognized the potential of these
mid-twentieth century, when the application requirements of polyol esters, both as high-stability basefluids in their own
the rapidly evolving military technology during World War II right and as blend components to optimize the properties of
began to exploit the potential of petrochemical raw materials synthetic hydrocarbon oils [8].
as building blocks for optimized lubricants. A more limited research program had also been under way
Some of the groundwork had been laid by the requirement in the United States led by W. A. Zisman at the U.S. Naval
for low volatility, low melting fluids for use as plasticizers for Research Laboratory, focusing on diester chemistries [5,6,9].
flexible polymers, particularly with the adoption of polyvinyl- The subsequent Cold War ensured continued military inter-
chloride (PVC) for electrical insulation. A wide range of mate- est and an active research program into synthetic ester base-
rials, including esters based on aromatic and aliphatic diacids, fluid and additive technology was maintained, building on the
were screened and commercialized as PVC plasticizers dur- results of the German program [10,11].
ing the 1920s and 1930s [4] and some had already found use The further development of ester lubricants was initially
in specialized lubricant applications, including vacuum pump closely linked to that of the aviation gas turbine, which could
and instrument oils by the late 1930s [5]. not be adequately lubricated with mineral oil due to extreme
During the 1930s, limited availability of suitable mineral temperature requirements. Although many alternative chem-
oil basestocks in Germany directed attention to large-scale istries were evaluated and some found limited application (see
production of synthetic lubricants from chemical feedstocks. Trends toward Synthetic Fluids and Lubricants in Aerospace),
The increasing performance demands of aircraft engine tech- first diester [6] and then polyol ester [11,12] basefluids became
nology also led to a need for lubricant basefluids with sta- the standard for military applications. The market grew fur-
bility and operating temperature ranges, which could not be ther as gas turbine engines were introduced in civil aviation,
achieved with the mineral oils of the time [5,6]. where requirements were originally set based on published
An extensive research program was instigated, led by Dr. military specifications and similar basefluid and formulation
Hermann Zorn at the Leuna works of I.G. Farben, and a wide technology was therefore required.
Esters 47

With their capability proven against the exacting standards


of the aviation industry, the use of synthetic ester basefluids has
now become widespread in many other lubricant applications.
In automotive engine and gear oils, the major use of ester
basefluids has been as components of blends with synthetic
hydrocarbons (see Polyalphaolefins; Gas to Liquids), or highly
refined (API Group III) mineral oils (see Chemically Modified
Mineral Oils), where the ester component is used to modify
the solvency, seal swell, and lubricity characteristics of the
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

hydrocarbon. Originally developed for use in extreme low- FIGURE 3.1  The structure and formation of the ester group.
temperature environments [5], such mixed SBFs have become
increasingly widely used in top tier conventional passenger the molecular s­ tructure than is currently possible with hydro-
car and heavy duty diesel formulations, driven by ever more carbon c­ hemistry [13].
stringent targets for fuel economy and extended fill since the The raw materials used for ester lubricants generally carry
1970s (see Automotive Crankcase Oils). relatively large hydrocarbon substituent groups and the prop-
In some cases, requirements for synthetic ester basefluid erties of ester lubricant basefluids are largely influenced by
have been created directly by environmental legislation. For the number and structure of these hydrocarbon groups [10,14].
example, the Montreal Convention of 1987 mandated the However, the presence of the ester group modifies the prop-
phase out of chlorofluorocarbon (CFC) and hydrochlorofluo- erties relative to those of an analogous hydrocarbon. In par-
rocarbon (HCFC) refrigerant gases. The replacement hydro- ticular, the ester group carries a permanent dipole moment of
fluorocarbon (HFC) gases are incompatible with mineral oil approximately 1.7 Debye units due to unsymmetrical distribu-
basefluids due to their higher polarity. Neopentyl ester base- tion of electron density in the carbon–oxygen bonds, and ester
fluids were found to be suitable and the demand for polyol basefluids are therefore more polar than hydrocarbons. Ester
ester HFC-compatible refrigeration compressor lubricants has polarity varies with the number and spacing of ester groups
grown strongly since the early 1990s as the gas changes have in the structure.
been implemented across the various sectors of the refrigera- Structure–property relationships for ester basefluids are
tion industry (see Refrigeration Lubes). discussed in more detail later.
Another major growth area for synthetic esters over the
last two decades has been in environmentally acceptable
3.2.2 Raw Materials
hydraulic fluids. The use of biodegradable fluids is increas-
ingly required by legislation where mobile hydraulic equip- 3.2.2.1 General
ment is used in environmentally sensitive areas. Because the Because the esterification process does not modify the
ester group provides a site for initial microbial attack, ester hydrocarbon backbone of the acid and alcohol raw materi-
basefluids can be designed to have acceptable biodegrad- als, the ability to construct a particular ester structure obvi-
ability while having better oxidative stability and low-tem- ously depends on availability of the required building blocks.
perature flow characteristics than straight vegetable oils (see Although almost every conceivable acid or alcohol has been
Environmentally Friendly Hydraulic Fluids). or can be prepared in the laboratory, only a limited range
Synthetic ester basefluids are now found across the range are available commercially at suitable scale and cost to be of
of lubricant applications, including automotive and marine interest as raw materials for ester lubricants.
engine oils, compressor oils, hydraulic fluids, gear oils, and Synthetic lubricant basefluids represent a small fraction of
grease formulations, wherever the ability to tailor properties the overall chemical industry. For raw materials to be avail-
by molecular design, or the inherent properties of the ester able at commercially viable costing it is generally necessary
group, offer a sufficient performance advantage to justify the that they be manufactured at very large scale for primary use
additional cost over mineral oils. in larger volume market sectors such as polymers, coatings,
or surfactants.
3.2 CHEMISTRY The chemical industry has undergone substantial restruc-
turing and globalization since the original development of
3.2.1 Ester Structure ester basefluids. While many new products have been intro-
The structure of the ester functional group is shown in duced, overall the diversity of the viable raw material range
Figure 3.1, which also illustrates how it is formed by chemi- has reduced over time as the larger chemical sectors have con-
cal reaction of carboxylic acid and alcohol raw materials, solidated their usage around preferred materials and smaller
with removal of water. The relative simplicity of the reac- scale processes have become uneconomic. For example, nei-
tion chemistry, and the ready availability of a wide range of ther the methyladipic acids nor the “Leuna alcohols” which
monofunctional and polyfunctional acid and alcohol building were the major raw materials used in the original German
blocks, mean that complex molecules can be constructed cost- 1940s diester lubricants are readily commercially available in
effectively using ester linking groups with better control of the twenty-first century.
48 Synthetics, Mineral Oils, and Bio-Based Lubricants

The following sections summarize the major raw materials deodorization of food-grade vegetable oils, and the “top
currently used in ester lubricant basefluids. and bottom” fractions from lauric oil fatty acids, which are
unsuitable for surfactant production. The composition and
3.2.2.2 Renewability availability of such sources varies, so extensive process-
Raw materials used for esters can be based either on petro- ing is required to produce derivatives of consistent quality.
chemical or renewable sources. Increasing environmental Furthermore, oleochemical processing also generally gives
awareness particularly in Europe and North America has cre- rise to coproducts for which economic applications must
ated a demand for fluids with reduced environmental impacts also be found. For example, separation of oleic acid from
[15,16]. Over the same period, overcapacity and reduction tallow for use in ester lubricants coproduces glycerol and
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

in subsidies for food production have prompted efforts by stearic acid. These considerations will limit the available
agricultural producer groups to seek nonfood applications range of renewable-derived raw materials for the foresee-
for their crops. Lubricant applications appeared an obvious able future.
industrial outlet for natural fats and oils, and consequently
much effort has been directed to investigating and promoting
3.2.2.3 Acids
opportunities for their use [17,18].
This work has largely resulted in the rediscovery of the 3.2.2.3.1 Monoacids
performance limitations of natural oils, which had led to The carbon number and functionality of acid and alcohol raw
their substitution by mineral basefluids for most lubricant materials are of key importance because they will determine
applications during the early twentieth century; poor oxida- the molecular weight of the product and the relative concen-
tive stability, high pour points, and limited viscosity range tration of ester groups and therefore the viscosity and polarity
[19,20]. Furthermore, recent years have seen a recovery in of the ester product.
farm gate prices for traditional food uses, as well as develop- Linear saturated carboxylic acids used in lubricant base-
ment of large-scale capacity for conversion of both oilseed fluids generally have carbon numbers in the range C5–C12.
and grain crops into fuels. As the potential market for fuels Shorter chain acids are not used because of their relatively
is approximately 10 times greater than that for organic mate- high corrosivity and offensive odor if released by degrada-
rials, it will be readily able to absorb any future overcapac- tion processes during use. Longer linear alkyl chains result
ity. Accordingly, producer interest in promoting renewable in products with unacceptable low-temperature flow proper-
raw materials for materials applications such as lubricants ties. Disruption of the linear alkyl chains by unsaturation or
has diminished, although the momentum imparted to gov- branching allows the use of higher carbon numbers.
ernment and academic research programs may take longer Available monocarboxylic acids include the linear natu-
to dissipate. ral fatty acids, which contain even numbers of carbon atoms
However, it has been sufficiently influential that some stan- in the range C8–C22. These may be used either as mixtures
dards for environmentally acceptable fluids include a require- in the naturally occurring ratio (e.g., tallow fatty acids, coco
ment that a proportion of the carbon content of the fluid fatty acids) or fractionated into narrower cuts or single chain
should be derived from renewable sources, as well as stipulat- lengths. The most widely used for manufacture of lubricant
ing minimum standards for biodegradability, ecotoxicity, and basefluids are C8–C10 and oleic acid (monounsaturated C18
application performance [21,22]. acid). Although petrochemical routes have been used, these
Fortunately, although natural triglycerides are unsuitable C-even linear acids are currently only commercially avail-
for most high-performance lubricant applications, they can able from renewable sources [24]. Because these acids arise
serve as sources of renewable raw materials for synthetic as components of a natural raw material in fixed ratio, their
esters through hydrolysis to the constituent fatty acids, which pricing and availability depends on the supply–demand bal-
can then be purified and derivatized as required. A significant ance for each of the components. Pricing for the C8–C10 frac-
proportion of acid and alcohol raw materials used for syn- tion has been highly volatile over the last years, which has led
thetic ester basefluids are of renewable origin, and this is indi- to some substitution by linear C9 and some reformulation to
cated in the following sections. alternative technologies.
Synthetic esters therefore represent an overlap between the Shorter chain linear C-odd (C5–C9) monocarboxylic acids
synthetic and bio-based lubricant product categories, since are manufactured from petrochemical raw materials by car-
they can be designed to be readily biodegradable and to incor- bonylation/oxidation. These processes give rise to smaller
porate renewable raw materials. However, they are not nor- quantities of branched isomers, which may remain in the
mally derived exclusively from renewables and indeed many product as minor components or be separated for sale as
of their useful properties depend on the use of raw materials coproducts. For example, conversion of C4 olefin 1-butene to
that are (currently) derived from petrochemical sources [23]. C5 acid gives the linear C5 valeric acid as major product, with
In general, cost-effective manufacture of such renew- minor amounts of branched isovaleric isomers as coproducts.
able oleochemical raw materials requires a low-cost source, Linear C7 and C9 acids are also available from renewable raw
which is not suitable for use in food or feed. Typical sources materials. The renewable routes do not generate branched
include tallow from slaughterhouse waste, acid oils from coproducts but do involve side reactions leading to linear
acids of different carbon numbers [24,25]. In some cases, this
Esters 49

different impurity profile between petrochemical and renew- limited by the tendency for longer chains to form waxes at low
able materials can impact the properties of derivative esters. temperatures, but higher carbon numbers can be used where
Fully branched C8 and C9 short chain acids are available carbon chains are branched. For linear alcohols, the practical
from petrochemical processing, in each case as single isomers maximum is approximately C11.
(2-ethylhexanoic and 3,5,5-trimethylnonanoic acids, respec- Pure single component short chain monoalcohols in the
tively) [25]. Mixed isomer branched C8 acid (isooctanoic range C1–C4 are readily available. Mid-chain alcohols in the
acid), which was used as a raw material for biodegradable range C8–C24 are produced in large volume for use in surfac-
polyol esters, is no longer commercially available following tants and plasticizers. The surfactants market typically prefers
the closure of Exxon’s manufacturing plant in Harnes, France, predominantly linear alcohols with carbon numbers C12–C16.
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

during 2004. C10 2-propylheptanoic acid [108] and higher car- The plasticizer demand is mainly for C8–C13 branched alco-
bon number α-branched “Guerbet” acids are available in prin- hols, although there is an increasing trend toward more linear
ciple and can be used similarly to 2-ethylhexanoic, but have structures [28].
not yet found high-volume application. Fully linear C-even surfactant alcohols are available from
Acids having the α,α-dibranched structure (“neo acids”) are both petrochemical and renewable raw materials. Both routes
available but generally not suitable for use as ester raw materi- generate a carbon number distribution from C8 upward, which
als because steric hindrance results in very low reactivity. can be fractionated into tighter cuts or single chain products;
Longer chain branched C18 isostearic acid is available from the C8–C10 fraction is preferred for lubricants. These alcohols
renewable raw materials as coproduct from the manufacture can be dimerized to monobranched Guerbet alcohols, with
of dimer and trimer acids [24]. Degree of branching and level useful carbon numbers in the range C16 –C20 [28]. As with the
of residual unsaturation in isostearic acids depend on the analogous acid, pricing for the C8–C10 alcohol fraction has
source, and products from different suppliers are generally been volatile over recent years leading to some substitution
not interchangeable. by linear C9, which had previously only been available as a
component of C9–C11 mixed linear/branched surfactant alco-
3.2.2.3.2 Di- and Polyfunctional Acids hol grades.
Linear aliphatic diacids in the range C4 –C12 are produced in C-odd surfactant alcohols are typically sold to a carbon
large volume for manufacture of polyamides and are suit- number range rather than as single chain length, and are nor-
able for production of diesters. The most widely available mally predominantly linear but with significant branched
are C6 adipic, C9 azelaic, C10 sebacic, and C12 dodecanedioic. content.
Currently, C6 and C12 are petrochemical-derived, whereas the Single isomer, fully branched monoalcohols include
C9 and C10 diacids are renewable-derived [26]. C8 (2-ethylhexanol), C9 (3,5,5-trimethylhexanol), and C10
C36 dimer and C54 trimer acids are branched mixed iso- (2-propylheptanol). Mixed isomer branched alcohols derived
mer materials manufactured from renewable polyunsaturated from polymerized C3 and/or C4 olefins are also available
fatty acids. Dimer and trimer acids have very complex prod- with carbon numbers ranging from approximately C8–C13.
uct distributions, including some residual unsaturation and Compositions of these mixed isomer materials may vary
formation of both alicyclic and aromatic ring structures. The significantly between different producers. Both C-odd and
composition of dimer and trimer acids and therefore the prop- fully branched monoalcohols are based on petrochemical raw
erties of their derivatives, depends significantly according to materials [28,29].
the raw material origin and degree of processing [24,26].
A range of aromatic diacids and polyacids are also readily 3.2.2.4.2 Di- and Polyfunctional Alcohols
available due to their high volume use in polymers, coatings, The polyols most commonly used in lubricant esters are the
and plasticizers. The most widely used in lubricants are the “neopentyl polyols,” which have primary alcohol methy-
unsymmetrical 1,2-difunctional phthalic and 1,2,4-trifunc- lol groups attached to a quaternary carbon atom, due to the
tional trimellitic acids. Tetrafunctional pyromellitic acid has superior thermal stability of esters based on this structure. All
also been used. All are marketed as the corresponding anhy- of the neopentyl polyols have historically been derived from
drides [27]. Esters of terephthalic acid, particularly bis(2- petrochemical raw materials, but fully or partially bio-based
ethylhexyl) terephthalate (DOTP), have longstanding use in grades are now commercially available.
plasticizer applications where volumes have increased signifi- Neopentyl polyols are available with functionality from
cantly in recent years due to substitution of phthalates [109], two to six. The most widely used are neopentyl glycol (NPG,
but have not yet found significant application as lubricants. difunctional), trimethylolpropane (TMP, trifunctional), and
pentaerythritol (PE, tetrafunctional) [30].
3.2.2.4 Alcohols Hexafunctional dipentaerythritol (DiPE) is a minor
3.2.2.4.1 Monoalcohols coproduct in the production of PE. PE may be sold as the
Lubricant esters are normally derived from primary alco- production mix technical PE (Tech PE), containing approxi-
hols, because esters of secondary or tertiary alcohols have mately 10% DiPE, or purified to monopentaerythritol (Mono
inferior thermal stability. There is no minimum chain length PE), and DiPE. Esters based on Tech PE have higher viscosity
limit for alcohol raw materials. As for acids, the maximum and lower pour point than corresponding Mono PE esters, and
chain length for linear alkyl groups in alcohol raw materials is Tech PE is therefore preferred for some applications.
50 Synthetics, Mineral Oils, and Bio-Based Lubricants

Trifunctional trimethylolethane (TME) was used in the monoalcohol and/or monoacid raw material feeds are used in
original German DFW polyol esters, but TMP has been more manufacture of a monoester, the composition of the resulting
widely used since it has been more available and less costly. monoester mix is identical to that which would be obtained
Esters of TMP have slightly higher viscosity and lower pour if the individual raw materials were reacted separately and
point, though slightly inferior oxidative stability to TME ana- blended after reaction. However, when mixed monofunc-
logues [31]. tional feeds are used in manufacture of a diester or polyol
Many other diols, including ethylene glycol, propylene gly- ester, cross products are formed, in which residues from the
col, and their oligomers, are readily available and have more different monofunctional raw materials are combined in the
limited use as ester raw materials. Higher polymers of ethylene same ester product molecule. Such mixed-feed reaction prod-
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

oxide and/or propylene oxide (see Polyalkylene Glycols) can be ucts therefore have a higher structural diversity than blends
esterified to PEG or PAG esters, but these materials are typi- of the same raw materials esterified separately and blended
cally amphiphilic and used in surfactant, rather than basefluid, post-reaction [34].
applications. Diesters of longer chain glycols such as hexane-
diol with monoacids have been evaluated, but properties were 3.2.3.1 Monoesters
generally inferior to those of analogues assembled using the Monoesters result from reaction of monofunctional acid (e.g.,
corresponding difunctional acid and monoalcohol [32]. oleic, isostearic) with monofunctional alcohol. They contain
Polyols other than neopentyl glycols are used to some only one ester group and only two hydrocarbon residues. They
extent, particularly the bio-based trifunctional glycerol and are therefore the lowest molecular weight and typically low-
hexafunctional sorbitol. (Petrochemical routes to glycerol est viscosity/highest volatility ester product group. Table 3.2
exist, but the present market is essentially all renewable due illustrates how the physical properties of monoesters of an
to coproduction in the manufacture of fatty acid methyl esters acid raw material, in this case oleic acid, can be systemati-
for biodiesel.) Glycerol can be reacted with desirable fatty cally varied by selection of the alcohol raw material.
acid fractions (e.g., short chain C8C10 or monounsaturated Monoleate esters are typically used in metalworking
oleic acid) to give synthetic triglycerides having much bet- applications.
ter stability and low-temperature characteristics than natural
oils, though inferior to TMP analogues. Sorbitol undergoes 3.2.3.2 Diesters
side reactions under typical esterification conditions, which Diesters result from reaction of di- or polyacid with (normally
partially convert the original hexitol to tetrafunctional sorbi- branched) monofunctional alcohol. Because of the wider
tan and difunctional isosorbide, leading to very complex reac- range of diacid structures available, diesters are normally sub-
tion mixtures [33]. Partially esterified sorbitol derivatives are divided into three distinct groups.
widely used as surfactants. Alternatively, sorbitol can be fully Diesters based on short chain (C6 –C12) linear diacids with
converted to isosorbide and esterified to diesters which have plasticizer alcohols (Figure 3.2) are relatively low-viscosity,
been promoted for plasticizer applications and may find some
use in lubricants [110]. Glycerol and sorbitol both contain sec-
ondary alcohol groups and their ester derivatives therefore TABLE 3.2
have relatively low thermal stability, limiting their suitability Typical Physical Properties of Mono-Oleates
for high-temperature applications.
Viscosity at Viscosity at
Alcohol 40°C (cSt) 100°C (cSt) VI Pour Point (°C)
3.2.3 Ester Types Methyl 4.5 1.8 – −12
Isopropyl 5.3 2.0 221 −21
Table 3.1 shows how the different ester basefluid structures
Iso-butyl 6.0 2.2 219 −50
are classified according to the combinations of mono- and
2-Ethyl hexyl 8.0 2.8 238 −35
polyfunctional acid and alcohol raw materials used.
Iso-octyl 9.1 2.9 192 −29
Ester basefluids are frequently based on mixed raw materi-
Decyl 10.2 3.4 246 −3
als with a distribution of chain lengths or isomer structures,
Glycerol 100 10.4 83 0
either because the raw materials are only available as mix- mono-oleate
tures, or because pure components are intentionally blended
prior to reaction to produce a mixed feed. Where mixed

Diesters n = 4 - adipates
O O
n = 7 - azelates
TABLE 3.1 C (CH2)n C n = 8 - sebacates
The Major Ester Basefluid Structural Types R O O R n = 10 - dodecanedioates
Monoacid Di- or Polyacid
R = C7 - C18 linear or branched alkyl groups
Monoalcohol Monoester Diester
Di- or polyol Polyol ester Complex ester FIGURE 3.2  Chemical structure of diesters of short chain linear
diacids.
Esters 51

TABLE 3.3
Typical Physical Properties of Diesters of Short Chain Linear Diacids
Viscosity at Viscosity at Pour Noack at 250°C/1 OECD 301
40°C (cSt) 100°C (cSt) VI Point (°C) h (% loss) Biodegradability
2-Ethyl hexyl
Adipate 8.0 2.4 124 −68 44 90–100
Azelate 10.7 3.0 137 −64 29 75–90
Sebacate 11.8 3.1 126 −60 18 85
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

Dodecanedioate 14.3 3.8 168 −57 11 –


2-Propylheptyl
Adipate 11.4 3.0 122 −41 28
Isodecyl
Adipate 15.2 3.6 121 −62 15 90–100
Azelate 18.1 4.3 151 −65 10 75
Sebacate 20.2 4.8 169 −60 6 80
Dodecanedioate 23.4 5.2 162 −41 4 −
Isotridecyl
Adipate 27.0 5.4 139 −51 5 55–70
Azelate 33.8 6.4 143 −55 4 –
Sebacate 36.7 6.7 141 −52 4 –
Dodecanedioate 40.7 7.6 156 −50 3 25–50

high-polarity materials with typically very good VIs and is more rigid and more polarizable than linear aliphatic chains.
pour points. Examples of commercially available short chain Accordingly, aromatic esters have different solvency charac-
diesters are given in Table 3.3 showing how their properties teristics, higher viscosity, and lower VI than linear aliphatic
can be varied by selection of acid and alcohol carbon number esters of the same alcohol raw materials. (The entire group
and degree of branching. Derivatives of C36 dimer acids have is generally described as diesters, although obviously deriva-
significantly higher viscosity and lower polarity due to the tives of trimellitic anhydride are in fact triesters.) Structures
higher carbon number acid raw material. Because of residual of typical aromatic esters are shown in Figure 3.3. Properties
unsaturation and branching on the diacid backbone, thermal of the aromatic esters are shown in Table 3.5, illustrating how
and oxidative stability is lower than for short chain diesters. viscosity depends both on the alcohol raw material and the
The alcohol carbon number makes proportionally less differ- acid functionality that determines how many alcohol groups
ence to the product molecular weight, so selection of alcohol are included in each ester molecule.
raw material makes less difference to final product molecular
weight, as shown in Table 3.4. The alcohol most widely used 3.2.3.3 Neopentyl Polyol Esters
is 2-ethyl hexanol. Dimer acids from different sources vary Reaction of a polyfunctional alcohol with monofunctional
widely in structure, depending on the degree of unsatura- acid(s) produces a polyol ester (Figure 3.4).
tion of the fatty acid raw material and the process conditions. Physical properties of a range of single-component neopen-
Lubricant dimerate esters are normally based on technical tyl polyol esters are given in Table 3.6, showing that viscos-
grades of dimer acid which contain relatively high levels of ity and other properties can be varied over a very wide range
trifunctional trimer acid and some residual unsaturation. by appropriate selection of the acid raw material(s) and the
Derivatives of aromatic acids or anhydrides have distinct polyol functionality. In practice, the majority of commercial
properties because of the presence of the aromatic ring, which polyol ester fluids are based on mixed acid raw materials, and

TABLE 3.4
Typical Physical Properties of Dimerate Esters
Viscosity at Viscosity at Viscosity Pour Noack at CEC-L-33-A-93%
Alcohol 40°C (cSt) 100°C (cSt) Index Point (°C) 250°C/1hr (% loss) Biodeg.
2-Ethyl hexyl 91.1 12.7 136 –50 0.8 49
2-Ethyl hexyl (C36 92.0 12.7 135 –41 0.2 24
hydrogenated)
C9/C11 94.9 14.0 151 –15 1.1 78
iso-tridecyl 140.0 17.0 132 –27 0.6 30
52 Synthetics, Mineral Oils, and Bio-Based Lubricants

O ester product has an oligomeric polyester backbone with a sta-


tistical distribution of molecular weights. Lubricant basefluid
C complex esters are normally manufactured using one mono-
OR
R = C7 - C18 linear or branched alkyl groups functional raw material of either type in addition to the poly-
OR functional reactants, to act as end capper to convert unreacted
C
acid or hydroxyl functionality.
O Final product average molecular weight is controlled by the
ratio of polyfunctional acid and alcohol reactants. Exclusively
O O O O difunctional reactants give linear backbone structures and
Trimellitate Pyromellitate
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

relatively narrow molecular weight distribution, whereas


C C C C inclusion of tri- or tetrafunctional reactants gives branched
RO OR RO OR
backbones with broader distributions [35].
OR RO OR
C C C Complex ester chemistry therefore represents a route to
products having very high viscosity while retaining good
O O O biodegradability due to the ester links in the backbone.
R = C7 - C18 linear or branched alkyl groups Furthermore, because of their polymeric nature, complex
esters are subject to less stringent new substance notification
FIGURE 3.3  Structures of the aromatic esters. requirements in major regulated territories, including the EU
and the United States. This reduces both the time and cost of
therefore contain a statistical distribution of mixed ester prod- new product introduction and therefore favors innovation in
ucts. The higher structural diversity of mixed polyol esters gen- this area. However, because of the distribution of molecular
erally gives improved low-temperature fluidity, whereas other weights, complex esters typically have higher volatility and
parameters such as viscosity vary in the same way as a blend poorer material compatibility (elastomer compatibility, misci-
of the pure single component esters. Neopentyl polyol esters bility with other basefluid types, and interference with func-
have superior thermal stability compared to diesters. Solvency tional additives) than simple esters of the same viscosity.
characteristics are also somewhat different because the ester Oligomeric polyesters related to complex esters can also be
groups are rigidly oriented at the center of the molecule. produced by condensation homopolymerisation of hydroxy-
acid monomers containing both acid and alcohol groups, e.g.,
3.2.3.4 Complex Esters ricinoleic or 12-hydroxystearic acids derived from castor oil.
The ester types described earlier contain relatively well- In this case, the polyfunctional acid and alcohol groups are
defined molecular structures, with molecular weight range necessarily present in 1:1 ratio and a defined ratio of at least
depending only on the carbon number distribution in the raw one nonvolatile monofunctional endcapper is required to con-
materials. However, if polyfunctional acids and polyfunc- trol molecular weight. For basefluid applications, it is nor-
tional alcohols are reacted together, the resulting complex mally necessary to minimize both unreacted acid and alcohol

TABLE 3.5
Typical Physical Properties of Aromatic Esters
Viscosity at Viscosity at Viscosity Pour Noack 250°C/1 OECD 301 Biodeg-radability
Ester 40°C (cSt) 100°C (cSt) Index Point °C h (% loss) % (28 days)

Bis(2-ethylhexyl) phthalate 26.4 4.2 21 −43 19 82


Di-isononyl phthalate 38.5 5.3 50 −44 12 81
Bis(2-propylheptyl) phthalate 37 5.1 40 −48 11 80–90
Di-isodecyl phthalate 45.5 5.8 49 −47 9 67
Di-isotridecyl phthalate (1) 80.5 8.2 46 −43 3 71
Bis(2-ethylhexyl) terephthalate 29 4.5 87 −45 10 73–78
Tris (octyl, decyl) trimellitate 52 8.1 126 −45(2) 0.4 21–45
Tris (2-ethylhexyl) trimellitate 90 9.7 82 −36 1.6 20
Tri-isodecyl trimellitate 145 13 79 −30 1 2
Tri-isotridecyl trimellitate(1) 305 20.5 76 −9 1.6 −
Tetrakis (2-ethylhexyl) 172 16.3 98 −27 0.4 0–10
pyromellitate

1. Derived from tridecanol based on propylene tetramer.


2. Freezes at ca 0°C on prolonged storage.
Esters 53

Neopentlyglycol ester O 3.2.3.5 Other Esters


CH3 The previous sections describe the main classes of esters
C
C CH2 O R intended for use as lubricant basefluids. In all of these, the
reactive and H-bonding acid and alcohol groups are essen-
CH3 tially fully converted to ester, substituent groups are hydro-
2
carbons, and no other functional groups are present. However,
it will be apparent that the flexibility of esterification chem-
Trimethyolpropane ester O istry also allows assembly of many other chemical structures
with additional functionality and reactivity, some of which
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

C
find application as lubricant and fuel additives
CH3 CH2 C CH2 O R
Significant examples include:

3 Partially esterified polyol esters with residual hydroxyl


functionality used as friction modifiers [1].
Pentaerythritol ester O Complex esters with unreacted acid or alcohol
functionality used as tribo-polymerizing antiwear
­
C
additives [36].
C CH2 O R
Amphiphilic PEG or PAG esters containing polyol seg-
ments based on polar polymers of ethylene oxide
4 and/or propylene oxide used as soluble oils and
Dipentaerythritol ester emulsifiers [37,38].
O

C 3.2.4 Synthesis and Production


CH2 O O R 3.2.4.1 Manufacturing Process
C
There are typically four distinct stages to manufacture of ester
R
lubricant basefluids:
O CH2 O C CH2 O O
1. Esterification
C
2. Stripping of excess reactants
CH2 O R
3. Neutralization
2 4. Filtration
FIGURE 3.4  Chemical structure of polyol esters. The basic chemical reaction for the esterification process is
shown in Figure 3.1. Esterification is an equilibrium process,
groups, which requires use of both alcohol and acid endcap- and formation of the ester group is carried out by mixing the
pers. The resulting product distribution therefore contains acid and alcohol raw material while removing the water of
some monoester derived from reaction between endcappers. reaction in order to drive the equilibrium to the right-hand
Complex esters based on hydroxyacids have found limited use side. Ester lubricant basefluids are normally manufactured
as basefluids but have established application as polymeric batchwise, using a stirred reaction vessel equipped with heat-
surfactant hydrophobes [111]. An alternative route based on ing, cooling, and overheads capable of condensing and sep-
unsaturated fatty acids, using addition of carboxylic acids to arating water of reaction from volatile organic reactants by
olefins instead of esterification, has recently been developed fractionation or decantation [13].
[112] and is described in a separate chapter (Estolides). The To facilitate removal of water and accelerate reaction,
estolide ester group forms part of the polymer backbone, so esterification processes are normally carried out at a tem-
these products are more closely related to complex esters perature of at least 100°C, and since the products have good
than to Polymeric Esters where the ester group is part of a thermal stability, higher temperatures up to about 250°C are
pendant chain. The backbone chain length between ester commonly used. Esterification catalysts may be used; these
groups is more variable for estolides manufactured by addi- include strong mineral acids such as sulfuric or p-toluene
tion to olefin than for poly 12-hydroxystearates and the ole- sulfonic acids for use at lower temperatures and Lewis acid
fin-addition estolides therefore have better low-temperature catalysts such as tin and titanium alcoholates at higher tem-
fluidity. Whichever manufacturing route is used, the estolide peratures. Azeotroping agents may also be used to assist in
ester group is equivalent to an ester of a secondary hydroxyl removal of water [13].
group and therefore has inherently lower thermal stability, but For diester products, the monoalcohol reactants normally
higher hydrolytic stability, than the primary hydroxyl ester have significant volatility, and a 10%–20% molar excess of
groups present in most other ester SBFs. monoalcohol is frequently used to drive the equilibrium.
54 Synthetics, Mineral Oils, and Bio-Based Lubricants

TABLE 3.6
Typical Physical Properties of Polyol Esters
Alcohol–Acid Viscosity at Viscosity at VI Pour Noack at 250°C/1 OECD 301 Biodegradability %
40°C (cSt) 100°C (cSt) Point (°C) h (% loss) (28 days)
Neopentyl Glycol
nC7 5.6 1.9 – −64 – 89
nC9 8.6 2.6 145 −55 31 76
nC8/nC10 8.4 2.5 129 −33 32 90
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

Oleic 30 7 207 −24 1 90–100


2-Ethyl hexanoic 7 2 76 −54 76 10–20
3,5,5-Trimethylhexanoic 12.8 3.1 100 −45 43 –
Trimethylolpropane
nC7 13.9 3.4 120 −60 12 85
nC9 21 4.6 139 −51 2 76
nC8/nC10 20.4 4.5 137 −43 3 80–90
Oleic 46.8 9.4 190 −39 2 80–90
2-Ethyl hexanoic 24.8 4.4 75 −50 7 –
3,5,5-Trimethylhexanoic 51.7 7.2 98 −32 7 –
Pentaerythritol
nC5 15.5 3.6 120 −54 85–95
nC7 22.2 4.9 151 −40 – 76–86
nC9 32.2 6.1 140 −7 1 –
nC8/nC10 30 5.9 145 −4 1 85
Oleic 64 10 141 −21 1 72–99
2-Ethyl hexanoic 44.8 6.4 88 8 – 18–22
3,5,5-Trimethylhexanoic 129.2 11.6 70 30 – <6

At the end of reaction, excess volatile reactant is removed by viscosity range at one or more relevant temperatures. Other
stripping under reduced pressure and recycled to subsequent parameters specifically relevant to the ester structure, or to
batches. downstream application requirements may also be included.
Similarly, for polyol esters, excess monoacid reactant For example, ester specifications often include Acid Value and
may be used and stripped at the end of reaction. Because Hydroxyl Value, which indicate degree of conversion of the
the excess acid serves to catalyze the reaction, polyol esters raw materials. Esters having a defined chemical composition
are frequently manufactured without use of additional may also be specified by nominal chemical purity.
catalyst. Trace impurities such as catalyst residues can significantly
The esterification and stripping process will typically give affect thermal, oxidative, and hydrolytic stability as well as
a conversion of acid groups of >99% in the crude reaction other application parameters including demulsibility, air
product. However, because very low acid values are gener- release, and foaming. The precise relationships are not always
ally required for lubricant basefluids, further reduction of acid well understood, so these issues may be addressed either by
concentration is frequently necessary. This can be achieved by specifying maximum levels for identified impurities such as
treatment of the crude reaction product by addition of a base heavy metals or ash content, or by specifying application test
such as sodium carbonate or calcium hydroxide, which con- performance criteria.
verts residual acid to carboxylate soaps that can be removed Application performance of ester basefluids is therefore
by water washing and/or filtration. dependent on processing parameters and upstream raw mate-
Adsorbents such as bleaching earths, activated aluminas, rial quality as well as on their nominal composition. Table 3.7
or activated carbons may be added prior to filtration to control lists typical specification parameters, indicating their depen-
color and concentration of specific impurities such as catalyst dence on each of these controlling factors.
residues. Final product purification is normally carried out by
cake filtration using a suitable filter aid in a press-type filter,
3.3 PROPERTIES AND PERFORMANCE
followed by polishing filtration to the required standard of
particulate cleanliness using bag or cartridge filters. CHARACTERISTICS

3.2.4.2 Specifications 3.3.1 Structure–Property Relationships


As for all lubricant basefluids, ester fluid specifications typi- Structure–property relationships for ester basefluids are
cally include pour point and flash point, which will normally broadly similar to those that have been established for mineral
define the useful application temperature range, as well as a oils and synthetic hydrocarbons [32].
Esters 55

Viscosity at 100°C in cSt


TABLE 3.7 6
Typical Ester Specification Parameters and Controlling 5
4
Factors
3
Specification Controlling factor 2
parameter 1
Nominal Raw Reaction Post-
composition material processing 0
4 5 6 7 8 9 10 11
Composition
Linear acid chain length
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

Acid value X X
Hydroxyl value X NPG TMP PE

Iodine value X X
Physical properties FIGURE 3.5  The effect of chain length on the viscosity of linear
Kinematic viscosity X acid polyol esters.
Viscosity index X
Figure 3.5 shows the effect of molecular weight by illustrat-
Density X
ing the change in viscosity at 100°C with acid chain length for
Low-temperature fluidity
a range of polyol esters. Table 3.8 demonstrates the variation
Pour point X X
Cloud point X
in viscosity due to degree of chain branching for PE esters of
High-temperature stability
C8 and C9 acids.
Flash point X X The effect of branching can also be seen from the data
Fire point X X in Table 3.9, which compare the physical properties of tri-
Cleanliness mellitate esters of isotridecanol from different sources.
Particle count X Isotridecanol is manufactured by carbonylation of C12 olefin,
Filterability X X which can be derived from C2, C3, or C4 raw materials, each
Demulsibility X X giving a characteristic isomer distribution. It is apparent that
Color X X X these nominally identical trimellitate esters have viscosities
that are sufficiently different that they do not fall within the
same ISO viscosity grade and are therefore not interchange-
However, a key advantage of ester basefluids is that this able materials.
understanding can be exploited more effectively to optimize
product properties because their compositions can be con- 3.3.1.2 Volatility
trolled to give more uniform molecular structures. This degree Volatility generally decreases with increasing molecular
of control is generally not possible with mineral oil basestocks weight but increases with reduced flexibility (branching and
that consist of complex mixtures of naturally occurring hydro- cyclics) since this prevents close packing and reduces the
carbons, nor with synthetic fluids manufactured by polymer- strength of intermolecular dispersion forces at high tempera-
ization processes (e.g., polyalphaolefins, polyalkylene glycols, ture [14,32].
polybutenes, Fischer–Tropsch gas to liquid isomerates), that Lubricant basefluids are typically specified in terms of
give a distribution of molecular weights. flashpoint or Noack volatility. For fluids containing a range
A corollary to this is that the physical properties (visco- of structures, both parameters are dominated by the behav-
metrics, density, etc.) associated with a defined ester structure ior of the most volatile component present. As a result, the
will, in principle, be fixed and invariant rather than varying flashpoint of a fluid containing a mixture of structures will be
within a controllable range as with a mineral oil distillation lower than that of a single-component fluid having the same
fraction. In practice, some variation in properties is generally average molecular weight. Furthermore, under high-tempera-
found due to batch-to-batch variation in raw material compo- ture service conditions where some of the fluid is volatilized,
sition. If properties are required that do not match those of
any single ester structure, then they can normally be achieved
by use of mixed raw materials (Section 3.2.2), or by blending TABLE 3.8
finished products. The Effect of Branching on Viscosity
Pentaerythritol Tetraester Viscosity at Viscosity at VI
3.3.1.1 Viscosity
40°C (cSt) 100°C (cSt)
Viscosity generally increases with increased molecular linear C8 25.5 5.5 141
weight and with reduced molecular flexibility. Ester molecu- branched C8 44 6.23 83
lar weight can be increased by either increasing the carbon (2-ethylhexanoic)
number of the raw materials, or by increasing raw material linear C9 31.5 6.1 149
functionality (more ester groups per molecule). Flexibility is branched C9 111 11.3 86
reduced by introducing branching, particularly quaternary (3,5,5-trimethylhexanoic)
branching and ring structures [14,32].
56 Synthetics, Mineral Oils, and Bio-Based Lubricants

TABLE 3.9
The Change in Physical Properties of a Trimellitate Using Different Sources of Isotridecanol
Isotridecanol type Viscosity at 40°C (cSt) Viscosity at 100°C (cSt) VI Flashpoint (°C) Pour Point (°C)
Propylene tetramer 315 19.9 67 298 −7
n-butene trimer 189 16.7 93 268 −17
Isobutene trimer 366 21.6 65 296 −23
Mixed C3/C4 320 20.8 73 234 −7
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

a mixed fluid will progressively increase in viscosity due to For the TMP series, the data show that although the satu-
selective loss of the lower molecular weight fraction, whereas rated stearate ester is a solid at room temperature, introduc-
a single structure fluid will maintain its initial viscosity until ing unsaturation (oleate) or branching (isostearate) reduces
oxidative degradation becomes significant. pour point to acceptably low values. Within the oleate series,
increasing symmetry by changing from trifunctional TMP to
3.3.1.3 Flow Properties (Viscosity even-numbered difunctional (NPG) or tetrafunctional (PE)
Index and Pour Point) leads to an increase in pour point. Data for glycerol triole-
In general, it is desirable for a lubricant basefluid to have both ate are included for comparison, showing that the triglyceride
a high VI and a low pour point, but the structural require- structure (selected by nature to pack into regular lipid mem-
ments for these two parameters are difficult to reconcile. VI is brane bilayers) has a much higher pour point than the syn-
highest for flexible structures containing long linear aliphatic thetic TMP trioleate.
chains, and is reduced by inflexible structural elements includ- There is therefore an inevitable trade-off between viscos-
ing branching and rings. However, at low temperatures, linear ity index (VI) and pour point. This can be resolved to some
aliphatic chains pack efficiently to form waxlike structures, extent by increasing structural diversity, but as noted earlier,
leading to gelling and pour point failure. An unhindered poly- this has a detrimental effect on volatility.
methylene (–CH2–)n structure with n = 9 or greater is typically Even with the structural control possible for synthetic
sufficient to induce solidification at temperatures within the esters, it is not possible to simultaneously optimize all per-
relevant range for high-performance fluids (i.e., >−40°C). To formance parameters. There is no single ideal lubricant struc-
prevent solidification, longer alkyl chains must be disrupted ture, only an optimum compromise that is unique to each
by introduction of unsaturation or branching [14,32,39]. intended application.
Like other viscous liquids, esters may supercool below
their true melting point and exhibit low pour point under 3.3.1.4 Lubricity
standard test conditions, but subsequently solidify during The factors affecting lubricity (i.e., coefficient of friction and
extended storage at temperatures above the measured pour wear rate) depend on the lubrication regime.
point [39]. Fluids based on a single molecular structure, and
particularly those that can pack with a high degree of sym- 3.3.1.4.1 Hydrodynamic Lubrication
metry, are more prone to solidification than those with a high Under hydrodynamic lubrication conditions, bearing surfaces
structural diversity. are separated by a relatively thick fluid-film, there is no wear,
Table 3.10 compares pour points for a range of polyol and the friction coefficient depends only on the fluid kine-
esters, all of which have very high VIs (>170) to illustrate matic viscosity at the operating conditions (temperature, pres-
these effects. sure, and shear). Hydrodynamic lubrication is therefore only
dependent on the physical properties of the lubricant basefluid
rather than its chemical composition. However, selection of a
TABLE 3.10 synthetic lubricant can be necessary in order to provide the
Pour Points for a Range of Polyol Esters best physical properties [40].
Longest To optimize energy efficiency, it is generally desirable to
CH2 Pour Point select a lubricant with the lowest viscosity sufficient to ensure
Polyol Acid Chain Symmetry (°C) full fluid-film lubrication at maximum expected operating
TMP Stearic 16 Low +45 temperature, combined with a high VI in order to minimize
TMP Oleic 7 Low −51 viscosity, and therefore viscous drag, at lower temperature
TMP Oleic/mixed 7 Very low −56 operation.
TMP Isostearic 9 (approx.) Low −30 Exact requirements will depend on specific application,
PE Oleic 7 High −21 but it is obvious from the structure–property relationships
NPG Oleic 7 High −24 outlined earlier that there will be a practical limit to the
Glycerol Oleic 7 Mid −16 lowest viscosity achievable before high volatility becomes a
limiting factor. Use of single component high VI ester will
Esters 57

normally represent a significantly more efficient viscosity/


volatility trade-off than will be achievable with mixed struc- TABLE 3.11
ture basefluids. Pressure–Viscosity Coefficients for a Variety of
In practice, lubricant formulators normally also use poly- Lubricants
meric VI improvers, rather than relying on basefluid VI alone.
Lubricant Pressure–Viscosity Coefficient GPa−1
This brings both advantages and disadvantages; the non-
Temperature 40°C 100°C
Newtonian rheology (reversible shear thinning) behavior of
polymer solutions may reduce hydrodynamic friction, but ISO 32 alkyl benzene 30.2 22.0
ISO 32 napthenic mineral oil 26.1 19.0
mechanical degradation (irreversible shear thinning) and oxi-
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

ISO 32 PE polyol ester 15.2 12.8


dative degradation of polymers can lead to deposit formation
ISO 68 PE polyol ester 19.3 14.9
and change of rheology during lubricant lifetime. Where VI
ISO 68 trimellitate 16.6 13.9
improvers are used in basefluids with significant ester con-
tent, their rheological properties will be modified because of
differential solvation by the more polar fluid. Polymer/base-
fluid interactions are very complex and require case-by-case their EHL behavior. Conversely, P–V coefficients that have
optimization. been determined indirectly by measurement of film thick-
ness under EHL conditions may be significantly dependent on
3.3.1.4.2 Elastohydrodynamic Lubrication presence of low-concentration components.
As the lubrication regime passes from hydrodynamic into Because of this complexity, it is very difficult to theoreti-
elastohydrodynamic lubrication (EHL), the thickness of the cally predict and rationalize EHL behavior, but the effects
fluid-film reduces and may approach molecular dimensions. described can be exploited practically by lubricant formula-
At the same time, the pressure in the contact zone increases tors to control lubricity in EHL contacts. For example, recent
dramatically. The rheology and even the composition of this work suggests that when a small amount of a high-viscosity
thin high-pressure lubricant film depend on the pressure–vis- polar ester is added to a low-viscosity nonpolar basefluid
cosity coefficient of the basefluid and on the surface affinity (PAO), the ester will preferentially stick to the surface. When
of the various components of a formulation and are generally the two metal surfaces are far apart, the bulk viscosity is con-
very different from those of the bulk lubricant. Lubricity under trolled by the PAO. When the surfaces come closer together,
EHL conditions is also much more dependent on the materials the PAO is squeezed out of the contact zone. The polar ester
of construction and surface finish of the contacting surfaces. sticks to the surface and stays in the contact area. As the ester
The key lubricity parameters are the EHL friction coefficient has high viscosity, the bulk viscosity of the oil will increase
and the maximum load that can be supported. A higher EHL as the surfaces come closer together. Finally, a point will be
film viscosity, that is, higher P–V coefficient, will typically reached where only the more viscous polar ester remains [45].
give a higher friction coefficient, but a stronger film [40]. Such an effect can be very beneficial as it will increase the
Pressure–viscosity coefficients typically have an inverse strength of the EHL film and may allow reduction in dose rate
relationship to VI (i.e., high VI normally implies low P–V of antiwear additives.
coefficient). Because current commercial synthetic esters are
generally designed to have higher VI than mineral oils, they 3.3.1.4.3 Mixed Film
have relatively lower P–V coefficients. The P–V coefficient Mixed film lubrication, as the term implies, is actually a com-
of a lubricant is increased by increasing degree of branching bination of boundary and hydrodynamic or EHL regimes. In
and number of cyclic structures, presence of aromaticity, and the mixed lubrication regime, the contact characteristics are
increasing length of side chains branches [32]. determined by varying combinations of fluid-film and bound-
Values for P–V coefficient at 40°C have been reported ary lubrication effects. Some asperity contact may occur and
[41–44] for TMP ester (8.4–9.8 GPa−1), PE esters (7.5–12.2 interaction takes place between mono- and multilayer bound-
GPa−1), phthalates (13.6 GPa−1), and diesters (6.6–7.6 GPa−1). ary lubricating films while a partial fluid-film lubrication
Note, however, that P–V coefficients vary with temperature as action develops in the bulk of the space between the metal
shown in the results in Table 3.11. surfaces. It follows that both the physical properties of the
Furthermore, in real life lubricant formulations where mul- lubricant (per EHL) and the chemical properties of the lubri-
tiple components are present, the more polar components are cant (per boundary) are important [46].
concentrated in the layers adjacent to the polar metal surface.
This effect is well known as the basis of action of friction 3.3.1.4.4 Boundary Lubrication
modifier additives, but has also been observed in blends of Under boundary conditions, there is asperity contact between
polar and nonpolar basefluids, for example, PAO/ester under the surfaces. The load is mainly carried through these solid
EHL conditions [45]. Because the thickness of the EHL film contacts, the major contribution to the frictional force is the
may be as low as a few tens of molecules, this effect can sig- energy required for deformation of contacting asperities, and
nificantly alter the composition and rheology of the EHL film boundary friction coefficients generally show little depen-
compared to that of the bulk composition. P–V coefficients dence on lubricant viscosity. However, lubricant baseflu-
of pure basefluids are therefore not necessarily related to ids or additives can modify the boundary friction and wear
58 Synthetics, Mineral Oils, and Bio-Based Lubricants

properties. In particular, components that form a coherent components that lead to deposit and varnish formation. For
chemisorbed or physisorbed layer that deforms more readily most ester basefluid structures, the polarity is only slightly
than the underlying metal surface will reduce the boundary higher than that of mineral oil and such esters have excellent
friction coefficients. If this film is strong and thick enough to compatibility with most common types of lubricant baseflu-
prevent metal–metal contact, it will also reduce wear. ids. Ester basefluids can therefore be blended with the very
Lubricants having a high polarity or affinity for metal oxide nonpolar synthetic hydrocarbon basefluids or Gp III and
surfaces, such as esters, have a greater tendency to form such III+ basefluids, to increase solubility of VI improver, deter-
adsorbed layers than less polar fluids, such as mineral oils gent, and friction modifier additives that would otherwise be
or synthetic hydrocarbons, and therefore have lower bound- insoluble. In particular, esters may be used as the basefluid in
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

ary friction coefficients. Esters containing predominantly additive packages, where the polar components are present at
linear alkyl substituents can form a more coherently packed much higher concentrations. Furthermore, esters can be used
adsorbed film, and consequently show lower boundary fric- in machinery that previously used mineral oil, PAO, PIBs, and
tion coefficients, than those with branched alkyl substituents. most PAGs without significant cross-contamination problems.
This principle can be extended to the use of components with However, for ester fluids with a high concentration of
longer linear alkyl chains and polar head-groups (acid, amide, ester groups and high molecular weight, for example, com-
or partial ester), which are widely used as friction-modifying plex esters derived from low carbon number raw materials,
additives in a range of lubrication applications, particularly in the polarity of the ester itself may be so high that it becomes
low-polarity basefluids such as hydrocarbons [1,46]. In some immiscible with low polarity hydrocarbon basefluids at lower
cases, the lower friction coefficients observed with ester base- temperatures.
fluids may in fact be due to the formation of trace levels of As noted earlier, polar esters may be preferentially
their carboxylic acid raw materials by thermolysis of the ester adsorbed on contacting metal surfaces and modify mixed
bonds under the very high local flash temperature conditions and boundary lubricity. However, they can also interact with
of asperity contact. Carboxylic acids have higher polarity than other engineering components, in particular elastomeric poly-
the ester group because they can act as hydrogen bond donors, mers used in seals and gaskets. Elastomeric polymers have
or even form surface-bound metal carboxylate soaps, and are relatively open cross-linked structures and basefluid compo-
therefore preferentially adsorbed even at very low concentra- nents may be absorbed into the polymer network, causing it to
tions [47,48]. swell and soften. This effect can either be desirable to ensure
However, although ester basefluids can reduce wear by effective sealing or undesirable if it leads to degradation of
increasing strength of an EHL film and reduce boundary mechanical properties through chemical reaction of the poly-
friction by formation of a low surface energy adsorbed layer, mer network or loss of low molecular weight polymer com-
they do not inherently contribute to the formation of the type ponents. Thus, esters are frequently used as seal swell agents
of thicker film formed by antiwear additives such as ZDDP. in synthetic hydrocarbon formulations to prevent elastomer
In fact, high polarity ester basefluids may interfere with the shrinkage, but care is required to ensure elastomer compat-
action of antiwear additives either by excluding the antiwear ibility where esters are used at high concentration.
additive from the immediate metal surface, or by solvating the Table 3.12 gives an overview of typical ester compatibility
initial products of antiwear additive decomposition such that with commonly used elastomers and other engineering mate-
they do not form a stable film [49]. rials. It is important to note that the processing of the elasto-
Under very severe extreme pressure conditions, asper- mer can have a major impact on its performance. As esters
ity contact inevitably results in metal–metal contact that can can be efficient solvents, they have the potential to extract
remove the protective metal oxide surface layer leading to any substances used during the manufacture of the elastomer.
welding between exposed metal surfaces. Under these con- Elastomers from different suppliers can have significant dif-
ditions, scuffing wear can be minimized by the presence of ference in terms of the degree of cross-linking, fillers, and
components that lead to rapid reformation of a protective process residuals. Compatibility will also depend on the spe-
nonmetallic layer or metal oxide, sulfide, or equivalent. The cific ester type and the contact conditions. Therefore, any
oxygen content of ester lubricants can help minimize scuffing information on ester compatibility with elastomers in general
wear by serving as an available reservoir of oxygen for rapid should be confirmed by appropriate tests on the specific mate-
oxide formation [46,47]. rials of interest.
Finally, because of the stronger intermolecular bonding,
3.3.1.5 Polarity and Material Compatibility ester basefluids typically have lower volatility than equivalent
The dipole moment of the ester functional group results in hydrocarbons, although the magnitude of this effect is small
dipole–dipole and dipole-induced dipole interactions between compared to the volatility reduction due to narrower molecu-
ester basefluid molecules and other components. These forces lar weight distribution.
are stronger and more localized than the dispersion forces The magnitudes of all of these polarity effects depend on
that act between aliphatic hydrocarbons [50]. This has a direct the number of ester groups per molecule and the size of the
impact on application performance in a number of ways. molecules, as well as on how the ester groups are distributed,
Esters are good solvents for polar additives and for the and the structures surrounding them. These effects of sol-
polar initial products of thermal decomposition of lubricant vency, polymer compatibility, and fluid miscibility are of great
Esters 59

practical interest in many other areas of technology, and many


TABLE 3.12 attempts have been made to improve the theoretical under-
Compatibility Data for Esters standing by quantitative group contribution methods [50].
Suitable Marginal Unsuitable However, the majority of these general approaches are either
too oversimplified or too complex for practical application.
Elastomers
A widely used measure of ester lubricant polarity is the
Nitrile rubber Nitrile rubber (Buna-N, Nitrile rubber
nonpolarity index or NPI, defined as [49]:
(Buna-N, NBR) NBR) (Buna-N, NBR)
Only if nitrile exceeds With nitrile content With nitrile content
36% 30%–36% below 30% Nonpolarity index ( NPI ) = Total no. of atoms ´ RMM
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

Fluorosilicone rubber No. of carboxylate groups ´ 100


Polyurethane Natural rubber where
Fluorocarbon (Viton, Ethyl propylene Styrene–butadiene RMM is the relative molecular mass.
Teflon) terpolymer (EPDM) rubber (SBR)
NPI of esters was originally developed to predict interac-
Polyester (Hytrel) Polyacrylate rubber Butyl rubber
tions of esters with polar system components, for example,
Ethylenepropylene Chlorosulfonated
copolymer (EPR) polyethylene (very
elastomer swell and interference with the activity of polar
marginal?) antiwear additives. NPI is a measure of nonpolarity, so a high
Silicone rubber Polychloroprene value implies a relatively low polarity [49]. High NPI esters
(Neoprene) (very have relatively low interactions with polar elastomers and
marginal?) antiwear additives.
Ethylene/acrylic However, because the NPI metric conflates two indepen-
(EAE) dent parameters, namely ester polarity and molecular weight,
Polysulfide (Thiokol) into a single value, it only correlates with experimental mea-
Paints surements when the effects of these two parameters are acting
Epoxy Oil-resistant alkyds Acrylic in the same direction. For prediction of compatibility of esters
Baked phenolic Phenolic Household latex with nonpolar components, such as solubility in low polarity
Two component Single component Polyvinyl chloride basefluids, it is not applicable.
(PVC)
Urethane Urethane
Moisture-cured Industrial latex Varnish 3.3.2 Stability
urethane
Lacquer
3.3.2.1 Thermal and Oxidative Stability
Plastics Because thermal and oxidative stress are normally encoun-
Nylon Polyurethane Polystyrene tered simultaneously in lubricant applications and their
Fluorocarbon Polyethylene Polyvinyl chloride practical effects are difficult to separate, they are normally
(PVC) discussed together, but they are in fact distinct processes that
Polyacetal (Delrin) Polyproylene Styrene (ABS) can operate independently.
Acrynitrile–butadiene Polysulfone Styrene acrylonitrile The ester group itself is highly resistant to oxidation, and
(Celcon) (SAN) slightly increases the oxidative stability of the adjacent car-
Acetals Melamine Polysulfones bon atoms [31,51]. The initial oxidation chemistry of ester
Polyamides Phenylene oxide (Noryl) Acrylic (Lucite, basefluids is therefore primarily that of the substituent hydro-
Plexiglas) (very carbon groups. In principle, during the later stages of oxida-
marginal?)
tive decomposition, hydroxyl and carboxyl functional groups
Polycarbonate
formed by oxidation may undergo transesterification reac-
(Lexan) (very
marginal?)
tions with the ester groups originally present to form degra-
Polyphenyloxide
dation products, which would not be formed so readily from
Metals hydrocarbons [52–54]. However, in practice, such extensive
Steel and alloys Cadmium Lead degradation would not occur during the normal lubricant
Aluminum and alloys Zinc lifetime.
Copper and alloys Magnesium The detailed processes involved in hydrocarbon oxidation
Nickel and alloys have been intensively studied and are relatively well under-
Titanium stood [55,56]. The chemistry will not be described in detail
Silver here, except to note that the initial attack on each molecule
Chromium is by free radical abstraction of a hydrogen atom. The rate of
Tin this process depends on the stability of the free radical thus
formed. The order of stability of C–H bonds in the structures
likely to be encountered in lubricant basefluids are as follows
(Ar = benzene ring):
60 Synthetics, Mineral Oils, and Bio-Based Lubricants

a). Ester with beta hydrogens (e.g. diesters)


CH3>ArH=-CH2->-CHR->-CH2-CHCH->

-CH2-Ar->-CHCH-CH2-CHCH- H H O H H O
R R’ R R’
The relative stability of the molecule with respect to initial oxida- O O
tive attack depends on the number of each of these bond types in H H
H H
the molecule and their steric accessibility. This leads to the fol-
lowing “rules of thumb” for oxidative stability of ester basefluids: R H HO
+ R’
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

Linear and quaternary branched esters are more stable O


than tertiary branched. H H

Short chains are more stable than long.


For polyol esters, the order of stability is b) Esters without beta hydrogens (e.g. polyols)

PE > DiPE > TMP > NPG.
Esters with olefinic unsaturation, for example, oleates, R R O R R O

have lower oxidative stability. R R’ R + R’


Esters containing olefinic polyunsaturation (conjugated O O
or methylene-interrupted), for example, most veg- H H H H
etable oils, have very low oxidative stability [57,58]. R R
O
+ R’
Because the fundamental oxidation processes are essentially
HO
identical, conventional lubricant antioxidants have similar H R
effectiveness in ester basefluids as in mineral oils or synthetic
hydrocarbons. Note that comparisons of oxidation stability FIGURE 3.6  Thermal decomposition mechanism of esters.
between fluids should only be carried out in the presence of
equivalent and realistically high levels of antioxidants, since in practical lubricant applications both processes occur simulta-
lightly processed mineral oil and vegetable oils may already neously and interact in complex ways. Both oxidation and ther-
contain trace (but variable) levels of naturally occurring free molysis of esters lead eventually to formation of carboxylic acid
radical inhibitors whereas synthetic fluids do not [59–61]. groups, which increase the acid value of the oil. Increased acid
Residual unreacted alcohol groups have been shown to value can accelerate both processes and can also result in cor-
have some antioxidant activity in ester basefluids due to side rosion. Corrosion leads to increase in dissolved metals, which
reactions, which decompose free radicals [62]. However, the further catalyze both oxidation and thermolysis. Alternatively,
effect does not become significant except at concentrations oxidation processes may result in deposition of lacquers or var-
sufficient to increase the polarity of the fluid to a level that nishes on metal surfaces that inhibit further corrosion and sur-
affects compatibility with other formulation components. face-catalyzed fluid decomposition processes [67,68].
When thermal stability is measured in the absence of oxi- Metal catalysis has two roles in ester oxidation: it promotes
dative stress or catalysts, the ester linkage is normally the first the degradation of ester molecules by accelerating hydroper-
site of thermolysis [11,12,63]. For esters with hydrogen atoms oxide decomposition and it affects condensation of oxidation
attached to the β carbon of the alcohol residue, the predomi- products to form sludge and varnish. The rate of catalytic activ-
nant decomposition pathway is β elimination to form carbox- ity on oxidation rate in decreasing order of activity was found
ylic acid and olefin. The rate of this process is insignificant to be the following: low-carbon steel > stainless steel​  > le​
at most lubricant operating temperatures but becomes detect- ad  > ​alumi​num   >​  bras​s   > c​opper​. The rate of catalytic activ-
able at temperatures >275°C. For esters without β hydrogens, ity on the deposit formation in decreasing order of activity
particularly the neopentyl polyol esters, this pathway is not was found to be the following: low-carbon steel > stainless
possible and decomposition is by a free radical route at signif- steel > aluminum > brass > copper > lead [68].
icantly higher temperatures (>315°C) [64,65]. These pathways Thus, practical optimization of thermal and oxidative stabil-
are illustrated in Figure 3.6. ity of a formulated ester lubricant requires more than just base-
In contact with untreated ferrous metal, the decomposition fluid structure optimization. Additives, including antioxidants
temperatures of polyol esters are significantly reduced and (e.g., phenolics and aminics), which act as free radical scaven-
decomposition is detected at temperatures as low as 200°C. gers, antiwear additives (e.g., phosphites can act as hydroperox-
However, in the presence of phosphate esters, such as tricre- ide scavengers), metal passivators (e.g., phosphates), and metal
sylphosphate (TCP), the metal surface is passivated and does deactivators (e.g., thiadiazoles), can act independently or syn-
not catalyze decomposition [66]. For this reason, neopentyl ergistically to stabilize the formulation [55,61]. Decomposition
polyol esters used in high-temperature applications such as avi- products of these additives, however, particularly strong acids,
ation gas turbine lubricants are normally addivated with TCP. may in turn have a detrimental effect on stability of the ester.
Although separate thermal and oxidative degradation pro- However, other things being equal, use of an ester base-
cesses can be identified for ester basefluids under ideal conditions, fluid designed and specified for inherently high thermal and
Esters 61

materials. In practice, this is rarely a problem in nonaqueous


TABLE 3.13 lubricant applications because at ambient temperatures the
Panel Coker Results on a Variety of Lubricants reaction rate constant is low and at elevated temperatures the
Addivated with 1% Aminic Antioxidant (275°C for 22 h residual concentration of water dissolved in the lubricant and
in Air) available for reaction is low. However, an understanding of
the fundamental processes involved in ester hydrolysis may
Lubricant Visual Demerit Rating
be useful to prevent any potential problems developing.
Mineral oil 3.95 Hydrolysis occurs via an acid-catalyzed mechanism and
PAO 6 >4.00 the reaction rate is proportional to the concentration of water
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

Standard diester 2.50 and the concentration of acid. A hydrogen ion adds to the car-
Standard phthalate 1.07 boxyl oxygen of the ester linkage, converting it transiently to a
Standard trimellitate 0.50 carbonium ion, which rapidly adds water to form a positively
Optimized trimellitate 0.04
charged tetrahedral intermediate. This tetrahedral interme-
Standard TMP polyol ester 0.57
diate then separates into carboxylic acid and alcohol, regen-
Optimized TMP polyol ester 0.18
erating a proton that can then catalyze further reaction. The
Visual rating system reaction is therefore autocatalytic. The more acid that forms,
>2.0   Thick carbonaceous deposits. the faster the breakdown, which in turn creates more acid.
2.0–0.5 Low levels of deposits. However, water is consumed in the reaction, which can only
<0.5   Virtually no deposits. progress if more water is available [69,70].
There are therefore two fundamental ways to control the
rate of ester hydrolysis: reduce the water concentration or
oxidative stability can significantly improve lubricant per- reduce the acid concentration. Other trace components such
formance. This is illustrated by the Panel Coker test results as residues of transition metal esterification catalysts may also
shown in Table 3.13. In the Panel Coker test, a lubricant for- act as catalysts and must be controlled, but unlike the acid,
mulation is dripped onto a heated steel plate at 275°C for 22 h their effect is not autocatalytic [71].
in air. The lubricant is collected at the bottom of the plate and The rate of hydrolysis can in principle also be limited by
recirculated for repeated application to the steel plate. At the restricting temperature, although this is not practically pos-
end of the test, the deposits are visually analyzed and given a sible for most lubricant applications.
demerit rating. Several lubricants addivated with 1% aminic Different ester structures have different inherent rates of
antioxidant have been evaluated [107] (Table 3.13). hydrolysis. Aromatic esters and esters with branching at the 2
The results clearly show the effects described earlier. position of the acid group hydrolyze more slowly.
Aromatic phthalates give superior results to aliphatic diesters The extent of hydrolysis can be measured by monitoring
and polyol esters give superior performance to diesters. Use of the acid value of the fluid. Figure 3.7 shows the progress of
highly oxidatively stable substituents such as 3,5,5-trimethyl- hydrolysis for polyol esters in a sealed system, modeling the
hexyl derivatives allows further optimization of oxidative sta- behavior of hermetically sealed refrigeration compressor
bility within each ester category. The effect of residual metals lubricants. As water is consumed, the acid value of the polyol
(catalysts, neutralizing agents, etc.) can be seen in Table 3.14. ester increases. When the water is completely consumed, the
It is clear that even small levels of certain contaminants can acid value reaches a plateau. The results illustrate that the
make a huge difference to the deposit-forming tendency of final degree of “de-esterification” or hydrolysis is directly
an ester. This reemphasizes the need for a tight specification. related to the amount of water present.
Use of a more hydrolytically stable structure reduces the
3.3.2.2 Hydrolytic Stability rate of hydrolysis but does not change the final acid value
As noted earlier, the esterification reaction is reversible (Figure (Figure 3.8).
3.1) and the ester products can in principle undergo hydro- In most cases, however, it is not possible to prevent water
lysis, that is, reaction with water to regenerate the starting ingress during lubricant service life. Therefore, where hydro-
lytic stability is a concern, it should be controlled by a tight
initial acid value specification and limiting the use of addi-
TABLE 3.14 tives that may themselves hydrolyze more rapidly to produce
The Effect of Process Residuals on the Deposit acid by-products (e.g., phosphate esters).
Formation of a Trimellitate in the Panel Coker Test There is also a theoretical possibility of interesterifica-
tion or transesterification between different ester groups. As
(275°C for 22 h in Air)
with simple hydrolysis, these processes are normally too slow
Residual Metals in ppm Visual Demerit Rating to be significant, but can be catalyzed by some additives or
0–5 0.5 contaminants. Transesterification may also occur if a fluid is
6–10 1.01 subjected to thermal cycling where partial hydrolysis is fol-
11–15 1.33 lowed by re-esterification at higher temperature. In blended
basefluids this would result in change of composition due to
62 Synthetics, Mineral Oils, and Bio-Based Lubricants

0.8 250 8 2500

Acid value (mg KOH/g)

Water value (ppm)


Acid value (mg KOH/g)
0.7 7

Water value (ppm)


200 2000
0.6 6
0.5 150 5 1500
0.4 4
0.3 100 3 1000
0.2 2
50 500
0.1 1
0 0 0 0
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

0 2 4 6 8 0 10 20 30 40
Time (Days) Time (Days)

Water Value Acid Value

FIGURE 3.7  Polyol ester tested at 2000 ppm and 200 ppm water at 150°C in a sealed tube containing a metal coupon.

redistribution [113]. For example in a blend of diester and 3.3.3 Safety, Health, and Environmental Properties
polyol ester, transesterification leads to formation of monoes-
ter and complex ester components. This may lead to change of 3.3.3.1 Biodegradability
viscosity and reduction of flash point during service. The biochemistry of microbial attack on esters is well known
in general outline and has been reviewed in detail [72–76].
3.3.2.3 Shear Stability The initial step is normally hydrolysis of an ester group, either
Ester basefluids show similar shear stability to hydrocarbons abiotically or catalyzed by microbial lipase enzymes. The ini-
of similar viscosity profiles. Materials of relatively low viscos- tial hydrolysis products have higher water solubility than the
ity intended for use as basefluids show Newtonian rheology, ester and are therefore more bio-available, so hydrolysis of
with no dynamic shear thinning effects at shear rates likely to the remaining ester groups occurs more rapidly, releasing the
be encountered in service and no permanent shear thinning. original acid and alcohol raw materials for biodegradation by
Blends containing higher viscosity materials such as high microbial oxidation.
molecular weight complex esters may exhibit dynamic shear The biodegradability of ester basefluids is therefore pri-
thinning. Mechanical degradation of long polymer chains marily determined by the relative hydrolytic stability of the
leading to permanent shear thinning may become significant ester groups, and the biodegradability of the acid and alcohol
under very high shear conditions for materials with molecular raw materials [77,78].
weights above about 10,000 amu [47]. Many different biodegradability test protocols have been
developed to simulate different real world environmental con-
ditions. Older literature (including previous editions of this
work) often refers to initial biodegradability in fresh water
determined by the CEC L-33-A-95 protocol, based on quan-
tification of solvent-extractable non-polar components. This
method is no longer used because it required halogenated sol-
vents for extraction. A replacement (CEC L-103-12) has been
developed [114]. However, most current specification criteria
refer to Ready Biodegradability assessed by OECD 301 or
equivalent methods which measure biodegradation by oxygen
uptake or CO2 formation over 29 days [115]. Since the hydro-
lysis reactions increase polarity but do not involve oxygen
uptake or CO2 formation, they contribute to initial biodegrad-
ability, but are not detected by ready biodegradability tests.
Biodegradability values by CEC L-33-A-95 are therefore typi-
cally higher than values measured by OECD 301.
Esters of predominantly linear raw materials are gener-
ally readily biodegradable, particularly those derived from
unmodified natural fatty acids. Esters with hydrolytically
stable ester groups (particularly esters of aromatic and
FIGURE 3.8  Hydrolytic stability of a linear acid versus a branched 2-alkyl substituted acids) or branched hydrocarbon chains
acid polyol ester tested at 2000 ppm water at 150°C in a sealed tube have much lower biodegradability. Some typical values
containing a metal coupon. for biodegradability are shown in Table 3.15. However, it
Esters 63

segregated. When esters are diluted in larger volumes of


TABLE 3.15 mineral oil, as in recovered automotive engine oils, the ester
Biodegradabilities of Various Ester Lubricant Groups content becomes a low-value trace component, which may
% Biodegradability % Biodegradability be removed as part of the heavy ends waste stream during
reprocessing. End of life fate of ester basefluids therefore
CEC-L-33-A-95 (21
depends on the application. However, the large majority of
Ester Type Days) OECD 301B (28 Days)
used ester basefluids are currently eventually disposed of
Monoesters 70–100 30–95 by combustion. Esters have slightly lower calorific value
Diesters 70–100 10–80 than mineral oil but otherwise have similar combustion
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

Phthalates 40–100 5–70 characteristics.


Dimerates 20–80 10–50
Trimellitates 20–80 0–40 3.3.3.4 Renewability
Linear polyols 80–100 50–99
Some specifications for biolubricant applications include a
Branched polyols 20–50 5–40
requirement for a minimum bio-based or renewable content.
Complex polyols 0–90 10–90
This is calculated on the basis of the number fraction of car-
bon atoms in the product which were derived from renewable
rather than petrochemical sources, not on the weight fraction
should be noted that biodegradability tests are characterized of raw materials used. As noted above, some ester raw mate-
by high variability. Significantly different results may be rials are commercially available based on either renewable
obtained for identical materials. For example, results rang- or petrochemical sources and may be used interchangeably.
ing from 56–80% have been reported for the diester fluid For esters based on such raw materials, suppliers normally
diisotridecyl adipate. Ready biodegradability data are nor- quote minimum and maximum values for renewable carbon
mally included in supplier literature and safety data sheets content. The actual renewable content of a specific sample of
for ester lubricants. ester lubricant can be verified by radiocarbon analysis (e.g.,
ASTM D6866 - 16 or equivalent).
3.3.3.2 Toxicity and Ecotoxicity
Esters used as lubricant basefluids generally have low acute
3.4 APPLICATION AREAS
toxicity by ingestion, skin absorption, or inhalation of air-
borne mist. However, concerns exist over possible long-term 3.4.1 Engine Oils
effects of some phthalate esters. Phthalates have been inten-
sively studied due to their high volume use as plasticizers Engine oils are the highest volume lubricant application for
in consumer products. This has led to legislation restricting mineral oils and synthetic hydrocarbons, but relatively less
applications of some phthalates, including bis(2-ethylhexyl) significant for esters. Application requirements for automo-
phthalate (DOP) which was historically the most widely used tive engine oils are reviewed in detail elsewhere in this book
PVC plasticizer. The materials of concern were not widely (Crankcase Lubricants). The long-term drivers for the auto-
used in lubricant applications and have been largely substi- motive lubricant market are cost of ownership, reliability, and
tuted by other aromatic esters. compliance with legislative standards for fuel economy and
Lubricating oils can act as solvents for natural fats present emissions (Automotive Trends). These application perfor-
in the skin and prolonged skin contact can result in physical mance drivers translate into the following requirements for
damage due to defatting and potential for contact dermatitis. lubricant formulation properties [79–81]:
Because of their higher polarity, esters are generally better
solvents than mineral oils of similar viscosity and can give Low viscosity at low temperatures for short-trip fuel
rise to such effects more rapidly. Use of appropriate personal economy, with acceptably low volatility even at
protective equipment is therefore particularly important when increasing engine operating temperatures.
working with ester fluids. Increased oxidation stability to increase lubricant life-
time and service intervals.
3.3.3.3 Recycling and Reuse Reductions in sulfated ash, phosphorus, and sulfur
Uncontrolled lubricant emissions to the environment due to (SAPS) to ensure exhaust catalyst durability and
losses in service and illegal disposal of used oils have been compliance with emissions regulations.
significantly reduced and globally approximately 50% of Improved wear protection, even with reduced levels
lubricant basefluids are now recovered. The majority of this of traditional antiwear additives and lower fluid
material is used as fuel oil replacement in heating or power viscosity.
generation [16], but an increasing proportion is recycled Retention of boundary friction modifier activity
for reuse as lubricant basefluid (see Re-refined Base Oils). throughout the lubricant lifetime.
However, because the application benefits of esters depend Lower deposit formation, particularly carbonaceous
on their chemical structure, beneficial re-use in lubricant cylinder deposits which may contribute to Low
applications is only possible where materials can be strictly Speed Pre-Ignition (LSPI).
64 Synthetics, Mineral Oils, and Bio-Based Lubricants

From the discussion of structure–property relationships polarity and solvency characteristics of the ester are therefore
in the previous sections, it is apparent that synthetic ester of more relevance than its viscosity and VI. Effectively the
technology has the potential to deliver optimized viscosity/ ester behaves more as a functional additive than a basefluid
volatility performance, inherently high VIs, good solvency component. However, because of the large size of the automo-
for oxidation products, and excellent frictional properties. tive engine oil market and the increasingly high penetration
Esters have therefore always attracted interest as potential of synthetics, this sector remains a large and growing applica-
engine oil basefluids and formulations have been developed tion area for ester fluids.
based on straight ester fluids [116] and even on vegetable
oils (Vegetable-Oil-Based Internal Combustion Engine Oil).
3.4.2 Automotive Gear
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

However, the required application properties of engine oils


cannot be delivered by basefluid properties alone; engine oils The automotive gear oil market is subject to the same overall
are complex formulations which rely on a range of additives industry drivers as crankcase oils [83], with the additional feature
(see Crankcase Lubricants). Additive technology has devel- that new technologies with distinct lubrication requirements, such
oped incrementally over many decades with mineral oils as automatic dual clutch (DCT) and continuously variable trans-
dominating the market. Conventional additive packages have missions (CVT), are increasingly important (see Transmissions
therefore been designed for activity in mineral oils and are and Transmission Fluids, Automotive Gear Lubricants).
not optimized for performance in higher polarity ester fluids. The expectation for fill-for-life gear lubrication and
Fully ester-based engine oil formulations have therefore not demands for improved fuel economy leading to smaller, lighter
found application except in niche areas such as racing oils. transmission systems with reduced lubricant charge demand
Current synthetic engine oils are based on PAOs, GTL optimized thermal, oxidative, and shear stability to cope with
basestocks, or Gp III+ oils as the main basestocks. These higher stresses for longer lifetimes. Fuel economy and user
highly aliphatic hydrocarbons have lower polarity than Gp expectations for smooth shifting demand optimized lubricity.
I and Gp II fluids and perform differently in terms of addi- The main advantages of esters in this sector are their rela-
tive compatibility, oxidation product solubility, and elastomer tively high VIs, and low-temperature flow properties. This
swell. Low levels of ester basefluids are therefore typically allows synthetic gear oils to operate over a much wider tem-
added to give a blended basefluid with optimized polarity. perature range than conventional oils. Also the good shear sta-
Products of this type have been established in the market- bility of ester-based oils gives major performance advantages.
place since the 1960s [62,82]. Initial formulations were based Synthetic manual transmission fluids, like synthetic crank-
on PAOs with a small amount of a phthalate (to act as a seal case oils, are typically based on PAO blended with ester as
swellant). These formulations were followed by PAO/adipate a minor component for polarity modification [84]. A typical
ester blends where the diester was used at between 5% and ester-based 75W gear oil can be found in Table 3.17. Fluids
20%. Here, the ester acted as a seal swellant, an additive sol- for automatic DCT systems are generally similar to manual
ubilizer, and made an important contribution to the desired transmission fluids with adjustments to optimize anti-shudder
deposit/volatility targets. A typical 5W/40 engine oil formu- performance.
lation of this type is shown in Table 3.16. Short chain TMP Fluids for toroidal CVT systems require very high EHL
esters were preferred over diesters due their superior thermal friction. This implies fluids having very high P–V coeffi-
stability [81]. However, pricing of the short chain C8C10 acid cients so that the lubricant transiently solidifies in the EHL
raw materials has been volatile over recent years which has contact, enabling it to transmit traction forces. The structural
disincentivized use of the TMP esters requirements for such traction fluids are well understood, but
Current formulations typically use ester dose rates of <10%. hydrocarbons with appropriate structures are difficult to man-
At these low levels the properties of the ester make a rela- ufacture. Ester-based fluids have therefore been extensively
tively low contribution to the viscosity of the formulation. The investigated [85]. However, although their synthesis is more
straightforward, the high flexibility of the ester linkage limits
the maximum traction coefficient, and hydrocarbon technology
TABLE 3.16 has become the most widely adopted (see Cyclohydrocarbons)
Typical Synthetic Passenger Car Motor Oil Formulation although esters are still of interest as minor basefluid compo-
% Dose nents for rheology optimization [86,87].
Component Rate Comments
Ester basestock 5–20 Diesters or TMP polyol esters
Friction modifier (FM) 0–3 Amide/ester organic FM MO-based TABLE 3.17
inorganic FM Typical Synthetic Gear Oil Formulation (e.g., 75 W)
Hydrocarbon oil 40–70 PAO, hydrocracked, alkyl
Component % Dose Rate Comments
naphthalene
VI improver 8–15 Ester 5–30 Diesters polyols
Additive pack 10–20 Antiwear, detergent inhibition pack, PAO 60–85 Blends of PAO 4, 6, 40, and 100
etc. Additive pack 7–13
Esters 65

3.4.3 Two Stroke in the former USSR. Esters rapidly became the preferred
basefluid for both civil and military aviation elsewhere [5].
Ester lubricants offer a number of advantages over mineral Efforts continued to identify even higher performing fluids
oils as the lubricant component of two-stroke engine mixtures. and a very wide range of alternative SBFs have been evalu-
The good solvency characteristics result in lower deposits and ated. Some have found limited military use (see Polyphenyl
less engine fouling. The lower volatility and boundary friction Ethers) [117] but esters have remained the most widely used
allow use of lower viscosity oils, removing the requirement fluid type, delivering performance adequate for applications
for high-viscosity mineral oils (brightstock) and allowing including civil supersonic transport (SST) between 1969 and
leaner burn ratios. In turn, this significantly reduces exhaust 2003 [118].
fume (smoke) due to unburnt or partially combusted lubricant
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

The first generation of oils (Type I) were diesters, typi-


[88]. The excellent solubility of esters also allows them to be cally based on bis(2-ethylhexyl) sebacate, but these have been
used without solvents (which are usually added to conven- replaced in most applications by more thermally stable polyol
tional two-stroke oil to help miscibility with the fuel and low- esters. Diesters are still used in less demanding applications
temperature fluidity). A typical ester-based formulation can such as small private aircraft or turbo-prop engines [91].
be seen in Table 3.18. Since gas turbine engines were originally developed for
The leaner burn ratios result in reduced oil emissions, which military aircraft, lubricants were first controlled by U.S. mili-
is a benefit in environmentally sensitive applications such as tary specifications MIL-L-7808 (3 & 4 cSt oils) and MIL-L-
marine outboard engines and chainsaw motors. The high 23699 (5 cSt oils). (Now MIL-PRF-7808L, MIL-PRF-23699G
biodegradabilities of esters, low ecotoxicity, and clean-burn [119].). 5 cSt oils were preferred for civil aviation turbines
characteristics of ester formulations make them excellent can- and MIL-PRF-23699 became the default specification in the
didates for “environmental considerate” labeling such as Blue civil aviation industry. Subsequently, it was recognized that it
Angel in Germany. Dimerate esters have been widely used was not appropriate for the U.S. military to manage material
[89]. However, despite their high renewable content, dimerate accreditations for a global civil application. Responsibility
esters have relatively poor biodegradability. Therefore, where for civil aviation specifications was been transferred to SAE
biodegradability is an important factor, NPG or TMP-based (Society of Automotive Engineers) and a civil specification
polyols have replaced dimerates [90]. Biodegradable polyol developed as SAE-AS5780 [120].
ester formulations (>70% OECD 301B) for use in chain saw SAE-AS5780 is closely based on MIL-PRF-23699 and
and Jet Ski applications are in commercial use. industry literature still refers to both specifications. MIL-
Low-temperature performance is important in some appli- PRF-23699 recognizes three sub-types of 5 cSt oils; standard
cations, such as engines used to power snowmobile-type vehi- (STD), high thermal stability (HTS), and corrosion inhib-
cles. Therefore, esters with low pour points of down to −56°C ited (C/I). SAE-AS5780 recognizes two classes, Standard
are very suitable for these applications. Performance Class (SPC) and High Performance Class (HPC).
SAE, SPC, and HPC requirements are very similar to require-
3.4.4 Aviation Turbine Oils ments for MIL, STD, and HTS, respectively, although there
are minor updates. (Civil aviation does not generally require
The bulk of aviation lubricant demand is for gas turbine (jet C/I performance. C/I conformance was generally achieved
engine) lubricants (see Gas Turbine Oils). Aviation turbine through formulation rather than basefluid modification.)
oils require an extraordinarily wide fluid temperature range AS5780 SPC products currently represent the large major-
because of the combination of high engine operating tempera- ity of the aviation turbine lubricant market, although other
tures and very low ambient temperature during high altitude grades are produced at significant volume for military avia-
flight. The oil must be stable and involatile enough to pro- tion, spin-off industrial applications (e.g., aero-derivative tur-
vide adequate lubrication and cooling during operation, while bines), and minor civil aviation applications such as auxiliary
remaining fluid at very low temperatures in order to allow power unit oils. AS-5780 HPC class is defined as “Lubricant
in-flight “re-lighting” in the event of temporary engine fail- intended for more demanding service in which engine operat-
ure. Mineral oils were found to be unsuitable although spe- ing conditions and/or service durations require higher thermal
cial naphthenic oils were used for some military applications capability.”
For both SPC and HPC fluids, the specification is based
TABLE 3.18 around test requirements which are divided into:
Typical Ester-Based Two-Stroke Formulation
• Physical Properties
Component % Dose Rate Comments • Chemical Properties
Ester basestock 50–60 Dimerate esters • Stability Properties
NPG and TMP polyol esters • Deposition Properties
PIB 10–15 Low smoke • Tribological Properties
Add pack 10–15
Solvent 15–20 Physical and tribological property requirements for HPC
lubricants are identical to those for SPC. Chemical property
66 Synthetics, Mineral Oils, and Bio-Based Lubricants

requirements are almost identical, but permit slightly more phosphate (TCP) to passivate metal surfaces and prevent cata-
rapid swelling of fluorocarbon elastomers by the HPC fluids. lytic decomposition and also provide some antiwear activity.
The distinction between SPC and HPC fluids is primarily There has been concern that use of TCP in combination with
associated with stability at very high temperatures as mea- fluids based on TMP esters could lead to formation of TMP
sured by change in fluid viscosity, increase in acid number, phosphate (TMP-P) which is neurotoxic [93]. SPC fluids are
and formation of solid sludge or deposits. generally formulated with a mixture of alkylated diphenyl-
The specifications define physical properties and applica- amine and phenyl α-napthylamine antioxidants for high-tem-
tion performance, but do not dictate chemical composition. perature stability. Minor amounts of other additives including
Section 3 of SAE-AS5780 requires only that the basefluid corrosion inhibitors and antifoams may also be present.
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

be based on polyol ester, plus some restrictions on permitted Reclamation of polyol ester aviation lubricants to original
additives. In particular, the standard allows for sale of prod- specification standards has been demonstrated [78] but is not
ucts having different chemical compositions under the same currently practiced.
marketing brand name, as long as each different formulation
has been independently qualified.
3.4.5 Hydraulic Fluids
There are many different combinations of POE raw mate-
rials which can be used to achieve the required physical Hydraulic fluids are used primarily to transmit power but are
properties. The range of raw materials with preferred cost or also required to provide adequate lubrication to moving parts,
availability has changed over time, leading to commercial- e.g., motors and actuators in the hydraulic circuit. In principle,
ization of different product chemistries. Most 5 cSt fluids are the only heat source is from frictional and viscous losses, so
based on PE esters of mixed short chain acids with carbon service temperatures and thermal stresses on the lubricant
numbers in the range C5–C10, [121] but there is some use of are generally lower than for engine oils. The large majority of
TMP esters of linear C7–C10 [122]. Aviation POE basefluids hydraulic fluids are formulated with MO basefluids, with vis-
were originally developed based on tech PE, having a com- cosity in the range ISO 22 to ISO 68, depending on the expected
position close to the crude reaction product. The DiPE and operating temperature. The reason for selection of esters or
other higher polyol components increase product viscosity. other SBF or natural triglycerides in hydraulics are most often
The higher structural diversity also inhibits solidification, so due to the potential impact of leaks from the high-pressure
tech PE esters have higher viscosity (and VI) and lower pour sections. In mobile equipment, these represent environmental
points, than the equivalent mono PE products. Consequently, pollution and biodegradability may be a legal requirement. In
aviation POEs could not easily be reformulated onto straight factory systems, high pressure oil leaks may form oil mists
mono PE and producers have continued to make use of tech which are a flammability hazard and high flash and fire points
PE, although tech PE has increasingly become a niche prod- may be required. These sectors therefore represent the major
uct which is no longer lower cost than pure mono PE. uses of esters in hydraulics. Esters and other SBFs may also
Because they have the same physical property require- be preferred for hydraulic systems operating in extremely low
ments, HPC fluids are based on the same general types of ambient temperatures, where MOs having sufficiently low vis-
ester basefluids as SPC fluids. However, the products and cosity would have unacceptably low flash point [Y2].
manufacturing process are varied using one or more of three
basic approaches to improve oxidative stability; POE struc- 3.4.5.1 Biodegradable Hydraulic Fluids
ture optimization, elimination of unstable impurities, and Biodegradable hydraulic fluids have been a major growth area
optimizing use of additives. for esters over recent decades (see Environmentally Friendly
The different acid raw materials used in POE manufac- Hydraulic Fluids). The major application is in mobile hydrau-
ture vary in inherent oxidative stability depending on their lic equipment used in environmentally sensitive areas, for
structure. POE stability can be improved by restricting the example, in forestry and water-course maintenance. Such
raw materials used to only the most stable acids. For lin- equipment pumps hydraulic fluid at high pressure through
ear acids, stability decreases with increasing chain length. flexible hoses and any damage to the hoses results in loss of
Most branched acids have lower stability, but one particular fluid to the environment.
branched acid, 3,5,5-trimethylhexanoic acid (also known as As noted earlier, some labelling schemes also require that
isononanoic acid, iC9) has particularly high stability. Inclusion the lubricant should have a minimum renewability content
of some iC9 acid in a formulation therefore tends to improve (i.e., to use natural-based feedstocks). Table 3.19 compares
oxidative stability [92]. Also, the branched structure makes a the biodegradability versus renewable content for a range of
higher contribution to viscosity, which allows a reduction in lubricants [94]. It can be clearly seen that esters allow for the
the average carbon number of the overall fatty acid mix. This development of high-biodegradability, renewable, and high-
further improves stability, although at the expense of some performance hydraulic fluids.
decrease in VI and lubricity and relatively higher volatility. There are three market segments for biohydraulics, defined
Use of a mixture of purified mono PE and purified DiPE gives by type of equipment and operating temperature. For low
better thermal stability than tech PE. severity, high loss equipment operating up to 60°C, mainly
Polyol ester turbine oils are always addivated with approxi- farming equipment, vegetable oils may be used. For medium
mately 1–2% of an aromatic phosphate ester, normally tricresyl severity, medium loss applications up to 100°C, mainly in
Esters 67

The alternative route is to redesign the molecular structure


TABLE 3.19 more fundamentally, using neopentyl polyol alcohols and sat-
Comparison of Lubricant Biodegradability and urated acids to optimize thermo-oxidative stability and biode-
Renewability gradability. This approach significantly reduces the renewable
raw material content but has led to development of products
% OECD 301B
Biodegradability % Renewability
combining ready biodegradability with thermo-oxidative sta-
bility, superior to that of mineral oil-based formulations [16].
Vegetable oil 70–100 100
In terms of hydrolytic and oxidative stability, the basefluids
Mineral oil 20–40 0
show the following trend (decreasing order of stability): satu-
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

PAO 20–60 0
rated polyols > diesters > modified TMP oleates > TMP ole-
Alkyl benzene 5–20 0
ates > high oleic sunflower oil > rapeseed.
Diesters 40–80 0–80
Aromatic ester 5–70 0
Table 3.20 gives a typical formulation for an ester-based
Polyol ester 20–99 0–85
biohydraulic fluid. The ester selected will be dependent on
Complex ester 20–90 0–100 the trade-off between cost, renewability, oxidative stability,
Polyalkylene glycol 10–70 0
biodegradability, and low-temperature pumpability.

3.4.5.2 Fire Resistant
forestry operations, synthetic esters with a high content of Ester-based fire-resistant hydraulic oils are classified as
renewable raw materials, for example, TMP oleates are used. HFDU fluids [91]. They compete in this market sector with
High severity applications, particularly in the construction natural triglycerides and with phosphate esters (HFDR flu-
industry, may require fluids capable of extended lifetimes at ids) (see Fire-Resistant Hydraulic Fluids, Neutral Phosphate
operating temperatures in excess of 100°C. Development of Esters). Ester HFDU fluids are normally based on long chain
biodegradable fluids with the necessary high oxidative and polyol esters such as TMP oleates, similar to those used in
thermal stability has been a major challenge for the industry. medium severity biodegradable hydraulic oils [131,132].
It is also essential for hydraulic fluids intended for outdoor Fire resistance of ester HFDUs is due only to their low
use in mobile equipment to possess satisfactory pumpabil- volatility and thermal stability, whereas the phosphorus
ity at the prevailing ambient temperatures during the initial content of phosphate esters further lowers their heat of com-
period of operation with cold fluid. All types of hydraulic flu- bustion, and formation of phosphorus oxide combustion
ids increase in viscosity and eventually solidify as tempera- products inhibits flame spread. Phosphate esters therefore
ture decreases. have superior inherent fire resistance. However, phosphate
An understanding of structure–property relationships has esters generally have inferior VI and pour point, lower ther-
been used to develop two distinct approaches to higher per- mal stability, and are more aggressive to seals and coatings
forming biodegradable hydraulic fluids. The first route is to than triglycerides and polyol esters. Hybrid fluids based on
improve TMP oleate type products by modifying the fatty blends of organic and phosphate esters have been evaluated
acid raw material composition so as to increase the degree [133] but most applications require the performance char-
of saturation, reduce the average alkyl chain length, and acteristics of one type or the other. To enable polyol esters
decrease molecular symmetry. Using this approach, it has to pass certain fire-resistance tests, for example, the factory
been possible to design products containing 85% renewable spray test, 0.5–2% of a VI improver-type polymeric additive
raw materials, but with oxidative stability and low-tempera- may be added. This additive helps modify the spray pattern
ture fluidity greatly superior to standard TMP oleate products. of the fluid pumped through the nozzle by increasing the
Figure 3.9 compares the change in viscosity with time during droplet size. Larger droplets of lubricant are much more dif-
storage at −30°C for standard TMP oleate and new-generation ficult to ignite. A typical HFDU fluid formulation can be seen
product [16,94]. in Table 3.21. [131]

Viscosity vs Storage time at - 30 °C


18000
16000
TABLE 3.20
14000
Viscosity (cSt)

12000 Typical Ester-Based Biodegradable Hydraulic Fluid


10000 TMPO
Modified Component % Dose Rate Comments
8000
6000 Ester basestock >95 TMP oleate ester or modified TMP
4000 oleates or diesters or polyols or
2000
complex polyols
0
0 50 100 150 200 Antioxidant 1–2 Phenolic + aminic
Time (h) Metal deactivator 0.2–0.5 Metal-free
Antiwear 0.4–1.0 Metal-free
FIGURE 3.9  Low-temperature storage stability of standard TMP Antifoam <50 ppm Metal-free, usually non-silicone
oleate vs modified.
68 Synthetics, Mineral Oils, and Bio-Based Lubricants

TABLE 3.21 TABLE 3.22


Typical Ester-Based HFDU Fluid Typical Ester-Based Air Compressor Formulation
Component % Dose Rate Comments Component % Dose Rate Chemistry
Ester basestock >96 NPG, TMP, PE polyol ester (e.g., TMP Ester basestock 5–98 Diester or aromatic esters or polyol
oleate) esters
VI improver 1–2 Droplet modifier Other basestock 0–93 PAO or PAG
Additive pack 1–2 Antiwear, antioxidant, Additive 1–5 Antioxidant, antiwear, anticorrosion,
anticorrosion, etc. antifoam
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

3.4.6 Air Compressor They generally operate in a sealed refrigeration circuit, hav-


ing no exposure to atmospheric oxygen. However, many sys-
The main advantages of esters in air compressor systems are tems, particularly domestic appliances, are sealed-for-life
their high stability, good solvency, and low volatility, which and expected to operate for 10–20 years. Domestic appliance
have important safety implications as well as offering poten- compressors are typically driven by an internal electric motor.
tial for lower total cost operation and cleaner downstream gas Some of the compressor lubricant is inevitably circulated with
for sensitive applications such as breathing air. the working fluid through the hot and cold sections of the cir-
Air compressor oils are required to act as lubricants, cool- cuit and must therefore have good miscibility with the work-
ants, and oil seals at moving parts. For safety reasons, it is ing fluid and good low-temperature fluidity in order to ensure
critically important to minimize formation of carbonaceous adequate oil return. The working fluid condenses at delivery
deposits around the outlet/discharge valves, where high tem- pressure to a low-viscosity liquid which dilutes the lubricant.
perature, high system pressure, and high partial pressure of Therefore, in addition to providing adequate lubrication, the
O2 present a real risk of fire and explosion [95]. Ester oils can compressor lubricant must have outstanding thermal stabil-
be designed for extremely high oxidative stability, and their ity and long-term compatibility with other system materials
solvency characteristics further reduce deposition of sludge including the working fluid, electrical insulation, elastomers
and varnish. etc. (see Refrigeration Lubricants, Compressors, and Pumps,
Selection of high oxidative stability basefluids has a sec- Natural and Process Compressors).
ondary benefit to the user in terms of extended drain intervals Polyol esters were long known to be suitable for use as
(by as much as 8- to 10-fold) and improved energy efficiency, refrigeration compressor lubricants [123], having lower ten-
as well as other associated cost savings, such as less frequent dency to floc formation at low temperature than napthenic
replacements of oil filters and air/oil separators [96]. mineral oils. However, they were not widely used prior to the
The low volatility of esters compared to mineral oils of phase out of chlorofluorocarbon (CFC) refrigerant gases. The
the same viscosity also results in reduced carryover of lubri- hydrofluorocarbon (HFC) refrigerants which replaced CFCs
cant into the gas. These properties, coupled with their higher in many applications because of their lower Ozone Depletion
flash and autoignition temperatures, and low order of toxicity Potential (ODP), also have a much higher polarity and tra-
for vapor inhalation and ingestion, make them considerably ditional refrigeration lubricants based on naphthenic mineral
safer lubricants to use than mineral oil. Their low ecotoxicity oils were found to have inadequate miscibility. Polyol esters
and high biodegradability can also lessen their environmental have become the most widely used basefluid for use with
impact [97]. most HFC gases in stationary applications, with polyvinyl
Because of their high solvency, aromatic esters and partic- ethers (PVEs) as the only real alternative [124]. Mobile air-
ularly trimellitate esters are widely used in air compressors. conditioning systems are typically mechanically driven and
Mixed diester/polyalkylene glycol systems can also be used, therefore do not require high dielectric strength and low con-
exploiting the clean-burn characteristics of the PAG [98]. ductivity in the lubricant. End-capped PAGs are more com-
Because of the high polarity of these systems, care is required monly used in these systems (see Polyalkylene Glycols) [99].
in the specification of elastomeric sealing materials: the use of A very wide range of compressor types including recipro-
Viton is usually recommended. cating, rotary, and screw compressors having widely differ-
A typical ester-based air compressor formulation can be ent lubricant requirements are used in refrigeration systems.
seen in Table 3.22. Together with the possibility for lubricant dilution by refriger-
ant gas, this requires a correspondingly broad range of fluid
viscosities. POE refrigeration lubricants are commercially
3.4.7 Refrigeration Compressor available in viscosity grades from ISO 7 to ISO 320.
A major growth sector for synthetic ester basefluids during Low-viscosity oils (<ISO 10) are based on NPG polyols;
the last two decades has been as HFC-compatible refrigera- high-viscosity oils (>ISO 150) are based on DiPE polyols or
tion compressor lubricants. Refrigeration compressor lubri- complex esters. Intermediate viscosity products are based on
cants present a range of unique performance requirements. pure PE polyols themselves or blended with NPG or DiPE
Esters 69

TABLE 3.23 TABLE 3.24


Typical Polyol Ester-Based Refrigeration Lubricant Typical Ester-Based Chain Oil Formulation
Component % Dose Rate Chemistry Component % Dose Rate Comments
Ester >95 Polyol ester (e.g., NPG, PE, DiPE) Ester >75 Dimerate
Antioxidant 0–1000 ppm Storage stabilizer Trimellitate ester
Antiwear 0–5 To reduce wear Polyol ester
Foaming agent 0–1000 ppm To reduce noise Tackifier/thickener 2–20 VI improver (PIB, PMMA)
polymeric esters
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

Additive pack 2–8 Antioxidant, antiwear, anticorrosion

polyols. TMP polyols have not been widely used because of


concerns over potential reaction with phosphate ester addi-
tives under fire conditions [92]. The acids used are generally esters tend to be favored [100]. Trimellitates tend to have
mixed C5–C8 linear or C8–C9 branched, with ester product a poorer stability than POEs but decompose to leave a soft
compositions tailored to ensure optimum compatibility with (easy to remove) deposit. Polyol esters decompose at higher
the gas [125]. temperature but can leave behind a harder deposit (varnish).
Because the fluid has no contact with atmosphere, high Often blends of trimellitates and polyols are used to obtain the
levels of antioxidant additives are not required, although low correct balance.
levels of phenolic antioxidant are normally used to ensure VI improvers are often used as tackifiers/thickeners. For
stability in storage. Aromatic phosphate esters are often very oxidative stable formulations, high-viscosity polymeric
used as antiwear agents. Refrigeration lubricants are manu- esters have been used as a tackifier.
factured to very low moisture content to prevent ice forma-
tion in cold sections, so hydrolytic stability is not a problem.
3.4.9 Metalworking Fluids
However, some manufacturers use epoxide acid catcher addi-
tives as a precautionary measure to control any increase in Metalworking fluids represent a highly diverse and frag-
acid value. Gas bubbles in the fluid have a noise-damping mented sector including high- and low-temperature rolling
effect and foam control agents are sometimes added to con- oils, coolants for cutting and drilling operations, and lubri-
trol compressor noise. A typical ester-based formulation can cants for stamping and forming operations (see Metalworking
be seen in Table 3.23. Fluids) as well as quenching oils used for controlled cooling
The refrigeration compressor lubricant market continues of castings [127]. The contact conditions normally involve
to evolve as regulatory attention has switched to minimiz- intentional deformation of the metal and generation of new
ing Global Warming Potential (GWP) as well as ODP. This metal surface and are therefore very different from the rolling
is likely to involve progressive phase out of R134a and blends and sliding contacts found in the majority of other applica-
over the coming years. Where the replacement gases are also tions. Esters can be used either as additives (typically used at
HFCs, for example R32 replacing R410A in air-conditioning, a treat-rate of 5%–15%) or as the base oil. The primary reason
it is likely that POEs will remain the preferred basefluid [126]. for using an ester as additive is to reduce the friction between
Esters may also be used as compressor lubricants in other tool and work piece, with the specific aims of improving sur-
non-air applications, such as gas transmission and chemi- face finish and/or extending tool life. Their use as base oils
cal plant process gas systems, wherever their compatibility is usually the result of some additional requirement, such as
(polarity, solvency, and reactivity) with the compressant is higher lubricity, minimizing misting, or a desire for a control-
better than that of mineral oils or other synthetics. lable level of biodegradability [134]. The use of esters is set
to increase as machining techniques develop and as greater
consideration is given to the environmental and health aspects
3.4.8 High-Temperature Chain Oils
of metalworking formulations.
Many manufacturing products today require extreme heat, A wide range of esters are used in neat oils, including
either in the manufacturing, finishing, curing, or drying pro- monoesters (e.g., methyl oleate or isopropyl palmitate), dies-
cess. Application areas such as textile factories, car plants, ters (e.g., propylene glycol dioleate), and polyol esters (e.g.,
and pottery/glass kilns use roller chains, stenter chains, and TMP trioleate).
sliding chains (see High-Temperature Chain Oils). Lubricants In water-miscible fluids, the most commonly used esters
for these chains see temperatures above 150°C, and some- are isopropyl oleate, iso-butyl stearate, neopentylglycol diole-
times as high as 1000°C. Esters are required that have high ate, and a number of TMP derivatives. A common feature of
oxidative stabilities, low volatilities, and excellent clean-burn the esters used in water-miscible formulations is their greater
properties. A typical ester-based chain oil formulation can be resistance to hydrolysis as compared to methyl esters.
seen in Table 3.24. Natural oils and fats, such as coconut and palm oils, and
Dimerates tend to be used in cost-effective formulations. synthetic esters, for example, esters of NPG, TMP, and PE,
For higher temperatures, trimellitates and PE/diPE polyol are widely used in many rolling formulations [135]. Complex
70 Synthetics, Mineral Oils, and Bio-Based Lubricants

esters have been evaluated for minimum quantity lubrication


in cutting oils, [101] and may find application in formulations TABLE 3.25
developed to substitute for medium chain chlorinated paraf- Typical Ester-Based Drilling Lubricant Formulation
fins [135], either alone or in combination with EP additives, Component % Dose Rate Chemistry
such as phosphate esters and sulfurized esters or olefins.
Ester 20–30 Fatty ester
Water 10–20
3.4.10 Greases Calcium chloride brine 3–8 Alkalinity
Viscosifier 0.5–3
Esters are commonly used as basestocks for greases when one
Primary emulsifier 0.5–3
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

or more of the following properties are required:


Secondary emulsifiers/wetting agent 0.25–1
Fluid loss control 0.25–1
Low-temperature flow (e.g., aircraft wheel bearings) Lime 0.5–2 Alkalinity
High-temperature applications Weighting agent 40–60 e.g., barite or clay
Biodegradability
Low toxicity (e.g., food-use applications)

A range of ester types (diesters, phthalates, trimellitates, 3.4.12 Transformer/Dielectric Fluids


pyromellitates, and polyols) are used in this application (see
Synthetic Greases). Esters can be used in combination with Electrically insulating dielectric fluids are used to provide
traditional carboxylate soap gellant technology, although cooling in high power transformers. Operating temperatures
there is a possibility of some ester saponification occurring, are generally relatively low, but very long service lifetimes
which may modify the properties. Ester basefluids are more (10–20 years) are expected. As with hydraulic fluids, the
commonly used in combination with nonstandard thickener majority of the market is mineral oil based, but there is a large
technologies such as polyureas [102]. sub-segment where fire resistance is necessary. Historically,
polychlorinated biphenyls (PCBs) were used in this applica-
tion and a market developed for other SBFs including esters
3.4.11 Drilling Muds when use of PCBs was prohibited. Short chain pentaerythritol
Ester-based organic compounds are one type of SBF added esters were found to be suitable and have taken and main-
to drilling muds used during offshore oil-drilling operations. tained a significant market share [104,105]. The ester fluids
Since 1990, the oil and gas extraction industry developed used are PE esters of mixed short chain fatty acids, similar
SBFs with synthetic and nonsynthetic oil-like materials as to those used in aviation turbine oils. In addition to fire resis-
the basefluid to provide the drilling performance charac- tance, secondary benefits of polyol esters include higher water
teristics of traditional oil-based fluids (OBFs) based on die- solubility, which extends the lifetime of cellulose-based solid
sel and mineral oil. Drilling fluids are needed to cool and insulator components, and biodegradability in the event of
lubricate the drill bit and to help bring rock cuttings to the spillage. The products are typically stabilized with low levels
surface. of phenolic antioxidant and corrosion inhibitor additives.
A long-standing industry standard specification for syn-
Ester-based drilling fluids have the following advan- thetic ester dielectric fluids specifies use of pentaerythritol
tages over OBFs: esters [129] and includes requirements for viscosity, pour
Faster and deeper drilling point, and flash point which effectively constrain the range
Greater worker safety through lower toxicity of suitable acid raw materials to mixed linear/branched C7–
Elimination of polynuclear aromatic hydrocarbons C10 acids. As for appliance refrigeration lubricants, appli-
(PAHs) cation-specific performance requirements for transformer
Excellent biodegradability and lower bioaccumulation fluids include very high electrical resistivity and dielectric
potential strength (breakdown voltage) and low dielectric dissipation
Potentially less drilling waste volume factor. These parameters are mainly controlled by impuri-
Reduced drilling costs ties rather than bulk composition and require low acid and
hydroxyl value as well as high levels of fluid cleanliness since
Government and industry research found that several syn- particulate contamination can provide a pathway for arcing.
thetic-based fluids used in mud formulations exhibited similar Dissipation factor can be impacted by traces of other ester
biodegradation profiles to mineral oils offering no apparent lubricant basefluid types, so control of batch to batch cross-
benefits. As a result, the UK government decided to reduce contamination is critically important where ester dielectric
the discharge of a number of synthetic fluids. Esters were fluids are manufactured on multi-product plants. Dielectric
not subjected to the same reduction program because of their fluids are normally supplied degassed and are vacuum-treated
rapid biodegradation [103] and continue to find application to remove dissolved gases before packing.
[128]. A typical ester-based drilling formulation can be seen More recently, highly purified natural triglycerides simi-
in Table 3.25. lar to those used in HETG hydraulics have also been found
Esters 71

TABLE 3.26
Ester Lubricant Manufacturers
Tradenames (Former
Company (Former Business Names) Tradenames) Manufacturing Region Main Product Types
Adeka Corporation Adeka™ Asia Polyol Ester
BASF (Cognis) Synative™ ES, Glissofluid™ Asia, Europe, Americas Polyol Esters, Diesters
Cargill Envirotemp Americas Oleochemical
Croda (Uniqema) Priolube™, Perfad™ Europe, Americas, Asia Polyol Esters, Diesters, Oleochemical
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

EmeryOleo (Cognis) DeHylub™ Americas, Europe Oleochemical


ExxonMobil Esterex™ Americas, Europe Polyol Esters, Diesters
Hercules (now Calumet Speciality Calester™ (Hercolube™) Americas Polyol Esters
Products)
Industria Quimica Lasem Docoil™ Europe Diesters, Oleochemicals
Italmatch (Undesa) Ketjenlube™ Europe Diesters, Oleochemicals
Kao Kaolube™ Asia Oleochemicals
KLK Oleo Palmester™ Asia Oleochemical
Lanxess (Chemtura, Hatco Corporation) Hatcol™ Americas Polyol Esters, Diesters
M&I Materials Midel™ Europe Polyol Ester
NOF Corporation Unister™ Asia Polyol Esters
Nyco Nycolube™, Nycobase™ Europe Polyol Esters, Diesters
Oleon Radialube™ Europe, Asia Oleochemical
Oxea Oxlube™ Europe, Asia Polyol Esters, Diesters
Patech Fine Chemicals PALUB™ Asia Oleochemical
Polyone SynPrime™ Americas Diesters
Zschimmer & Schwartz (Inolex) Lexolube™ Americas Polyol Esters

to be suitable as fire-resistant and biodegradable transformer Italmatch acquired Union Derivan (Undesa) in 2013.
fluids [130] and these products now represent a higher overall The group is mainly focussed on Polymer Esters
volume than the longer established synthetics. However, their and additive chemistries, but also offer some diester
lower oxidative stability and relatively high pour points mean (Ketjenlube) and oleate (Dapralube) basefluids.
that they are mainly used in sealed transformers, taking mar- KLK Oleo commissioned new ester capacity for a range
ket share from MO rather than synthetic esters. including lubricant basefluids at their site in Klang,
Malaysia, during 2017.
Oxea have commissioned a new specialty ester plant in
3.5 MARKETING AND ECONOMICS
Nanjing, China, for a product range including polyol
3.5.1 Manufacturers esters which will be marketed in refrigeration com-
pressor applications by Lubrizol.
A list of companies who both produce and market synthetic Lanxess acquired Chemtura in 2017. The Hatco divi-
ester basefluids, the relevant trade names, and main product sion including the Hatcol product range is now part
types (diesters, short chain polyol esters, oleochemicals) are of Lanxess lubricant additives business unit.
given in Table 3.26. Table 3.26 does not include companies Emery Oleo (formerly the upstream raw materials
who produce esters for internal use only. business of Cognis) have introduced a lubricant
Significant changes in recent years include the following: ester product range based on their captive natural
fatty acid and dimer acid raw materials, branded as
Nyco increased ester capacity at Tournai, Belgium, in Dehylub. Esterification capacity has been developed
2012 and commissioned a plant manufacturing anti- or uprated at their production sites in Cincinnati,
oxidant additives optimized for performance with OH, Loxstedt, Germany, and Telok, Malaysia.
ester fluids in 2016. Calumet announced doubling of ester capacity at their
Industria Quimica Lasem were acquired by Nisshin Louisiana, MO, plant in 2015. The former Hercolube
OilliO Group in 2011, and continue to offer a range products have been rebadged as Calester. In 2018,
of ester basefluids under the Docoil tradename. Calumet acquired Biosynthetic Technologies includ-
Oleon closed their ester plant in Sandefjord, Norway, ing intellectual property rights to estolide technology
during 2014, with products transferred to other assets and will transfer production activities to the MO site.
in Europe and Asia.
72 Synthetics, Mineral Oils, and Bio-Based Lubricants

Zschimmer & Schwartz acquired the Inolex Lexolube in use were first developed and commercialized, followed by
ester business in 2017. Production will be transferred a much slower process of progressive spread of the technol-
to Milledgeville, GA. ogy across all areas of the lubricant industry, with some major
Canadian plasticizer producer PolyOne have entered increases in demand triggered by legislative changes.
the lubricant market with a range of diester fluids Overall, this pattern looks likely to continue for the fore-
under the Synprime tradename. seeable future. The range of potential raw materials offered
ExxonMobil closed the former Mobil ester plant in by the petrochemical and traditional oleochemical industry is
Edison, NJ, and transferred production to assets in TX. mature and stabilized by established high levels of demand.
Raw material innovation is therefore likely to continue to
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

focus on optimizing production economics for current prod-


3.5.2 Markets
ucts rather than introducing new substances.
The overall size of the ester lubricant market is very difficult At the esterification level, although there is still scope for
to define, both in terms of producer capacity and market sales innovation through combining the available raw material slate
volume. in different ways, the broad range of existing materials and
The majority of ester basefluid products are manufactured high costs of global notification for new substances will tend
on multipurpose production assets by producers who also to direct attention toward optimizing specifications and find-
manufacture for other application areas (e.g., surfactants, ing new applications for existing products.
coatings, plasticizers, biofuels, and fuel additives). Production However, for the lubricant formulator and end user, market
rates may be significantly different for different ester prod- requirements to optimize lubricity, durability, and environmen-
ucts manufactured on the same assets, so actual capacity will tal acceptability of lubricant formulations will continue to drive
depend on product mix. Declared plant capacity is therefore progressive replacement of mineral oil formulations with higher
not a good guide, even where it is available, complete, and performing synthetics. Because of their extreme structural flex-
correct. ibility and environmental performance, esters are expected to
Similarly, many of the materials marketed as ester lubri- take an increasing share of this increasing overall demand for
cant basefluids also have major applications in other market synthetics in all lubricant sectors. Some specific areas expected
segments, although typically with different specification to see further demand are listed in the following.
requirements. Some lubricant formulators may manufacture In the major volume areas of automotive crankcase and
their own basefluids internally. Estimates for total production gear oils, there will be increasing competition from electric
volumes of specific substances therefore do not necessarily vehicle technology. The threat of large scale substitution will
relate to lubricant consumption. make OEMs and formulators more reluctant to invest in long-
Published market estimates for total lubricant volume typi- term R&D directed to radical changes in internal combustion
cally do not differentiate between specific basefluid types and engine technology. However, internal combustion engines
are inconsistent over whether some major functional fluid will retain the large majority of the market for at least the
applications (e.g., transformer oils, hydraulic oils) are included next decade and will continue to be subject to increasing leg-
or not. Some sources may include solid wax ester lubricants islative pressure on fuel economy and emissions standards,
used as polymer processing additives. as well as economic drivers to increase lifetime and service
Approximate annual volume and growth rate for the intervals. This will continue the current trends to incremental
product types and market sectors discussed in this chapter substitution of traditional MO-based formulations by higher
are shown in Table 3.27. (These values are author estimates performing synthetics and semi-synthetics based on PAO,
drawn from a range of sources.) Group III+, and GTL fluids and will drive further demand for
esters as polarity modifiers. In industrial sectors, demand for
polyol esters for HFC-compatible refrigeration lubricants will
3.6 OUTLOOK
initially level off as the industrial refrigeration sector com-
The history of synthetic ester lubricants has been character- pletes the switchover. The overall refrigeration lubricant mar-
ized by an initial phase of very rapid product discovery and ket will continue to grow as economic growth in Asia drives
development when essentially all of the materials currently air-conditioning demand. However, market volume for esters
or other SBFs will continue to depend on legislation relat-
ing to the refrigerant gases and in particular the adoption of
TABLE 3.27 fluorinated refrigerants having low GWP. Meanwhile biode-
Market Volumes and Growth Rates for Ester gradable lubricants will be increasingly widely specified for
potentially dispersive industrial applications in shipping and
Lubricant Basefluids (2017 Basis)
mobile hydraulic. The high level of current interest in natural
Region Annual Volume (ktes) Annual Growth Rate (%) and bio-based lubricants will result in opportunities for syn-
EMEA 150 3 thetic esters as higher performing alternatives to natural oils.
Americas 135 5 Sustained development and growth of synthetic ester base-
Asia 85 5 fluid chemistry can therefore be expected to continue over the
coming decade.
Esters 73

ACKNOWLEDGMENTS 17. Martin, J., 2006, Price considerations for use of soy feedstocks
for the chemicals industry, Ind. Biotechnol., 2(2), 92–95.
The author would like to thank Steven Randles for his origi- 18. Yeong, S. K., Ooi, T. L., and Salmiah, A., 2005, Palm-based
nal chapter in the first edition of this book [106] and earlier hydraulic fluid, Malaysian Palm Oil Board, MPOB Information
works [107], from which this chapter has been expanded, Series, MPOB TT No. 281, ISSN:1511–7871, June 2005.
updated, and revised [108–120]. Thanks are also due to the 19. Erhan, S. Z., Sharma, B. K., and Perez, J. M., 2006, Oxidation
and low temperature stability of vegetable oil-based lubri-
many industry contacts who have provided useful information
cants, Ind. Crop. Prod., 24, 292–299.
or comment on the content of this chapter [121–135]. 20. Fox, N. J. and Stachowiak, G. W., 2007, Vegetable oil-based
lubricants—A review of oxidation, Tribol. Int., 40, 1035–1046.
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

REFERENCES 21. EU Ecolabel for Lubricants, 2005 (Decision 2005/360/EC),


http:​//eur​-lex.​europ​a.eu/​L exUr​iServ​/site​/en/o​j/200​5/l_1​18/1_​
1. Kenbeek, D. and Buenemann, T., 2017, Chapter 5: Organic 11820​05050​5en00​26003​4.pdf​, Accessed 17 July, 2012.
friction modifiers. In Lubricant Additives: Chemistry and 22. USDA, 2010, BioPreferred Program Product Categories—
Applications, 3rd edn., ed. Rudnick, L. R., CRC Press, p. 38. October 2010, www.b​iopre​ferre​d.gov​/file​s/Bio​Prefe​r red_​
2. Davison, C. St. C., 1957, Wear prevention, 1400 BC–14 AD, produ​ct_ca​tegor​ies_O​ctobe​r_201​0_FIN​A L.pd​f.
Wear, 1, 155–159. 23. Boyde, S., 2002, High performance lubricants from renew-
3. Archbutt, L. and Deeley, R. M., 1927, Lubrication and ables. In Lipid Technology Newsletter, PJ Barnes & Associates,
Lubricants, 5th edn., Charles Griffin & Co., Ltd., London, June 2002, pp. 61–65.
U.K. 24. Johnson, R. W. and Fritz, E., 1989, Fatty Acids in Industry:
4. Sears, J. K. and Darby, J. R., 1982, The Technology of Processes, Properties, Derivatives, Applications, Marcel
Plasticizers, John Wiley, New York, NY, pp. 1–34. Dekker Inc., New York, NY.
5. Zisman, W. A., 1962, Historical review, lubricants and lubrica- 25. Kubitsche, J., Lange, H., and Strutz, H., 2014, Carboxylic Acids,
tion. In Synthetic Lubricants, eds. Gunderson, R. C. and Hart, Aliphatic, Ullmann’s Encyclopedia of Industrial Chemistry,
A. W., Reinhold Publishing Corp., New York, NY, pp. 6–60. Wiley-VCH Verlag GmbH & Co., KGaA, Published Online:
6. Dudek, W. G. and Popkin, A. H., 1962, Dibasic acid esters. In May 30, 2014, doi:10.1002/14356007.a05_235.pub2.
Synthetic Lubricants, eds. Gunderson, R. C. and Hart, A. W., 26. Johnson, R. W., Pollock, C. M., and Cantrell, R. C., 2010,
Reinhold Publishing Corp., New York, NY, pp. 151–241. Dicarboxylic acids. In Kirk-Othmer Encyclopedia of
7. (a) Jantzen, E., 1996, The origins of synthetic lubricants: Chemical Technology, John Wiley & Sons, Inc., New York,
The work of Hermann Zorn in Germany. Part 2: Esters and NY, Published Online: September 17, 2010, doi: 0.100​2/047​
additives for synthetic lubricants, J. Synth. Lubr., 13(2), 113– 12389​61.04​09030​11015​0814.​a01.p​ub2.
128. (b) Zorn, H., 1947, Esters as lubricants, US Air Force 27. Park, C.-M., Sheehan, R. J., 2000, Phthalic acids and other
Translation Report F-TS-957-RE. (c) Report PB-110034, benzenepolycarboxylic acids. In Kirk-Othmer Encyclopedia
Tables of Physical Characteristics of a Wide Range of Esters, of Chemical Technology, John Wiley & Sons, Inc., New York,
I.G. Farbenindustrie, Library of Congress. NY, Published Online: December 4, 2000, doi:1​0.100​2/047​
8. Spaght, M. E., 1945, The Manufacture and Application of 12389​61.16​08200​81601​1811.​a01.
Lubricants in Germany, Combined Intelligence Objectives 28. Komp, H.-D. and Kubersky, H. P., 1982, Technical processes
Sub-Committee, Nav Tec Miseu, CIOS TARGET NO. 30.303, for the manufacture of fatty alcohols. In Fatty Alcohols,
Fuels and Lubricants, www.f​i sche​r-tro​psch.​org/p​r imar​y_doc​ Raw Materials, Methods, Uses, Henkel KGaA, Dusseldorf,
ument​s/gvt​_ repo​r ts/C​IOSC/​cios_​30_32​_68.h​tm, Accessed 17 Germany, pp. 49–74.
July, 2012. 29. (a) Falbe, J., Bahrmann, H., Lipps, W., Mayer, D., and
9. Cohen, G., Murphy, C. M., O’Rear, J. G., Ravner, H., and Frey, G. D., 2002, Alcohols, Aliphatic, Ullmann’s
Zisman, W. A., 1953, Aliphatic esters, Ind. Eng. Chem., 45, Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag
1766. GmbH & Co., KGaA, Published Online: January 15, 2013,
10. McTurk, W. E., 1953, Synthetic lubricants. WADC Technical doi:10.1002/14356007.a01_279.pub230. (b) Hunter, W. N.,
Report 53–88, Contract AF 33(038)-14593, RDO No. 613–15, 2000, Alcohols, Polyhydric, Kirk-Othmer Encyclopedia of
October 1953. Chemical Technology, John Wiley & Sons, Inc., New York,
11. Barnes, R. S. and Fainman, M. Z., 1957, Synthetic ester lubri- NY, Published Online: December 4, 2000, doi:1​0.100​2/047​
cants, Lubr. Eng., 13, 454–458. 12389​61.01​12031​50821​1420.​a01.
12. Smith, T. G., 1962, Neopentyl polyol esters. In Synthetic 30. Schlosberg, R. H., Chu, J. W., Knudsen, G. A., Suciu, E. N.,
Lubricants, eds. Gunderson, R. C. and Hart, A. W., Reinhold and Aldrich, H. S., 2001, High stability esters for synthetic
Publishing Corp., New York, NY, pp. 388–401. lubricant applications, Lubr. Eng., 57(2), 21–26.
13. Aslam, M., Torrence, G. P., and Zey, E. G., 2000, Esterification. 31. Briant, J., Denis, J., and Parc, G., 1989, Rheological Properties
In Kirk-Othmer Encyclopedia of Chemical Technology, John of Lubricants, Editions Tecnip, Paris, France, pp. 123–163.
Wiley & Sons Inc., New York, NY, Published Online: December 32. van Os, N. M. (Ed.), 1998, Nonionic Surfactants: Organic
4, 2000, doi:1​0.100​2/047​12389​61.05​19200​50119​1201.​a01. Chemistry, Marcel Dekker, New York, NY.
14. Murphy, C. M. and Zisman, W. A., 1950, Structural guides 33. (a) Niedzielski, E. L., 1976, Neopentyl polyol ester lubri-
for synthetic lubricant development, Ind. Eng. Chem., 42, cants—Bulk property optimisation, Ind. Eng. Chem. Prod.
2415–2420. Res. Dev., 15, 54–58. (b) Niedzielski, E. L., 1977, Neopentyl
15. Bartz, W. J., 1997, Lubricants and the environment. In New polyol ester lubricants—Boundary composition limits, Ind.
Directions in Tribology, ed. Hutchings, I. M., IMechE, Eng. Chem. Prod. Res. Dev., 16, 300–305.
London, U.K., p. 103. 34. Solomon, D. H., 1972, Polyesterification. In Step-Growth
16. Boyde, S., 2002, Green lubricants. Environmental benefits and Polymerisation, ed. Solomon, D. H., Marcel Dekker, New
impacts of lubrication, Green Chem., 4, 293–307, doi:10.1039/ York, NY, pp. 1–41.
B202272A.
74 Synthetics, Mineral Oils, and Bio-Based Lubricants

35. Furey, M. J., 1973, The formation of polymer films directly on 55. Jensen, R. K., Korcek, S., Zinbo, M., and Johnson, M. D.,
rubbing surfaces to prevent wear, Wear, 26, 369–392. 1990, Initiation in hydrocarbon autoxidation at elevated tem-
36. Reck, R. A., 1989, Polyoxyethylene esters of fatty acids. In peratures, Int. J. Chem. Kinet., 22(10), 1095–1107.
Fatty Acids in Industry: Processes, Properties, Derivatives, 56. Gunstone, F., 1996, Fatty Acid and Lipid Chemistry, Blackie
Applications, eds. Johnson, R. W. and Fritz, E., Marcel Academic and Professional, London, U.K., pp. 162–179.
Dekker Inc., New York, NY, pp. 201–214. 57. Erhan, S. Z. and Asadauskas, S., 2000, Lubricant basestocks
37. Mathiesen, T. and Jensen, J. W., 1993, Self-emulsifying ester from vegetable oils, Ind. Crops Prod., 11, 277–282.
compounds, US 5219479. 58. Hunter, M., Klaus, E. E., and Duda, J. L., 1993, A kinetic
38. Boydes, S., 2001, Low temperature characteristics of synthetic study of antioxidation mechanisms, Lubr. Eng., 49, 492.
fluids, Synth. Lubr., 18, 99. 59. Bakunin, V. N. and Parenago, O. P., 1992, A mechanism of
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

39. Hamrock, B. J., Schmid, S. R., and Jacobson, B. O., 2004, thermo-oxidative degradation of polyol ester lubricants, J.
Fundamentals of Fluid Film Lubrication, 2nd edn., Marcel Synth. Lubr., 9, 127.
Dekker, New York, NY. 60. Hamblin, P., 1999, Oxidative stabilisation of synthetic fluids
40. Chang, H. S., Spikes, H. A., and Buenemann, T. F., 1991, The and vegetable oils, J. Synth. Lubr., 16, 157–181.
shear stress properties of ester lubricants in elastohydrody- 61. (a) Schick, J. W. and Kaminski, J. M., 1979, Synthetic ester and
namic contacts, J. Synth. Lubr., 9(2), 91–114. hydrogenated olefin oligomer lubricant and method of reduc-
41. Anderin, M., Johnston, G. J., Spikes, H. A., and Caporiccio, ing fuel consumption therewith, US 4175047. (b) Aldrich, H.
G., 1992, The elastohydrodynamic properties of some S., Gordon, F. H., Holt, G. L., Krevalis, M. A., Leta, D. P.,
advanced non hydrocarbon-based lubricants, Lubr. Eng., Schlosberg, R. L. H., Sherwood-Williams, L. D., and Szobota,
48(8), 633–638. J. S., 1996, Polyol ester compositions with unconverted
42. Gunsel, S., Spikes, H. A., and Anderin, M., 1993, Tribol. hydroxy groups for lubricants, US 5665686.
Trans., 36(2), 276–282. 62. Sommers, E. A. and Crowell, T. L., 1953, High tempera-
43. Guangteng, G. and Spikes, H. A., 1997, The control of fric- ture anti-oxidants for synthetic base oils, Part 3. The ther-
tion by molecular fractionation of base fluid mixtures at metal mal decomposition of Di (2-ethyl hexyl) sebacate, W.A.D.C
surfaces, Tribol. Trans., 40, 461–469. Technical Report, pp. 53–293.
44. Smeeth, M., Spikes, H. A., and Gunsel, S., 1996, The forma- 63. Lansdown, A. R., 1994, High Temperature Lubrication,
tion of viscous surface films by polymer solutions: Boundary Mechanical Engineering Publications Limited, London, U.K.,
or elastrodynamic lubrication? Tribol. Trans., 39, 720–725. ISBN:0 085298 897 464.
45. Boyde, S., Randles, S. J., and Gibb, P., 1999, The effect of 64. Blake, E. S., Hammann, W. C., Edwards, J. W., Reichard, T.
molecular structure on boundary and mixed lubrication by E., and Ort, M. R., 1961, Thermal stability as a function of
synthetic fluids—An overview. In Lubrication at the Frontier, chemical structure, J. Chem. Eng. Data, 6, 87–98.
eds. Dowson, D. et al., Elsevier, Amsterdam, The Netherlands, 65. (a) Cottington, R. I. and Ravner, H., 1968, Neopentyl polyol
p. 799. esters for jet engine lubricants—Effects of tricresyl phos-
46. Dorinson, A. and Ludema, K. C., 1985, Mechanics and phate on thermal stability and corrosivity, Naval Research
Chemistry in Lubrication, Elsevier, Amsterdam, The Laboratory Report No. 6667, http:​//tor​pedo.​nrl.n​avy.m​il/tu​/
Netherlands, pp. 198–252. ps/p​df/pd​f-loa​der?d​sn=28​80830​, Accessed July 17, 2012. (b)
47. Bovington, C. H., 2010, Friction, wear and the role of addi- Cottington, R. L. and Ravner, H., 1969, Interactions in neo-
tives. In Chemistry and Technology of Lubricants, 2nd edn., pentyl polyol ester – Tricresyl phosphate-iron systems at 500
eds. Mortier, R. M., Fox, M., and Orzulik, S. T., Springer, pp. F, ASLE Trans., 12, 280–286.
320–347. 66. Lockwood, F. E. and Klaus, E. E., 1981, Ester oxidation under
48. Van der Waal, G., 1985, The relationship between chemical simulated boundary lubrication conditions, ASLE Trans, 24,
structure of ester base fluids and their influence on elastomers 276–284.
seals and wear characteristics, J. Synth. Lubr., 1(4), 281. 67. Lahijani, J., Lockwood, F. E., and Klaus, E. E., 1980, The
49. Hansen, C. M., 2000, Hansen Solubility Parameters, A Users influence of metals on sludge, ASLE Trans., 25(1), 25–32.
Handbook, CRC Press, Boca Raton, FL, pp. 1–24. 68. Boyde, S., 2000, Hydrolytic stability of synthetic ester lubri-
50. (a) Sniegoski, P. J., 1976, Selectivity of the oxidative attack cants, J. Synth. Lubr., 16, 297–312.
on a model ester lubricant, ASLE Trans., 20(4), 282–286. (b) 69. Euranto, E. K., 1969, Esterification and ester hydrolysis. In
Rigg, M. W. and Gisser, H., 1953, Autoxidation of the satu- The Chemistry of Carboxylic Acids and Esters, ed. Patai, S.,
rated aliphatic diesters, J. Am. Chem. Soc., 75, 1415–1420. Interscience-Publishers, London, U.K.
51. Jensen, R. K., Korcek, S., and Zinbo, M., 1984, Oxidation 70. Murphy, C. M., Ravner, H., and Timmons, C. O., 1961, Factors
and inhibition of pentaerythritol esters, J. Synth. Lubr., 1(2), influencing the lead corrosivity of ester oils during long term
91–105. storage, J. Chem. Eng. Data, 6(1), 135–141.
52. Lindsay Smith, J. R., Nagatomi, E., and Waddington, D. J., 71. (a) Ratledge, C., 1992, Biodegradation and biotransformations of
2000, The autoxidation of aliphatic esters. Part 2. The autoxi- oils and fats – Introduction, J. Chem. Technol. Biotechnol., 55,
dation of neopentyl esters, J. Chem. Soc. Perkin Trans., 2, 2248. 397–398. (b) Haigh, S. D., 1992, Biodegradation of synthetic lubri-
53. Karis, T. E., Miller, J. L., Hunziker, H. E., De Vries, M. S., cants in soil, J. Chem. Technol. Biotechnol., 55, 406–409. (c) Cain,
Hopper, D. A., and Nagaraj, H. S., 1999, Oxidation chemis- R. B., 1992, Designing biodegradable molecules of the oleochem-
try of a pentaerythritol tetraester oil, Tribol. Trans., 42(3), ical industries, J. Chem. Technol. Biotechnol., 55, 411–414.
431–442. 72. Wright, M. A., Taylor, F., Randles, S. J., Brown, D. E., and
54. Rasberger, M., 1997, Oxidative degradation and stabilisation Higgins, I. J., 1993, Biodegradation of a synthetic lubricant
of mineral oil based lubricants. In Chemistry and Technology by micrococcus roseus, Appl. Environ. Microbiol., 59(4),
of Lubricants, 2nd edn., eds. Mortier, R. M. and Orzulik, S. 1072–1076.
T., Blackie Academic and Professional, London, U.K., pp. 73. Wyatt, J. M., Cain, R. B., and Higgins, I. J., 1987, Formation
98–143. from synthetic two-stroke lubricants and degradation of
Esters 75

2-ethyl hexanol by lakewater bacteria, Appl. Microbiol. 90. (a) Lansdown, A. R., 1997, Aviation lubricants. In Chemistry
Biotechnol., 25, 558–567. and Technology of Lubricants, 2nd edn., eds. Mortier, R.
74. Witton, K. E., 1988, The Degradation of a Synthetic 2-Stroke M. and Orzulik, S. T., Blackie Academic and Professional,
Lubricating Oil in Freshwater Systems, PhD thesis, University London, U.K., pp. 264–286. (b) Phillips, W. D., 2003, Turbine
of Kent, Canterbury, U.K. lubricating oils and hydraulic fluids. In Fuels and Lubricants
75. Cain, R. B., 1991, Biodegradation of lubricants. In Handbook, ed. Totten, G. E., ASTM International, West
Biodegradation and Biodeterioration, Vol. 8, ed. Rossmore, Conshohocken, PA, pp. 297–353.
H. W., Elsevier, Amsterdam, The Netherlands. 91. (a) Wright, L., 1996, Formation of the neurotoxin TMPP from
76. Randles, S. J. and Penman, M. G., 1995, Handling and dis- TMPE-phosphate formulations, Tribol. Trans., 39, 827–834.
posal of polyol ester refrigeration lubricants, ASHRAE Trans., (b) Wyman, F. J., Porvaznik, M., Serve, P., Hobson, D., and
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

SD-95-8-4. Uddin, D. E., 1987, High temperature decomposition of mili-


77. Glasgow, G. and Bruns, R. J., 1979, Reclamation of synthetic tary specification L-23699 synthetic aircraft lubricants, J. Fire
turbine engine oil mixtures, USAF Report AFAPL-TR-78-50, Sci., 5, 162–177.
MRC-DA-780, 1979. 92. Bunemann, T. F., Kardol., A. D., and de Mooij, A. C., 2004,
78. Coffin, P. S., Lindsay, C. M., Mills, A. J., Lindencamp, H., Hydraulic fluids, US 6693064.
and Fuhrmann, J., 1979, The application of synthetic fluids to 93. Carr, D. D. and Schaefer, T. G., 1992, Synthetic ester lubricant
automotive lubricant development trends today and tomorrow, base stock used as chain oils and lubricants for gas turbine
J. Synth. Lubr., 7(2), 123. engines, EP 051856.
79. Boyde, S. and Randles, S. J., 2009, Engine lubricants, EP 94. (a) Hampson, D. F. G., 1975, Reducing the risk of fires in
2118246. (b) Kitching, J., Randles, S. J., Boyde, S., and large reciprocating air compressor systems, Wear, 34, 399. (b)
Lefevre, J., 2009, Engine lubricants, EP 2268778. Thoenes, H. W., 1975, Safety aspects for selection and testing
80. Korcek, S., Sorab, J., Johnson, M. D., and Jensen, R. K., 2000, of air compressor lubricants, Lubr. Eng., 26, 409.
Automotive lubricants for the next millennium, Ind. Lubr. 95. van Ormer, H. P., 1987, Trim compressed-air cost with syn-
Tribol., 52, 209. thetic lubricants, Power, 43.
81. Coant, P. M. and Munns, G. W., 1979, Synthetic lubricant, US 96. (a) Witts, J. J., 1989, Diester lubricants in petroleum and chem-
4175046. ical plant service, J. Synth. Lubr., 5(4), 321. (b) Nippon Oils
82. Bartz, W. J., 1998, Fuel economy improvement by engine and and Fats Co. Ltd., 1980, Neopentyl polyol ester production, JP
gear oils. In Tribology for Energy Conservation. Proceedings 55105644.
of Leeds Lyon Symposium 1997, London, U.K., Elsevier, 97. Carswell, R., McGraw, P. W., and Ward, E. L., 1983, A blend
Amsterdam, The Netherlands, pp. 13–20. of a polyglycol and a tetraester as a lubricant for rotary screw
83. (a) Beimesch, B. J. and Lakes, S. C., 1998, Base stock for syn- air compressors, Lubr. Eng., 39, 684.
thetic gear or transmission lubricant, WO 9804658. (b) Watts, 98. Michels, H. H. and Sienel, T. H., 2003, Refrigeration lubri-
R. F., 2003, Power transmission fluid composition for reducing cants – Properties and applications. In Fuels and Lubricants
energy consumption in transmission, EP1369470. (c) Hartley, Handbook, ed. Totten, G. E., ASTM International, West
R. J. and Srinivasan, S., 1994, Synthetic gear oil for manual Conshohocken, PA, pp. 413–430.
transmissions, US 5358650. 99. (a) Nakanishi, H., Onodera, Y., and Yasushi, N., 1997, Heat-
84. (a) Yoshimura, N., Komatsu, Y., and Tomizawa, H., 1999, resistant dipentaerythritol ester-based lubricating oil compo-
Traction fluid composition comprising a cyclohexyl diester sitions for chains, JP 09217078. (b) Carr, D. D., Hutter, J. A.,
and branched poly alpha olefin, US 5259978. (b) Capehart, and McHenry, M. A., 2002, Synthetic polyol ester based lubri-
T. W., Chapaton, T. J., Linden, J. L., and Mance, A. M., 2004, cant, for high temperature chain oil applications, US 6436881.
Composition useful in toroidal continuously variable trans- (c) Keller, H., 2006, Synthetic high-temperature lubricant for
mission systems, US 6797680. (c) Dare-Edwards, M. P., 1991, chains, DE 202006014282.
A novel family of traction fluids deriving from molecular 100. Yokota, H., et al, 1999, Cutting or grinding oil composition,
design, J. Synth. Lubr., 8, 197. US 20020035043.
85. (a) Tipton, C. D., 2002, Traction fluid formulation, US 101. Coffin, P. S., 1984, Characteristics of synthetic lubricat-
6372696. (b) Chiu, I., 2004, Continuously variable transmis- ing greases, J. Synth. Lubr., 1, 34–60. (b) Stempfel, E. M.,
sion fluid and method of making same, WO 2004031330. (c) 1998, Practical experience with highly biodegradable lubri-
Hata, H., Koga, H., and Tsubouchi, T., 2002, Lubricant base cants, especially hydraulic oils and lubricating greases, NLGI
oil composition comprises naphthene synthetic base oil and Spokesman, 62(1), 8–23.
fatty acid ester of alcohol having gem type dimethyl struc- 102. Spencer, S. J., 2000, Governments, operators eyeing effects of
tures, EP 1391499. synthetic-based drilling fluids, Oil Gas J., 88–89.
86. Hata, H. and Tsubouchi, T., 1998, Molecular structures of trac- 103. McShane, C. P., 2000, New safety dielectric coolants for dis-
tion fluids in relation to traction properties, Tribol. Lett., 5, 69. tribution and power transformers, IEEE Industry Application
87. Suigura, K. and Kagaya, M., 1977, A study of visible Magazine, May/June 2000, pp. 24–32.
smoke reduction from a small two-stroke engine using vari- 104. Waddington, F. B., 1983, High temperature esters: New
ous lubricants, SAE Paper 770623, SAE International, dielectric fluids for power engineering applications, GEC J.
doi:10.4271/770623. Sci. Technol., 49(1), 18–22.
88. Beimesch, B. J. and Zehler, E. R., 1997, Ester base stock for 105. Randles, S. J., 2006, Esters. In Synthetics, Mineral Oils, and
smokeless two cycle engine lubricant, US 6197731. Bio-Based Lubricants, ed. Rudnick, L. R., CRC Press, Boca
89. (a) Duncan, C. B., Hartley, R. J., and Tiffany, G. M., 2000, Raton, FL.
Biodegradable lubricant or functional fluid composition for 106. Randles, S. J., 1999, Esters. In Synthetic Lubricants and
two-cycle engines, US 6054420. (b) Carr, D. D., McHenry, High-Performance Functional Fluids, eds. Rudnick, L.
M. A., and Styer, J. P., 2002, Biodegradable 2-stroke engine R. and Shubkin, R. L., Marcel Dekker, New York, NY, pp.
lubricant base stock composition, US 6551968. 63–101.
76 Synthetics, Mineral Oils, and Bio-Based Lubricants

107. Kishi, J., Nakai, Y., Tanaka, S., and Inayama, T., 2016, Mixed refrigerants, International Refrigeration and Air Conditioning
ester, US 9328306. Conference, Paper 600, http://docs.lib.purdue.edu/iracc/600
108. Storzum, U., Breitscheidel, B., Todd, K. S., and Veinot, J. D., 124. (a) Allgood, C. C. and Yokozeki, A., 1994, Interactions of
2017, Plasticizer composition comprising di(2-ethylhexyl) hydrofluorocarbons and polyolester lubricants, International
terephthalate, US 9670128. Compressor Engineering Conference, Paper 1000, http://docs.
109. (a) Luitjes, H. and Jansen, J. C., 1999, Bicyclooctane deriv- lib.purdue.edu/icec/1000 (b) Wahlstrom, A. and Vamling, L.,
atives as plasticisers, WO1999045060A1. (b) Chen, C., 2000, Solubility of hydrofluorocarbons in pentaerythritol tet-
Donaghy, C., and Kurchan, A. N., 2018, Seal swell additive, raalkyl esters, J. Chem. Eng. Data, 45, 97–103. (c) Inoue, K.
US 9862908. and Iwamoto, A., 1992, Mutual solubility of HFC-134a and
110. (a) Hentschel, K., Dhein, R., and Schule, W., 1980, Lubricating synthetic ester base stocks, Sekiyu Gakkaishi, 35(1), 1992.
Downloaded By: 10.3.98.104 At: 16:14 22 Jun 2021; For: 9781315158150, chapter3, 10.1201/9781315158150-3

oils for the working of metals, US 4292187. (b) Topham, 125. (a) Pham, H. M. and Monnier, K., 2016, Interim and long-
A., 1980, Dispersing agents, US 4224212. (c) Ansmann, A., term low-GWP refrigerant solutions for air conditioning,
Kawa, R., Gondek, H., Strauss, G., and Von Kries, R., 1998, International Refrigeration and Air Conditioning Conference,
Polyglycerol polyricinoleates, US 5736581. (d) Kersbulck, J., Paper 1734, http://docs.lib.purdue.edu/iracc/1734 (b) Leon, R.
Vinci, D., and Thoen, J. A., 2013, Esters of secondary hydroxy A. U., Benanti, T. L., and Hessell, E. T., 2014, Solution proper-
fatty acid oligomers and reparation thereof, US 8580984. ties of polyol ester lubricants designed for use with R-32 and
111. Isbell, T. A. and Kleiman, R., 1994, Characterization of related low GWP refrigerant blends, International Compressor
estolides produced from the acid‐catalyzed condensation of Engineering Conference, Paper 2355, http://docs.lib.purdue.
oleic acid, J. Am. Oil Chem. Soc., 71, 379. edu/icec/2355 (c) Chen, Y., Hung, J., Tang, H., and Tsaih, J.,
112. Beyer, J. C. and Kurchan, A. N., 2017, Lubricant base stock, 2016, POE lubricant candidates for low GWP refrigerants,
US 9677022. International Refrigeration and Air Conditioning Conference,
113. CEC L-103-12. 2014, Biological Degradability of Lubricants Paper 1783, http://docs.lib.purdue.edu/iracc/1783
in Natural Environment, CEC, Available online at: www.cect- 126. Liscic, B., Tensi, H. M., Totten, G. E., and Webster, G. A.,
ests.org/disptestdoc1.asp 2003, Non-lubricating process fluids: Steel quenching tech-
114. OECD, 1992, Test No. 301: Ready biodegradability. In OECD nology. In Fuels and Lubricants Handbook, ed. Totten, G. E.,
Guidelines for the Testing of Chemicals, Section 3, No. 290, ASTM International, West Conshohocken, PA, pp. 587–634.
OECD Publishing, Paris, http:​//dx.​doi.o​rg/10​.1787​/9789​26407​ 127. Razali, S. Z., 2018, Review of biodegradable synthetic-based
0349-​en. drilling fluid: Progression, performance and future prospect,
115. Appelman, E., Kenbeek, D., and Rieffe, H. L., 2002, Lubricant, Renew. Sust. Energy Rev., 90, 171.
US 6346504. 128. International Electrotechnical Commission, 2010, Insulating
116. Mahoney, C. L. and Barnum, E. R., 1962, Polyphenyl ethers. liquids – Specifications for unused synthetic organic esters for
In Synthetic Lubricants, eds. Gunderson, R. C. and Hart, A. electrical purposes, IEC 61099:2010, https://webstore.iec.ch/
W., Reinhold Publishing Corp., New York, NY, pp. 402–463. publication/4511
117. Byford, D. C. and Edgington, P. G., 1971, The development 129. International Electrotechnical Commission, 2013, Fluids for
of special oils for the power plants of supersonic transports, electrotechnical applications – Unused natural esters for trans-
Document WPC-14412, 8th World Petroleum Congress, June formers and similar electrical equipment, IEC 62770:2013,
13–18, Moscow. https://webstore.iec.ch/publication/7422
118. (a) MIL-PRF-7808L, 1997, USA Military Specification, 130. Abe, K., Seiki, H., and Iwata, M., 1992, Flame retardant
Lubricating oil, aircraft turbine engine, synthetic base. hydraulic oil containing a synthetic ester, US 6361711.
(b) MIL-PRF-23699G, 2014, USA Military Specification, 131. Schnur, N. E. and Beimesch, B. J., 1982, Low temperature
Lubricating oil, aircraft turbine engine, synthetic base, NATO hydraulic fluids, US 4519932.
Code Numbers: O-152, O-154, O-156, and O-167, http:// 132. Drake, H. N., 1986, Fire-resistant hydraulic fluid, US 4645615.
qpldocs.dla.mil 133. (a) Evans, R. D. and Craft, R. D., 1994, Method for lubricating
119. SAE Aviation, 2018, Specification for aero and aero-derived metal-metal contact systems in metalworking operations, US
gas turbine engine lubricants, SAE AS5780D, www.sae.org/ 5318711. (b) Benda, R., Godwin, A. D., and Wallaert, B. V. R.,
standards/content/as5780d/ 1990, Esters and fluids containing them, EP 0523122.
120. Schlosberg, R. H., Duncan, C. B., Turner, D. W., Ashcraft, T. 134. (a) Zhu, T., Schellingerhout, P., Zhang, Y., and Ma, J., 2017,
L., and Munley, W. J., 1996, Reduced odor and high stability Small particle size oil in water lubricant fluid, US 9707605.
aircraft turbine oil base stock, WO 1998008801. (b) Mulder, H. L., Smeulders, J. B., Herrendorf, R. A., and
121. Carr, D. and DeGeorge, N., 1986, Synthetic lubricant base Robert, A., 2012, Metal processing lubricant composition, US
stock of monopentaerythritol and trimethylolpropane esters, 8293691. (c) Takahama, Y., Shiraishi, T., Ogawa, S., Vanel,
US 4826633. L., Hauret, G., Laugier, M., and Masson, P., 2014, Method of
122. Williamitis, V. A., 1957, Working fluids in refrigeration appli- supplying lubrication oil in cold rolling, US 8720244.
cations, US 2807155. 135. Miller, P. R. and Patel, H., 1997, Using complex polymeric
123. Tominaga, S., Tazaki, T., and Hara, S., 2002, Various per- esters as multifunctional replacements for chlorine and other
formance of polyvinylether (PVE) lubricants with HFC additives in metalworking, Lubr. Eng., 53, 31–33.

You might also like