You are on page 1of 13

Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Contents lists available at ScienceDirect

Renewable and Sustainable Energy Reviews


journal homepage: www.elsevier.com/locate/rser

Tribological behavior of biolubricant base stocks and additives T


⁎ ⁎
Chung-Hung Chan , Sook Wah Tang, Noor Khairin Mohd, Wen Huei Lim , Shoot Kian Yeong,
Zainab Idris
Advanced Oleochemical Technology Division, Malaysian Palm Oil Board, 43000 Kajang, Selangor, Malaysia

A R T I C LE I N FO A B S T R A C T

Keywords: Biolubricants are gaining popularity and acceptance globally due to their sustainable and environmentally
Vegetable oil friendly properties; being derived from feedstocks from vegetable oils. Indeed, the potential for biolubricants to
Biolubricant eventually replace conventional lubricants is currently viewed in the literature as a real possibility. This study
Additives will provide valuable information pertaining to the formulation of biolubricants, by assessing and evidencing the
Tribology
tribological performances of various types of biolubricant base stocks and their related additives. This study
Friction coefficient
begins with a presentation of the basic tribological parameters in lubrication. Following that, the criteria for the
Wear
molecular structure of biolubricant base stocks for high tribological performance are discussed, based both on
the tabulation of the friction coefficient and on the measurement of wear scar from experimental studies. The
biolubricant base stocks under review in this study include vegetable oils (VO), epoxidized VO, ring-opened
products from epoxidized VO, estolides, and polyol esters. This review also discusses recent advances in eco-
friendly tribological additives such as plant-derived compounds and polymers, particulate and layered materials,
and ionic liquids. The performance and various applications of these additives are also reviewed.

1. Introduction 28/EC were mandated to promote the use of energy from renewable
sources in various sectors including fuels and lubricants [5]. At a con-
Biolubricants, or bio-based lubricants, are generally perceived as sumer level, environmental awareness is being raised by the awarding
environmentally friendly, biodegradable and non-toxic lubricants of environmental labels such as Blue Angel, Nordic Swan and European
which are derived from renewable resources such as plant oils and Eco-label for lubricant products [3], with the aim of highlighting the
animal fats. Most biolubricant products currently on the market are consequences of consumer purchasing and hiring decisions for the
made up either entirely or partially from bio-based oils, providing these planet. In addition to the above initiatives, the market growth of bio-
oils fulfill the requirements of international standards in terms of re- lubricants is further substantiated by the increase in global lubricant
newability, biodegradability, toxicity and technical performance. demand due to the specific development of end-user industries in
According to European standards [1,2], a certified biolubricant must China, India, South Africa, Brazil and Iran [6,7]. In the current scenario,
have a bio-based carbon content of at least 25% and biodegradability of the global market for lubricants is growing at approximately 2% an-
at least 60%. The biolubricant must also be non-toxic to the environ- nually; amounting to approximately 144.45 billion USD in 2015 [7].
ment and fit for the purpose of an application. The market demand for Indeed, the global biolubricant market has always been rising steadily
biolubricants is driven by various factors, which include: consumer at 10% annually, in spite of it constituting only about 1% of the total
environmental awareness, government directives and the global de- market in lubricants [8,9]. From this perspective, the shift towards bio-
mand for lubricants. A long-running and increasingly urgent issue is the based lubricants represents a global trend in the global lubricant
environmental cost of hazardous and non-degradable lubricants en- market; as part of the global effort to mitigate the environmental im-
tering the ecosystem through total loss applications, spillages and so on pact of hazardous and non-degradable conventional lubricants.
[3]. It is, in fact, well-attested that about half of the total production of The successful formulation of any lubricant product is built upon a
lubricants in the world accumulates in the environment every year multi-objective optimization of the types and concentrations of base
[3,4]. Continuing efforts have been made to fight this crisis through stocks and additives, to meet specific application specifications and
various government directives and environmental legislation. For in- requirements [10]. On average, a typical lubricating oil consists of 93%
stance, the European Union (EU) Renewable Energy Directives 2009/ base stock and 7% additives, though the additive content may vary


Corresponding authors.
E-mail addresses: rykenz87@yahoo.com, chan@mpob.gov.my (C.-H. Chan), limwen@mpob.gov.my (W.H. Lim).

https://doi.org/10.1016/j.rser.2018.05.024
Received 17 March 2017; Received in revised form 27 March 2018; Accepted 13 May 2018
Available online 23 May 2018
1364-0321/ © 2018 Elsevier Ltd. All rights reserved.
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

from 1%, as in simple compressor oils, or up to 30%, as in gear oil and 2. Basic tribological parameters
metal-working fluids [11,12]. In any lubricant formulation, base stock
is viewed as the essential part of the lubricating oil, as base stock The most important consideration for selecting a suitable lubricant
possesses the vital lubricant physical properties such as viscosity, is that it fulfills the lubrication requirements of the particular applica-
thermal stability, lubricity and so forth. Among all base stocks, vege- tion, especially in its tribological aspects. Tribological performance in
table oil (VO) is the simplest choice for a biolubricant [13–19]. It has in lubrication concerns lubricity, friction and wear. Generally speaking,
fact always been the alternative to mineral oil due to its superior lu- lubricity is associated with the formation of the lubrication layer, or
bricity, viscosity index, and wear resistance; not to mention its biode- tribofilm, on sliding surfaces. High lubricity reduces the direct contacts
gradability and the sustainability of its feedstock [13,20–22]. Un- of the surface asperities and thereby lowers friction and energy loss
fortunately, VO is known to have poor thermal and oxidative stability [11,96]. High lubricity is not always accompanied by better wear
[23–25], but by means of chemical modifications, VO may be converted protection, since the mechanism of forming a protective layer on the
into better-performing products such as epoxidized VO [26–28], ring- sliding surfaces occurs through the adsorption of the surface active
opened products from epoxidized VO [23,29–31], estolides [32–35], materials of the lubricant, which can be certain types of base stock or
and polyol esters [36–42]. These base stocks have been seen as pro- additives [97]. Whenever there is surface asperity contact during the
mising for use in biolubricant formulations. Further improvement to sliding activity, there is the possibility that three types of mechanical
VO-based lubricant performance is possible through additivation. Ad- wear are produced: abrasion, adhesion, and fatigue [46,97]. Abrasive
ditives are able to enhance the physical properties of the lubricant by wear (e.g. grooves) is produced via the removal of a certain volume of
providing additional features, including anti-corrosion, metal scaven- surface material by hard asperities on the sliding surface [98]; adhesive
ging, and anti-wear to the base stock [12]. The most important group of wear (e.g. attachment of wear debris) is the result of plastic deforma-
additives are tribological additives, as far as lubrication is concerned, tion and strong bonding of the interface materials [98]; fatigue wear
since the ability to reduce wear and damage to machines is ultimately (e.g. crack, fracture) is due to the weakening of surface materials after
the goal. Indeed, the aim of any new formulation is to achieve better repeatedly contact with high local stress for a large number of times
performance, higher energy efficiency, and a longer life cycle for the [98]. The obvious solution to the wear problem is to employ thicker
machinery, as well as longer intervals for the replacement of the lu- tribofilm, though this is not always efficient and desirable in some
bricating oil [43]. It should be mentioned here, of course, that not all applications, such as metalworking. Prior to any discussion of tribolo-
conventionally-used additives are compatible with biolubricant base gical performance, it is essential to understand what are known as the
stock [12], but tribological additives for biolubricants have so far been three basic tribological parameters: the mechanical properties of a tri-
advanced in the field with favorable regard to their lubricity, wear bosystem; the lubrication regime and its conditions; and also the phy-
protection and load-carrying capacity. The additives from this group sical properties and tribochemistry of the lubricant as illustrated in
which are currently enjoying most research focus are those containing Fig. 1. The details of each parameter are elucidated as follows.
sulfur and phosphorus [44–47], environmental-friendly polymers
[15,48–50], plant-derived compounds [35,51], particulates [52–56], 2.1. Mechanical properties of a tribosystem
two-dimensional (2D) layered materials [57–60], and ionic liquids
[61–64]. The mechanical characteristics of a tribosystem can essentially be
In practice, the varying performance of different lubrications in described as the material hardness, the surface roughness, the contact
terms of friction and wear is usually assessed based on the coefficient of geometry, and the sliding mechanism of the interacting part and its
friction (COF) and wear scar diameter (WSD), or volume (WSV), re- counterpart, as presented in Fig. 1(A). Limiting the discussion to the
spectively [13,35,65,66]. COF denotes the ratio of friction force be- laboratory tribometer, most sliding materials have a 30–64 Rockwell C
tween two bodies and the normal force pressing the bodies together Hardness (HRC) and a 0.01–1.0 µm surface roughness, depending on
[67], whereas WSD and WSV are respectively the diameter and volume the following: the type of material (e.g. steel, magnesium, copper), the
of wear after sliding activity. These measurements are often used in hardening process (e.g. nitriding, carburizing), and the coating (e.g.
sub-scale machine or laboratory tests to investigate the tribological carbon-based coating, ultra-high molecular weight polyethylene)
behavior of lubricant products in the workings of real machinery. When [15–17,26,37,66,99]. Basically, a higher hardness value provides a
evaluating and comparing the tribological behavior of different lu- lower wear rate, while a lower roughness value provides better lubricity
brications, it is important to specify the type of lubricant, the tribo- [100]. Comparing tribological results across different types of trib-
system, and the lubrication conditions to minimize the discrepancy ometers is not possible due to varying sliding geometries (e.g. four-ball,
between the laboratory test and the field performance [65,67]. In the pin-on-disk, ball-on-disk) and the particular sliding mechanism (e.g.
context of biolubricants, most of the literature simply discusses over- unidirectional or reciprocating, sliding to rolling ratio) [65,67]. For a
views, developments, synthesis and applications of biolubricants in a valid comparison, the measurements should be conducted using a si-
very general context [24,28,68–76], or in specific applications [77–87], milar type of machine and method (e.g. ASTM D5183, ASTM D4172,
while other studies focus only on the development and performance of a and ASTM D6079). Even if all of these conditions are closely matched,
specific type of additive [88–95]. In fact, a thorough search of the re- researchers are required to reproduce the measurements in their in-
levant research yielded no reviews which provided any adequate detail struments to ensure accuracy.
about aspects of tribological behaviors. This review, therefore, seeks to
address this gap in the literature by providing comprehensive evidence 2.2. Lubrication regimes and conditions
and valuable discussion pertaining to the tribological performances of
recent biolubricant base stocks, and related additives. This study fo- The operating conditions of any tribosystem play a significant role
cuses on their influencing parameters, molecular structure criteria, sy- in characterizing tribofilm and, subsequently, tribological behavior.
nergisms, and mechanisms; concerning lubricity, wear protection, and According to the famous Stribeck curve [67,101] there are three lu-
load-carrying capacity. In addition, a detailed tabulation of friction and brication regimes, namely: boundary; mixed and elastohydrodynamic
wear responses of various biolubricants is presented for comparison (EHL); and hydrodynamic lubrication regimes, as illustrated in
purposes. The aim is for interested parties to be able to utilize the tri- Fig. 1(B). Boundary lubrication usually occurs at high contact load or
bological information presented in this review as a guide to selecting low sliding speed lubrication, whereby thin or unstable tribofilm is
and formulating biolubricants for their own applications. formed, causing the contacting surfaces to rub against each other and
generate high friction and wear [102]. As the film thickness in this
regime is less than the surface roughness [102], the presence of surface

146
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Fig. 1. Three basic parameters affecting friction and wear behaviors of a lubricated tribosystem.

active materials is vital in order to protect the contacting surfaces from ultimately of tribological performance. For as more viscous lubricant
sliding damage. Further increase in the sliding speed or decrease in the creates thicker tribofilm but gives lower efficiency due to higher flow
contact load shifts the lubrication to a mixed/EHL regime, and then to a resistance [3], the selection of lubricant viscosity for an application is a
hydrodynamic lubrication regime. Along with this transition, the as- trade-off between lubrication and energy performance. In cases where
sociated friction and wear rate decrease since the contacting surfaces elevated temperature and high contact load are involved, a high visc-
are now separated by a thicker tribofilm due to hydrodynamic lift osity index and high pressure-viscosity coefficient are preferred in lu-
[102]. The film thickness is about 1–3 times that of the surface bricants to ensure stable tribofilm [3]. Other properties that affect the
roughness under mixed/EHL regime, and 3 times greater than the strength and the stability of tribofilm include: pour point, volatility,
surface roughness when under a hydrodynamic lubrication regime thermal degradation temperature, and oxidative stability [3,98]. On the
[102]. In the latter regime, the sliding contacts are now fully lubricated other hand, where the tribochemical properties of a lubricant are
and the associated friction is dominated by the viscous drag and re- concerned; lubricity, wear protection, and load carrying capacity are
sistance of the lubricant. In this manner, there is no significant me- exhibited by so-called surface active materials. These surface active
chanical wear on the moving parts except fatigue [98]. According to materials mainly act as a friction modifier (FM), an anti-wear additive
rough tabulations [97], a typical boundary lubrication has a coefficient (AW), and an extreme pressure additive (EP). More generally, FM
of friction (COF) value greater than 0.1, whereas mixed/EHL lubrica- promotes softer and easily sheared protection layers which reduce
tion produces a COF in the range of 0.01–0.10, and hydrodynamic lu- sliding friction [20,103], whereas AW and EP actually form a sacrificial
brication gives COF less than 0.01. Note that the sliding speed and protective layer on the sliding surface to minimize wear [45,46,103].
contact load to achieve a certain lubrication regime vary according to The typical COF of an FM layer (0.01–0.05) is generally lower than that
both the tribosystem and the physical properties of the lubricant; of EP and AW layers (0.06–0.15), but higher if compared to tribofilm in
especially the physical property of viscosity. Other influential operating a hydrodynamic regime (< 0.001) [103]. The differences between AW
parameters such as lubrication temperature and assessment time also and EP are that the latter has a higher reactivity and its protective layer
need to be specified. In most tribological assessments, the lubrication is tougher and thicker, allowing for a higher load tolerance [103]. The
operates within a boundary or mixed/EHL regimes, at an operating above additives are mostly employed to minimize wear and enhance
temperature which closely matches the required equilibrium tempera- lubricity for boundary lubrication and mixed/EHL lubrication regimes.
ture of an application. The friction and wear measurements are taken
after reaching the steady state, which is about an hour of operation, as
3. Molecular structure criteria of biolubricant base stocks
adopted in most standard practices.
The physical and tribochemical properties of biolubricant base
2.3. Physical and tribochemical properties of lubricant stocks depend very much on the molecular structure of their constituent
compounds. In most biolubricant base stocks, there are polar and non-
The physical properties and the tribochemical properties of any polar regions in the molecules [12]. The polar region is responsible for
lubricant are the most crucial parameters for tribological performance, adsorption or adhesion on a sliding surface, whereas the non-polar re-
as depicted in Fig. 1(C). In view of the physical properties of base stock, gion is responsible for the strength and oiliness of the lubricating fluid
viscosity and its relationship with temperature and pressure is the [3]. Some examples of the molecular structure of the base stocks can be
primary property responsible for the thickness of tribofilm, and found in Fig. 2. The simplest biolubricant base stock of all is vegetable

147
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Fig. 2. Molecular structures of biolubricant base stocks and their basic property-structure relationships. Molecular structure: (1) triglyceride, (2) oleic acid, (3)
ricinoleic acid, (4) epoxidized oleic acid, (5) epoxidized ricinoleic acid, (6) oleic estolide, (7) 9-hydroxy-10-acyloxyoctadecanoic acid, (8) octyl 9-hydroxy-10-
acyloxyoctadecanoate, (9) octyl 9-(lauroyloxy)-10-(acyloxy)octadecanoate, (10) 10,12-dihydroxy-9-acyloxystearic acid, (11) octyl 10,12-dihydroxy-9-acylox-
ystearate, (12) trimethylolpropane (TMP) ester, (13) pentaerythritol (PE) ester, (14) neopentyl glycol (NPG) ester.

oil (VO). VO is triglyceride consisting of three straight chained fatty products from epoxidized VO, estolides, and polyol esters [69]. Epox-
acids attached to a glycerol backbone by ester linkage. Typically, the idized VO is actually produced by forming a highly reactive oxirane
fatty acid chains in VO triglyceride have a 14–24 carbon atoms chain ring at the unsaturation site of the fatty acid chain [105]. Subsequently,
length, with varying degrees of unsaturation [24], although coconut oil the oxirane ring reacts with suitable organic acid or aliphatic alcohols
has mostly 12 carbon atoms in its fatty acid chain [104] and castor oil to produce poly-functional molecules [105,106]. Estolide, on the other
has an additional hydroxyl group (-OH) in the fatty acid chain (i.e. ri- hand, is produced through homopolymerization of fatty acids at its
cinoleic acid) [33]. Due to the intrinsic instability of the β-hydrogen unsaturation sites by ester linkage. It is characterized by the degree of
atom in the glycerol backbone coupled with the unsaturation site at the polymerization or estolide number [32,107]. Lastly, polyol esters can
fatty acid chains in VO, it is accepted that chemical modifications on be produced from VO methyl ester or fatty acid through transester-
such sites are possible with the aim of generating better oxidative and ification or esterification with polyols, including neopentyl glycol
thermal stabilities for products such as, epoxidized VO, ring-opened (NPG), trimethylolpropane (TMP) and pentaerythritol (PE) [108–111].

148
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

The performances of the aforementioned base stocks can essentially be [39], where incorporating a longer chain of alkyl (C5 to C18) into the
based on four criteria: degree of unsaturation, carbon chain length, ester actually enhances both its pour point (−75 °C to −42 °C) and
functional group(s) and branching effect [112,113]. A schematic dia- viscosity index (80–208). A different length of alkyl chain can tailor the
gram showing the effects of the molecular structure criteria on the ester to a specific application [39]: Valeric acid-TMP ester (C5) is sui-
physical and tribological properties of biolubricant base stocks is pre- table, as cutting fluids at low-temperature application due to low
sented in Fig. 2. The details of each criterion will be elucidated as viscosity; Caprylic acid–TMP ester (C8) is used for high viscosity and
follows. stability applications like dielectric coolants and rail wheel lubricants;
in addition to TMP-oleate (C18), which is mostly employed for hy-
3.1. Degree of unsaturation draulic fluid. It is important to know that the above relationship be-
tween the carbon chain length and the pour point is not applicable for
Degree of unsaturation refers to the number of unsaturation sites in biolubricant molecules that have significant steric hindrance, such as
the carbon chains of a molecule. This criterion is often used to assess the base stocks from ring-opening of epoxidized VO. In this case, the re-
non-polar regions of VO molecules, since the oil contains different lationship is in reverse: the increase in the chain length creates a larger
compositions of saturated and unsaturated fatty acid chains. A thorough steric barrier to inhibit the molecule from crystallization [29,120], thus
study performed by Reeves et al. [16] found that the higher the degree improving its cold flow property [105].
of unsaturation of VO, the higher were the COF and WSV produced.
This is because the unsaturation site disrupts and prevents the carbon 3.3. Functional groups
chains of VO from forming a densely packed bundle, weakening the
stability of the tribofilm [16,17,114]. It is also evidenced that highly The polarity in the polar region of a biolubricant molecule affects its
unsaturated VO tends to have a higher composition of free fatty acids, intermolecular interaction and its interaction with sliding surfaces. This
which may induce chemical wear on the sliding material due to its polarity influences both the tribological behavior and the physical
reactive carboxyl group (RCOOH) [17,49]. Nevertheless, at extreme properties of the biolubricant molecule. The most common functional
pressure lubrication, the reactive carboxyl group of the free fatty acid group in biolubricants is the ester group (RCOOR). The ester group has
may have an excellent anti-wear effect [18], as it allows better coupling only a mild polarity, but it is able to provide sufficient adhesive
of the fatty acid with the sliding material to produce low friction car- strength to form tribofilm with a sliding surface. It exhibits better anti-
boxylate soap protective film [115]. As a whole, the effect of un- friction capability [13,17,37,99] and promotes lower operating tem-
saturation of VO on tribological performance is significant only in perature in gear compared to mineral oils [121–123]. However, the
metallic-to-metallic contact, since the metallic oxide layer allows che- ester group does not have any superior anti-wear effect on the boundary
misorption of VO or free fatty acid [98]. Otherwise, it does not result in lubrication regime, due to its mild polarity and adhesiveness [13,14].
any substantial difference, as is reported in the lubrication of ultra- Another functional group which is commonly found in biolubricant
high-molecular-weight polyethylene (UHMWPE) sliding contacts using molecules is the hydroxyl group (OH). The hydroxyl group creates a
sesame and Nigella sativa oils [66]. In spite of the fact that a low degree thick and viscous fluid as a result of hydrogen intermolecular bonding.
of unsaturation is preferred in tribological performance, and also for A biolubricant with a hydroxyl group such as castor oil is suitable for
better thermal and oxidative stabilities, it is desirable to keep certain applications which operate in a mixed/EHL and hydrodynamic lu-
degree of unsaturation, so that the unsaturation sites disrupt the brication, such as hydraulic fluid [15,48]. Other than strong inter-
stacking of molecules and crystallization to prevent any hardening of molecular bonding, the presence of a hydroxyl group favors micro-
the base stock and subsequently to improve the cold flow property (e.g. crystal particle solubilization [124], but it compromises the cold flow
a low pour point) [116]. It is therefore agreed that a monounsaturated property and oxidation stability, unless the molecule possesses steric
carbon chain like the oleic acid chain is the best tradeoff in most base hindrance [29,106]. Among all groups, the strongest polar group
stocks, including VO [16,19,29] and other synthetic esters available in biolubricant base stocks is the carboxyl group (RCOOH)
[109,117,118]. such as fatty acids and their derivatives [106,120]. The molecule with
the carboxyl group has a strong adhesion to metallic surfaces in addi-
3.2. Carbon chain length and molecular weight tion to possessing good anti-wear properties in boundary lubrication
[16,20,125,126]. If the oxide layer of a sliding surface is worn off and
Carbon chain length in the non-polar region of biolubricant base there is no chemisorption site for adhesion, a chemical attack on the
stock molecules usually amounts to 8–22 carbon atoms chain length. material may occur in which the reactive carboxyl group causes soft-
Longer chain length molecules tend to create thicker tribofilm due to ening and melting of the metallic soap [98]. The above effects of
their increase in molecular weight and viscosity. Longer chain length functional groups can be multiplied and intensified with the increase of
also reduces the contact of asperities while increasing the protected the number of functional group in the molecule. Taking the ester group
surface area. With the increase in molecular weight and viscosity, long as an example, ring-opened products from epoxidized oil that have
chain length base stocks tend to shift the lubrication regime (as in more ester functional groups provide greater resistance to shear force
Fig. 1(B)) to the right to give lower COF, WSD and WSV [66,119]. and better wear protection [119].
While increasing the chain length of biolubricant molecules is desirable
for better tribological performance, it nonetheless affects some of the 3.4. Branching and steric hindrance
physical properties of the biolubricant. Usually, a long chain length in
biolubricant molecules improves tribological performance and viscosity Introducing branching around the polar region of a biolubricant
index, but at the cost of both pour point and oxidative stability [34,39]. molecule helps to shield the molecule from physical and chemical in-
These effects are not obvious for VOs, since their average chain length teractions due to steric hindrance. As mentioned previously, biolu-
spans only a narrow range, typically in between 16 and 18 carbon bricant molecules with steric hindrance such as ring-opened products
atoms [17,104], but the effects are profound and significant in esto- from epoxidized oil [105,106] and polyol esters [108] are able to in-
lides, polyol esters and the ring-opened products from epoxidized fatty hibit the stacking of molecules and crystallization. At low temperature
acids. For example, estolide capped with shorter chain fatty acids (C2 to conditions, the molecules instead form a microcrystalline structure.
C10) gives a better pour point (−12 to −30 °C) [34,107] when it is This structure allows the molecules to tumble and glide over one an-
capped with a long chain fatty acid or when it undergoes a large degree other to give fluidity [25]. In addition to that, other known advantages
of polymerization and its anti-wear property is enhanced due to viscous of having a branching structure in the molecule include better thermal,
tribofilm [35]. Another example can be seen in the case of TMP ester oxidative and hydrolytic stabilities [53,116]. In spite of the favorable

149
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Table 1
Tribological performances of biolubricant base stocks.
a a
Ref. Tribosystem Lubrication Base stock, viscosity (40 °C) COF WSD / WCV
conditions

[16] Pin-on-disk test: pin (oxygen-free electronic 10 N, 36 mm/s, 25 °C soybean oil, 25.9 cP 0.406 0.384 mm3
copper C101, 6.35 mm diameter, 50 mm length, corn oil, 26.8 cP 0.446 0.297 mm3
hemispherical tip); disk (aluminum alloy 2024, sesame oil, 28.6 cP 0.318 0.215 mm3
70 mm diameter, 6.35 mm thickness, 0.3 µm canola oil, 29.9 cP 0.087 0.201 mm3
roughness) peanut oil, 31.9 cP 0.098 0.168 mm3
genetically-modified safflower oil, 31.9 cP 0.494 0.195 mm3
avocado oil, 31.9 cP 0.020 0.104 mm3
olive oil, 32.7 cP 0.078 0.175 mm3
[17] Four-ball test: ball (AISI E-52100 chrome alloy 2574 MPa, 1200 rpm, coconut oil, 24.8 cSt 0.101 0.601 mm
steel, 12.7 mm diameter, extra polish EP grade 75 °C sunflower oil, 27.8 cSt 0.060 0.616 mm
25, 64–66 HRC hardness) rice bran oil, 40.6 cSt 0.073 0.585 mm
mineral oil, 105 cSt 0.117 0.549 mm
[66] Pin-on-disk test: pin (stainless steel, 0.06 µm 20 N, 30 mm/s, 25 °C sesame oil, 58.0 cP b 0.028 0.024 mm3
roughness); disk (ultra-high-molecular weight Nigella sativa oil, 64.5 cP b 0.032 0.028 mm3
polyethylene UHMWP, 0.31 µm roughness) dry or no lubricant 0.290 0.051 mm3
[18] Four-ball test: ball (Chrome alloy steel, 12.7 mm load, 1200 rpm, 75 °C olive oil, 37.042 cSt 0.052 0.621 mm
diameter, 0.1 µm roughness, 62 HRC hardness) load = 400 N commercial SAE15W40 oil, 103 cSt 0.080 0.717 mm
load = 500 N olive oil, 37.042 cSt 0.061 0.850 mm
commercial SAE15W40 oil, 103 cSt 0.084 1.172 mm
load = 600 N olive oil, 37.042 cSt 0.073 1.254 mm
commercial SAE15W40 oil, 103 cSt 0.090 1.640 mm
load = 800 N olive oil, 37.042 cSt 0.077 1.501 mm
commercial SAE15W40 oil, 103 cSt 0.097 2.095 mm
[19] Ball-on-disk (High frequency reciprocating rig load, speed (8 mm Jatropha Curcas L. oil, 36.61 cSt 0.126 < 0.001 mm3
(HFRR)): ball (AISI E52100 steel, 10 mm stroke), 20 °C hydrotreated rapeseed oil (paraffinic hydrocarbon containing free 0.164 < 0.001 mm3
diameter), disk (X210Cr12 steel) load = 10 N fatty acid and triglycerides), 4.68 cSt
speed = 10 Hz. rapeseed methyl ester (biodiesel), 2.83 cSt 0.139 0.001 mm3
load = 10 N Jatropha Curcas L. oil, 36.61 cSt 0.109 0.001 mm3
speed = 20 Hz hydrotreated rapeseed oil, 4.68 cSt 0.134 < 0.001 mm3
rapeseed methyl ester (biodiesel), 2.83 cSt 0.115 0.002 mm3
load = 20 N Jatropha Curcas L. oil, 36.61 cSt 0.129 0.003 mm3
speed = 10 Hz hydrotreated rapeseed oil, 4.68 cSt 0.153 < 0.001 mm3
rapeseed methyl ester (biodiesel), 2.83 cSt 0.132 0.002 mm3
load = 20 N Jatropha Curcas L. oil, 36.61 cSt 0.106 0.010 mm3
speed = 20 Hz hydrotreated rapeseed oil, 4.68 cSt 0.134 < 0.001 mm3
rapeseed methyl ester (biodiesel), 2.83 cSt 0.125 0.008 mm3
[26] Ball-on-disk test (HFRR): ball (AISI 52100 steel, 2 N, 20 Hz (1 mm moringa oil, 44.9 cSt 0.084 0.15 mm
570–750 HV hardness, 6 mm diameter); disk stroke), 60 °C passion fruit oil, 31.8 cSt 0.060 0.076 mm
(AISI 52100 steel, 190–210 HV hardness, 10 mm epoxidized moringa oil, 80.4 cSt 0.080 0.146 mm
diameter, #1200 polished surface) epoxidized passion fruit oil, 185.7 cSt 0.068 0.131 mm
commercial hydraulic oil A, 80–100 cSt 0.123 0.181 mm
commercial hydraulic oil B, > 150 cSt 0.103 0.164 mm
[29] Four-ball test: ball (AISI E-52100 chrome alloy 392 N, 1200 rpm, epoxidized oleic acid, 0.804 g/cm3 c 0.58 0.89 mm
steel, 12.7 mm diameter, extra polish EP grade 75 °C monoester
25, 64–66 HRC hardness) 9-hydroxy-10-octyloxyoctadecanoic acid, 0.813 g/cm3 c 0.47 0.83 mm
9-hydroxy-10-nonanoxyoctadecanoic acid, 0.819 g/cm3 c 0.42 0.75 mm
9-hydroxy-10-lauroxyoctadecanoic acid, 0.822 g/cm3 c 0.38 0.68 mm
9-hydroxy-10-myristoxyoctadecanoic acid, 0.826 g/cm3 c 0.27 0.56 mm
9-hydroxy-10-palmitoxyoctadecanoic acid, 0.830 g/cm3 c 0.21 0.48 mm
9-hydroxy-10-stearoxyoctadecanoic acid, 0.833 g/cm3 c 0.18 0.35 mm
9-hydroxy-10-behenoxyoctadecanoic acid, 0.836 g/cm3 c 0.14 0.28 mm
diester
octyl 9-hydroxy-10-octyloctadecanoate, 0.841 g/cm3 c 0.46 0.66 mm
octyl 9-hydroxy-10-nonanoxyoctadecanoate, 0.843 g/cm3 c 0.40 0.59 mm
octyl 9-hydroxy-10-lauroxyoctadecanoate, 0.846 g/cm3 c 0.33 0.53 mm
octyl 9-hydroxy-10-myristoxyoctadecanoate, 0.851 g/cm3 c 0.23 0.46 mm
octyl 9-hydroxy-10-palmitoxyoctadecanoate, 0.855 g/cm3 c 0.17 0.39 mm
octyl 9-hydroxy-10-stearoxyoctadecanoate, 0.858 g/cm3 c 0.13 0.34 mm
octyl 9-hydroxy-10-behenoxyoctadecanoate, 0.862 g/cm3 c 0.09 0.25 mm
Triester
octyl 9-(lauroyloxy)−10-(octyloxy)octadecanoate, 0.905 g/cm3 c 0.35 0.61 mm
octyl 9-(lauroyloxy)−10-(nonanoxy)octadecanoate, 0.913 g/cm3 c 0.31 0.53 mm
octyl 9-(lauroyloxy)−10-(lauroxy)octadecanoate, 0.920 g/cm3 c 0.25 0.49 mm
octyl 9-(lauroyloxy)−10-(myristoxy)octadecanoate, 0.925 g/cm3 c 0.20 0.42 mm
octyl 9-(lauroyloxy)−10-(palmitoxy)octadecanoate, 0.931 g/cm3 c 0.14 0.31 mm
octyl 9-(lauroyloxy)−10-(stearoxy)octadecanoate, 0.939 g/cm3 c 0.08 0.22 mm
octyl 9-(lauroyloxy)−10-(behenoxy)octadecanoate, 0.946 g/cm3 c 0.06 0.14 mm
(continued on next page)

150
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Table 1 (continued)
a a
Ref. Tribosystem Lubrication Base stock, viscosity (40 °C) COF WSD / WCV
conditions

[30] Ball-on-disk test (HFRR): ball (AISI 52100 steel, 200 g, 50 Hz (1 mm low lubricity diesel fuel (LLDF), n/a 0.400 0.656 mm
570–750 HV hardness, 6 mm diameter); disk stroke), 60 °C blend of LLDF and ring opening products of epoxidized canola oil 0.119 0.179 mm
(AISI 52100 steel, 190–210 HV, 10 mm diameter, with n-butanol at ratio 95:5 w/w, 251.7 cSt d
#1200 polished surface) blend of LLDF and ring opening products of epoxidized canola oil 0.099 0.142 mm
with amyl alcohol at ratio 95:5 w/w, 190.5 cSt d
blend of LLDF and ring opening products of epoxidized canola oil 0.081 0.120 mm
with 2-ethylhexanol at ratio 95:5 w/w, 85.5 cSt d
[35] Ball-on-three-plates test: plate (45° inclined, 20 N, 400 rpm, 25 °C oleic acids-estolides with degree of polymerization of 6.33 (O1), n/a 0.247 mm
C45E-1.1191 steel, 25–30 HRC hardness); ball 415.59 cSt
(100Cr6 steel, 12.7 mm diameter) ricinoleic acids-estolides with degree of polymerization of 6.92 (R1), n/a 0.174 mm
6712.98 cSt
blend of high-oleic sunflower oil (HOSO) and O1 at ratio: 85:15 w/ n/a 0.332 mm
w, 51.31 cSt
blend of HOSO and R1 at ratio: 85:15 w/w, 69.15 cSt n/a 0.242 mm
HOSO, 41.94 cSt n/a 0.357 mm
[42] Ball-on-disk test (HFRR): ball (detail n/a); disk 1 kg, 50 Hz (1 mm Lunaria annua oil-based TMP ester, 29.97 cSt 0.079 0.269 mm
(detail n/a) stroke), 25 °C mineral oil, 28.10 cSt 0.149 0.344 mm
[38] Four-ball test: ball (AISI 52–100 steel, 12.7 mm 40 kg, 1200 rpm, palm oil-based TMP ester (POTMP), 40.03 cSt 0.081 0.8053 mm
diameter, 64–66 HRC hardness) 75 °C ordinary lubricant (OL), 101.86 cSt 0.068 0.3511 mm
blend of OL and POTMP at ratio: 99:1 v/v, 108.75 cSt 0.064 0.3441 mm
blend of OL and POTMP at ratio: 97:3 v/v, 107.21 cSt 0.052 0.2789 mm
blend of OL and POTMP at ratio: 95:5 v/v, 101.42 cSt 0.056 0.3125 mm
blend of OL and POTMP at ratio: 93:7 v/v, 102.15 cSt 0.053 0.3631 mm
blend of OL and POTMP at ratio: 90:10 v/v, 97.637 cSt 0.065 0.3898 mm
[37] Four-ball test: ball (AISI 52–100 steel, 12.7 mm 1000 N, 1770 rpm, palm oil-based TMP ester, 40.03 cSt 0.25 2.47 mm
diameter, 64–66 HRC hardness) 25 °C palm oil-based PE ester, 68.40 cSt 0.06 2.54 mm
mineral oil, 30.61 cSt welded welded
commercial fully formulated lubricant, 101.86 cSt 0.08 0.54 mm
[36] Four-ball test: ball (AISI 52–100 steel, 12.7 mm 40 kg, 1200 rpm, palm oil-based NPG ester, 27.92 cSt 0.092 0.659 mm
diameter, 64–66 HRC hardness) Temperature (T) palm oil-based PE ester, 72.78 cSt 0.094 0.669 mm
T= 50 °C palm oil-based PE ester with anti-wear and anti-corrosion additives, 0.063 0.377 mm
75.37 cSt
commercial fully formulated lubricant, 131.76 cSt 0.087 0.218 mm
T= 100 °C palm oil-based NPG ester, 27.92 cSt 0.093 1.213 mm
palm oil-based PE ester, 72.78 cSt 0.050 0.584 mm
palm oil-based PE ester with anti-wear and anti-corrosion additives, 0.046 0.338 mm
75.37 cSt
commercial fully formulated lubricant, 131.76 cSt 0.101 0.434 mm

a
Values obtained at steady state or after sufficiently long sliding duration
b
Viscosity at 20 °C
c
Density is given instead since viscosity data is not available
d
Kinematic viscosity of the ring opening products at 40 °C rather than the mixture

physical properties above, excessive molecular branching puts the tri- A biolubricant base stock with ester functional groups such as VO is
bological properties at risk, as the adhesion of the molecule on the more efficient in terms of its frictional properties and wear protection
sliding surface could be disrupted and the probability of asperity con- than mineral oils [18]. VOs with a low degree of unsaturation such as
tact increased [98]. Molecular branching is nonetheless still beneficial avocado oil (0.985 unsaturation), olive oil (0.948 unsaturation), and
in lubrication involving high contact loads or extreme pressure. Indeed, peanut oil (1.102 unsaturation) exhibit superior frictional properties
in extreme cases, as observed in polyol esters; particularly TMP and PE compared to high unsaturation VOs such as soybean oil (1.451 un-
ester [38], molecular branching actually improves the load-carrying saturation), corn oil (1.381 unsaturation) and sesame oil (1.232 un-
capacity of the base stocks when benchmarking against mineral oils. saturation) [16]. Their wear protection effects, nonetheless, are not
This is because molecular branching helps to stabilize the tribofilm promising, particularly under extreme pressure conditions in a four-ball
from breaking apart [122]. test [17]. It is interesting to note that VOs derived from the breeding of
genetically-modified organisms (GMOs) such as safflower oil (1.010
4. Tribological performances of biolubricant base stocks unsaturation) in work [16] do not follow the trends of other natural
oils, probably due to their distinct structures in terms of their fatty
Base stock is the primary element which determines the final acids. On the other hand, short-chained VOs such as coconut oil are not
properties of a lubricant product. Even with the best additives, a poorly efficient as far as frictional properties are concerned [17]. Should a
performing base stock can only be improved with difficulty. The se- thicker fluid be required for an application such as hydraulic fluid,
lection of suitable base stock either in its neat form or in a combination epoxidized VO may be considered, as it has about twice the viscosity
of various types of base stocks (blend or emulsion) is usually based on a value of the original VO [26]. It also performs better in COF and WSD
list of properties, viz. physical properties, material compatibility, re- relative to commercial hydraulic oils within the same viscosity range
newability, degradability and so on. As one of these properties, tribo- [26]. Another alternative to resolve the low viscosity problem is to
logical performance is the most critical criterion. Recognizing its im- incorporate estolide as a thickening agent, as seen in the case whereby
portance, Table 1 tabulates the tribological performances of various oleic acid-based and ricinoleic acid-based estolides can be blended into
types of lubricant base stocks. Their key tribological behaviors will be a low viscosity fluid to improve wear performance [35]. When lu-
elucidated subsequently. bricating at extreme operating temperatures and enduring a heavy

151
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

contact load, biolubricant base stocks with significant molecular achieve what the current conventional additives can offer. On the other
branching and steric hindrance, namely the ring-opened products from hand, ethylene-vinyl acetate copolymer (EVA) and ethyl cellulose (EC)
epoxidized oil and the polyol ester, can be employed. The molecular are known as viscosity modifiers. Even when constituting only a slight
branching resists crystallization at a low temperature and degradation weight percentage of the additives in a base stock, they are able to boost
at an elevated temperature and pressure, producing stable tribofilm the viscosity of base stock at least by two-fold [15,48]. They are also
which are versatile and may be used for various applications [53,116]. able to form a thicker film for better frictional, wear, and load-carrying
According to the data tabulated in Table 1, it is suggested that the properties [15,50]. In fact, both polymers excel in different applica-
dominant molecular structure criteria affecting lubricity and wear tions. EVA helps to reduce friction and wear mainly in the mixed lu-
protection at boundary lubrication is the degree of molecular brication regime, whereas EC is more effective in extreme boundary
branching, followed by molecular chain length, and number of ester lubrication [15]. In view of other improvements [50], EC is able to
functional groups, in descending order. The latter two parameters are sustain tribofilm at a high temperature, in addition to acting as an ex-
more profound in mixed/EHL and hydrodynamic regimes as they affect cellent pour point depressant, which is proven to be more effective than
the viscosity of the fluid and the adhesion of tribofilm, respectively EVA and mineral oil additives with a polymethacrylate backbone. The
[101]. It should be noted that the effect of viscosity on friction is not above advantages are nonetheless dependent on the solubility of the
significant in boundary lubrication [98]. The above observation is ob- additives in the base stock. In one study [50], blending high oleic
tained based on the following cases: The triester product of ring- sunflower oil with a more polar oil, e.g. castor oil, can boost the solu-
opening of epoxidized oleic acid produced lower COF and smaller WSD bility of EC. This blend became a potentially eco-friendly base stock
in boundary lubrication as compared to the monoester and the diester which was suitable for various applications due to a wide range of
[29]. This comparison was done at the basis of similar chain lengths viscosity, good low-temperature properties and low degree of tem-
attached to the ester functional group. In addition, the ring-opened perature-viscosity dependency.
product of epoxidized canola oil with branched alcohol such as 2-
ethylhexanol exhibits better wear protection when blended into a low 5.2. Particulate and layered materials
lubricity oil, compared with that blended with other more viscous
products derived from linear alcohols [30]. In the studies involving Nano-sized particulates such as CuO, TiO2 and ZnO are non-toxic
polyol esters, PE ester gives better lubricity than TMP and NPG esters, anti-wear additives for biolubricants. With their sizes ranging from
especially at an elevated temperature (e.g. 100 °C) and under extreme around 2 to 120 nm [3], these materials are able to coalesce in asperity
pressure (e.g. 100 kg load). This is because PE possesses a higher valleys, creating a thin, smooth and solid lamellar film between con-
branching degree and therefore a much more stable molecular structure tacting surfaces [52,57]. As a result, these solid particulates provide
[36,37]. TMP and NPG esters may perform better at mild operating superior frictional and wear properties [52,57]. Their performance
conditions, or in mixed/EHL lubrication regimes. In actual fact, it is depends very much on the surface roughness of the sliding materials. If
difficult to isolate and investigate the tribological performances of NPG, the particulate size is slightly smaller than the surface roughness, the
TMP and PE esters, since their molecular weight and degree of steric wear volume decreases with the decrease of particulate size [54,57].
hindrance are varying, giving rise to different viscosities, thermal sta- When the particulate size is extremely small (e.g. five-fold smaller)
bility, and cold flow property [36,37]. Each type of ester has its own relative to the surface roughness, the wear volume increases with the
advantages: they can be designed and tailored to give specific physical decrease in the particle size [57]. In the case whereby particulate size is
properties and tribological advantages to suit a specific application. larger than the surface roughness, grooves from abrasions are produced
[57]. Another parameter affecting frictional and wear behavior is the
5. Tribological enhancement using biolubricant additives concentration of the additive. The presence of excessive additive par-
ticles in the asperities valleys could either give insignificant improve-
In the current scenario, environmental concerns continue to drive ment in tribological performance [56], or produce abrasive wear
original equipment manufacturers to change their specifications for (ploughing effects) by increasing the shear-concentration at the contact
lubricant performances. Lubricant formulation has been developed in surface [60]. The latter may be due to the interference of the additive
such a way that the elements of chlorine, phosphorus, sulfur, and me- particles causing poor adsorption of base stock at the contact area re-
tals in the formulation are already reduced [47]. This helps to mitigate sulting in inadequate lubrication, as observed in the case of nano-
environmental impact, as it is well known that the toxic and not readily particles in polar oils such as ester-based oil [56] and VO [128]. The
degradable additives may accumulate in the environment and harm the above relationships are also applicable for two-dimensional (2D)
ecosystem [127]. There are various types of additives available nowa- layered materials, including hexagonal boron nitride (hBN), boric ni-
days, but most of them are not environmentally friendly. For instance, tride and graphene [54,60]. The nature of 2D layered materials and the
zinc dialkyl dithiophosphate (ZDDP) has been used in engines both as particulates is different, whereby the former are known for their low
an anti-wear and antioxidant additive for decades, despite its non-de- interlayer friction ability and the ability to accommodate relative sur-
gradability and toxicity in nature [3,98]. As far as environmental face velocity to produce low COF during sliding [57,58]. Taking gra-
friendliness is concerned, the tribological additives for biolubricant phene as an example [58], its atoms lying on the same layer are closely
which are currently receiving most attention from researchers are bonded together by strong covalent bonds, while the layers are attached
plant-derived compounds and polymers, particulate and layered mate- together by weak Van der Waals. When employing nanoparticles and
rials and ionic liquids. Their tribological performances are tabulated in layered materials additives, replenishment of the solid additives is re-
Table 2. These additives will be briefly discussed in the following quired as the additive may be molten, welded or forced out of the tri-
subsections. bosystem after a certain time. Therefore, an effective suspension or
colloidal system that is stable and compatible with the solid additives is
5.1. Plant-derived compounds and polymers crucial.

Plant-derived compounds such as cystine schiff base ester are po- 5.3. Ionic liquids
tential anti-friction, anti-wear and anti-corrosion additives [51]. They
are comprised of disulfide groups, along with amine and carboxylic Ionic liquids have been applied in almost every field, including lu-
functionalities, which are capable of forming a surface-complex film, brication, due to their limitless pool of structural variations and prop-
particularly when metal-to-metal contact is involved [51]. As this type erties. Among the variety of ionic liquids, imidazolium-based ionic li-
of additive is relatively new, its tribological properties are yet to quid is claimed to have a great potential and to be a versatile lubricant

152
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Table 2
Tribological enhancement of biolubricants using additives.
a a
Ref. Tribosystem Lubrication conditions Base stock, viscosity (40 °C) COF WSD / WCV

[15] Ball-on-disk test (HFRR): ball (AISI 52100 4 N, 50 Hz (1 mm stroke), T high-oleic sunflower oil (HOSO), 38.5 cSt 0.049 0.18 mm
steel, 800 VPN hardness, 6 mm diameter); HOSO with 4% w/w ethylene-vinyl acetate copolymer (EVA), 0.047 0.19 mm
disk (AISI 52100 steel, 750 VPN, 10 mm 168.14 cSt
diameter) HOSO with 1% w/w ethyl cellulose (EC), 50.6 cSt 0.075 0.20 mm
T = 40 °C castor oil (CO), 242.5 cSt 0.087 0.34 mm
CO with 4% w/w EVA, 427.1 cSt 0.057 0.16 mm
CO with 1% w/w EC, 546.7 cSt 0.050 0.17 mm
T = 100 °C HOSO, 9.9 cSt b 0.085 0.20 mm
HOSO with 4% w/w EVA, 29.6 cSt b 0.067 0.22 mm
HOSO with 1% w/w EC, 15.1 cSt b 0.071 0.41 mm
CO, 21 cSt b 0.043 0.14 mm
CO with 4% w/w EVA, 50.9 cSt b 0.085 0.55 mm
CO with 1% w/w EC, 40.3 cSt b 0.034 0.16 mm
[50] Ball-on-disk test (minitraction machine 30 N, entrainment speed HOSO, 39.0 cSt 0.034 n/a
(MTM)): ball (AISI 52100 steel, 19.05 mm (50% slide-to-roll ratio), T HOSO with 1% w/w EC, 51 cSt 0.028 n/a
diameter 11 nm roughness); disk (AISI 52100 T = 40 °C CO, 242 cSt 0.033 n/a
steel, 50 mm diameter, 11 nm roughness) Entrainment speed = CO with 1% w/w EC, 547 cSt 0.030 n/a
0.01 m/s blend of HOSO and CO at ratio 1:1 w/w, n/a 0.027 n/a
blend of HOSO and CO at ratio 1:1 w/w with 1% w/w EC, 174 cSt 0.028 n/a
T = 40 °C HOSO, 39.0 cSt 0.021 n/a
HOSO with 1% w/w EC, 51 cSt 0.031 n/a
Entrainment speed = CO, 242 cSt 0.034 n/a
0.1 m/s CO with 1% w/w EC, 547 cSt 0.030 n/a
blend of HOSO and CO at ratio 1:1 w/w, n/a 0.025 n/a
blend of HOSO and CO at ratio 1:1 w/w with 1% w/w EC, 174 cSt 0.026 n/a
T = 40 °C HOSO, 39.0 cSt 0.021 n/a
HOSO with 1% w/w EC, 51 cSt 0.022 n/a
Entrainment speed = 1 m/s CO, 242 cSt 0.027 n/a
CO with 1% w/w EC, 547 cSt 0.024 n/a
blend of HOSO and CO at ratio 1:1 w/w, n/a 0.024 n/a
blend of HOSO and CO at ratio 1:1 w/w with 1% w/w EC, 174 cSt 0.025 n/a
T = 100 °C HOSO, 10 cSt b 0.052 n/a
HOSO with 1% w/w EC, 15 cSt b 0.016 n/a
Entrainment speed = CO, 21 cSt b 0.051 n/a
0.01 m/s CO with 1% w/w EC, 40 cSt b 0.051 n/a
blend of HOSO and CO at ratio 1:1 w/w, n/a 0.049 n/a
blend of HOSO and CO at ratio 1:1 w/w with 1% w/w EC, 26 cSt b 0.032 n/a
T = 100 °C HOSO, 10 cSt b 0.034 n/a
HOSO with 1% w/w EC, 15 cSt b 0.016 n/a
Entrainment speed = CO, 21 cSt b 0.022 n/a
0.1 m/s CO with 1% w/w EC, 40 cSt b 0.027 n/a
blend of HOSO and CO at ratio 1:1 w/w, n/a 0.016 n/a
blend of HOSO and CO at ratio 1:1 w/w with 1% w/w EC, 26 cSt b 0.016 n/a
T = 100 °C HOSO, 10 cSt b 0.011 n/a
Entrainment speed = 1 m/s HOSO with 1% w/w EC, 15 cSt b 0.013 n/a
CO, 21 cSt b 0.015 n/a
CO with 1% w/w EC, 40 cSt b 0.014 n/a
blend of HOSO and CO at ratio 1:1 w/w, n/a 0.012 n/a
blend of HOSO and CO at ratio 1:1 w/w with 1% w/w EC, 26 cSt b 0.013 n/a
[51] Four-ball test: ball (AISI 52–100 steel, 198 N load, 1200 rpm, 75 °C polyol base oil, n/a 0.068 0.90 mm
12.7 mm diameter, 64–66 HRC hardness) polyol base oil with 1000 ppm cysteine schiff base ester prepared via 0.075 0.91 mm
two step synthesis using 3,5-di-t-butyl-4-hydroxybenzaldehyde
followed by lauroyl alcohol (CySBE), n/a
polyol base oil with 2000 ppm CySBE, n/a 0.068 0.81 mm
polyol base oil with 3000 ppm CySBE, n/a 0.055 1.00 mm
[52] Four-ball test: ball (AISI 52–100 steel, load, 1770 rpm, 25 °C polyol ester oil, 40.03 cSt n/a 0.58 mm
12.7 mm diameter, 64–66 HRC hardness) polyol ester oil with 1% w/w oleic acid, 39.27 cSt n/a 0.58 mm
polyol ester oil with 1% w/w CuO nanoparticles, 40.25 cSt n/a 0.55 mm
load = 40 kg polyol ester oil with 1% w/w MoS2 nanoparticles, 40.96 cSt n/a 0.48 mm
polyol ester oil with 1% w/w CuO nanoparticles and 1% w/w oleic n/a 0.54 mm
acid, 40.17 cSt
polyol ester oil with 1% w/w MoS2 nanoparticles and 1% w/w oleic n/a 0.51 mm
acid, 40.09 cSt
load = 160 kg polyol ester oil, 40.03 cSt n/a welded
polyol ester oil with 1% w/w oleic acid, 39.27 cSt n/a welded
polyol ester oil with 1% w/w CuO nanoparticles, 40.25 cSt n/a 3.26 mm
polyol ester oil with 1% w/w MoS2 nanoparticles, 40.96 cSt n/a 3.04 mm
polyol ester oil with 1% w/w CuO nanoparticles and 1% w/w oleic n/a 3.14 mm
acid, 40.17 cSt
polyol ester oil with 1% w/w MoS2 nanoparticles and 1% w/w oleic n/a 2.88 mm
acid, 40.09 cSt
(continued on next page)

153
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Table 2 (continued)
a a
Ref. Tribosystem Lubrication conditions Base stock, viscosity (40 °C) COF WSD / WCV

[56] Pin-on-disk test: pin (AISI 1020 steel, load, 0.5 m/s, T mineral oil, 103.55 cSt 0.063 n/a
7.92 mm diameter); disk (AISI 52100 steel, T = 40 °C mineral oil with 0.3% w/w Cu nano-additive, 107.10 cSt 0.014 n/a
100 mm diameter, 64–66 HRC roughness) mineral oil with 3.0% w/w Cu nano-additive, 107.98 cSt 0.006 n/a
load = 392 N synthetic ester, 107.54 cSt 0.009 n/a
synthetic ester with 0.3% w/w Cu nano-additive, 108.50 cSt 0.004 n/a
synthetic ester with 3.0% w/w Cu nano-additive, 110.88 cSt 0.008 n/a
T = 40 °C mineral oil, 103.55 cSt 0.067 n/a
mineral oil with 0.3% w/w Cu nano-additive, 107.10 cSt 0.012 n/a
mineral oil with 3.0% w/w Cu nano-additive, 107.98 cSt 0.080 n/a
load = 588 N synthetic ester, 107.54 cSt 0.027 n/a
synthetic ester with 0.3% w/w Cu nano-additive, 108.50 cSt 0.006 n/a
synthetic ester with 3.0% w/w Cu nano-additive, 110.88 cSt 0.090 n/a
T = 100 °C mineral oil, 11.55 cSt b 0.111 n/a
mineral oil with 0.3% w/w Cu nano-additive, 11.87 cSt b 0.106 n/a
mineral oil with 3.0% w/w Cu nano-additive, 12.06 cSt b 0.107 n/a
load = 588 N synthetic ester, 14.70 cSt b 0.077 n/a
synthetic ester with 0.3% w/w Cu nano-additive, 14.81 cSt b 0.074 n/a
synthetic ester with 3.0% w/w Cu nano-additive, 15.05 cSt b 0.078 n/a
Four-ball test: ball (AISI 52100 steel, 12.7 mm 392 N, 1200 rpm, 75 °C mineral oil, 103.55 cSt n/a 0.88 mm
diameter, 64–66 HRC hardness) mineral oil with 0.3% w/w Cu nano-additive, 107.10 cSt n/a 0.70 mm
mineral oil with 3.0% w/w Cu nano-additive, 107.98 cSt n/a 0.74 mm
synthetic ester, 107.54 cSt n/a 0.79 mm
synthetic ester with 0.3% w/w Cu nano-additive, 108.50 cSt n/a 0.86 mm
synthetic ester with 3.0% w/w Cu nano-additive, 110.88 cSt n/a 1.11 mm
[57] Pin-on-disk test: pin (oxygen-free electronic 10 N, 33 mm/s, 25 °C canola oil mixed with 5% w/w of 5μm hexagonal boron nitride (hBN) 0.123 0.132 mm3
copper C101, 6.35 mm diameter, 50 mm particles, n/a
length, hemispherical tip); disk (aluminum canola oil mixed with 5% w/w of 1.5μm hBN particles, n/a 0.078 0.080 mm3
alloy 2024, 70 mm diameter, 6.35 mm canola oil mixed with 5% w/w of 0.5 μm hBN particles, n/a 0.056 0.069 mm3
thickness, Ra roughness) canola oil mixed with 5% w/w of 0.07μm hBN particles, n/a 0.035 0.025 mm3
Ra = 0.22 μm
Ra = 0.43 μm canola oil mixed with 5% w/w of 5μm hBN particles, n/a 0.072 0.359 mm3
canola oil mixed with 5% w/w of 1.5μm hBN particles, n/a 0.073 0.392 mm3
canola oil mixed with 5% w/w of 0.5 μm hBN particles, n/a 0.090 0.465 mm3
canola oil mixed with 5% w/w of 0.07μm hBN particles, n/a 0.093 0.434 mm3
[60] Four-ball test: ball (AISI 52–100 steel, 392 N, 1200 rpm, 75 °C jatropha oil-based TMP triester, 16.87 cSt 0.022 1.280 μm
12.7 mm diameter, 64–66 HRC hardness jatropha oil-based TMP triester with 0.05% w/w of 2–5 μm hBN 0.022 1.238 μm
particles, 17.13 cSt
jatropha oil-based TMP triester with 0.1% w/w of 2–5 μm hBN 0.026 1.514 μm
particles, 17.20 cSt
jatropha oil-based TMP triester with 0.5% w/w of 2–5 μm hBN 0.029 1.590 μm
particles, 17.56 cSt
commercial synthetic ester, 19.05 cSt 0.090 1.553 μm
[63] Four-ball test: ball (GCr15 steel, 61–64 HRC load, 1450 rpm, 75 °C Ionic liquid emulsion consists of 6% w/w I-butyl-3-methyl- 0.079 0.652 mm
hardness) load = 392 N imidazolium tetrafluoroborate [BMIM][BF4], 19% w/w castor oil,
60% w/w Tx-100 surfactant, 15% w/w 1-butanol, n/a
Commercial lubricant base oil, n/a 0.107 0.801 mm
load = 588 N Ionic liquid emulsion consists of 6% w/w [BMIM][BF4], 19% w/w 0.054 0.859 mm
castor oil, 60% w/w Tx-100 surfactant, 15% w/w 1-butanol, n/a
Commercial lubricant base oil, n/a 0.098 0.842 mm
load = 784 N Ionic liquid emulsion consists of 6% w/w [BMIM][BF4], 19% w/w 0.064 0.901 mm
castor oil, 60% w/w Tx-100 surfactant, 15% w/w 1-butanol, n/a
Commercial lubricant base oil, n/a 0.088 0.883 mm
[62] Ball-on-disk test (HFRR): ball (steel); disk 100 N, 25 Hz (1 mm stroke), trimethylolpropane oleate (TMPTO), 49.6 cSt 0.143 0.005 mm3
(steel) 150 °C TMPTO with 3% 1-tetradecyl-3-(2-ethylhexyl)imidazolium bis(2- 0.122 0.0098 mm3
ethylhexyl) phosphate [TDEHIM][DEHP], 51.23 cSt
TMPTO with 5% [TDEHIM][DEHP], 52.5 cSt 0.114 0.0057 mm3
TMPTO with 7% [TDEHIM][DEHP], 53.37 cSt 0.112 0.0095 mm3
[64] Four-ball test: ball (AISI 52–100 steel, 392 N, 1200 rpm, 75 °C polyol ester oil, 45.82 cSt 0.132 0.954 mm
12.7 mm diameter, 64–66 HRC hardness) polyol ester oil with 2% w/w aspartic acid and glutamic acid derived 0.0681 0.750 mm
ionic liquids (ILs) 1, 46.28 cSt
polyol ester oil with 2% w/w ILs 2, 46.38 cSt 0.0681 0.562 mm

a
Values obtained at steady state after sufficient sliding duration
b
Viscosity at 100 °C

which is applicable to various types of sliding materials [129,130]. On microemulsion [61,63]. For example, ionic liquid microemulsion
top of its unique properties such as low volatility, non-flammability, comprising of 1-butyl-3-methyl-imidazolium tetrafluoroborate, [BMIM]
negligible vapor pressure, and superior chemical and thermal stabilities [BF4], a surfactant, and a vegetable oil, produced better tribological
[130], this ionic liquid possesses excellent lubricating properties, in- results than conventional base stock [63]. In the latest development, the
cluding friction reduction, anti-wear and high load-carrying capacity dual roles of ionic liquid can been seen in the synthesis of polyol ester
[130–132]. Nevertheless, the immiscibility of ionic liquid in non-polar base stocks, whereby ionic liquid, namely 1-tetradecyl-3-(2-ethylhexyl)
mediums like fatty hydrocarbon and vegetable oil is the major draw- imidazolium bis(2-ethylhexyl) phosphate [TDEHIM][DEHP], acts as an
back. This technical restriction can be overcome with ionic liquid efficient catalyst and at the same time works as an in situ anti-wear

154
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Table 3
Tribological influences of biolubricant base stocks and additives.
Effectsa Base stocksb Additivesc

Increase Increase chain length Increase Increase branching Plant-derived Particulate and Ionic
unsaturation (molecular weight) polarity degree (steric polymer (EC, layered materialse liquid
hindrance) EVA)d

Tribofilm characteristics
Film thickness + + +
Adhesion strength – + – –
Low temperature flow + – – + +
Viscosity index +
Thermo-oxidative stability – – +
Tribological performance
Lubricity – + + + + +
Wear protection – + + + + +
Load-carrying capacity + + +
(extreme pressure)
Application
Boundary lubrication + + + + + +
Mixed lubrication + + + + +
High temperature – – +

a
Positive effect [+ ]; negative effect [-]; unknown or insignificant effect [].
b
Base stocks: vegetable oil (VO), epoxidized VO, ring opening products from epoxidized VO, estolides and polyol esters.
c
Effectiveness depends on solubility and compatibility.
d
EC: ethyl cellulose; EVA: ethylene-vinyl acetate copolymer.
e
Effectiveness depends on topological condition of contact surface.

additive, eliminating the need for catalyst separation from the synthe- References
sized ester [62]. In addition, biodegradable and non-toxic amino-acid
based ionic liquids have been synthesized as anti-wear additive and [1] European Committee for Standardization. Liquid petroleum products-Bio-lu-
friction modifier. They are able to form a stable homogeneous disper- bricants-Recommendation for terminology and characterisation of bio-lubricants
and bio-based lubricants. CEN/TR 162272011.
sion with polyol ester base stock [64]. [2] European Committee for Standardization. Liquid Petroleum Products-Bio-
Lubricants-Criteria And Requirements Of Bio-Lubricants And Bio-Based Lubricants.
prEN 168072014.
6. Final remarks [3] Rudnick LR. Synthetics, mineral oils, and bio-based lubricants: chemistry and
technology. New York: CRC Press; 2005.
It is clear, then, that biolubricants derived from renewable energy [4] Hinman ML. Environmental characteristics of fuel and lubricants. In: Totten GE,
Westbrook SR, Shah RJ, editors. Fuels and lubricants handbook: technology,
sources like plant oils are becoming more viable with the development properties, performance, and testing. West Conshohocken: ASTM International;
of various types of environmentally friendly base stocks and additives. 2003. p. 885–909.
They are able to perform as well as conventional lubricants, and can [5] European Commission. Directive 2009/28/EC of the European Parliament and of
the Council of 23 April 2009 on the promotion of the use of energy from renewable
even be applied and incorporated into broader applications, including sources and amending and subsequently repealing Directives 2001/77/EC and
biofuels and bioenergy. A successful formulation of biolubricant relies 2003/30/EC. Official Journal of the European Union; 2009. p. 16-62.
on the selection of suitable biolubricant base stocks and additives to [6] Gill G Slowly, The World’s Lube Thirst Grows. Lubes‘n’Greases; 2014. January
issue. p. 37-42.
meet a list of specifications required by an application. In terms of the [7] Markets And Markets. Lubricants Market by Type (Mineral Oil, Synthetic
tribological aspect, a good biolubricant base stock can be ensured by Lubricants, Bio-Based, and Greases), by Application (Transportation and Industrial
optimizing two characteristics: the stability and the adhesiveness of its Machinery & Equipment), and by Region (APAC, EU, NA, MEA, AND SA) - Global
Forecast to 2021.
generated tribofilm within the lubricating conditions, provided that the
[8] Infiniti Research Limited. Global Biolubricant Market 2015–2019; 2015.
viscosity range of the lubricant and the lubrication regime are suitable [9] Grand View Research. Biolubricants Market Analysis By Raw Material (Vegetable
and tally with the application requirements. Tribological additives can & Animal Oil), By Application (Automotive (Automotive Engine Oils, Gear Oils,
be then added to further improve the base stock. In summary, this re- Hydraulic Oils, Transmission Fluids, Greases, Chainsaw Oils), Industrial (Process
Oils, Demolding Oils, Industrial Gear Oils, Industrial Greases, Metal Working
view offers an in-depth discussion and comprehensive evidence of the Fluids)), By End-Use (Industrial, Commercial Transportation, Consumer
tribological behaviors of biolubricant base stocks and additives. The Automotive) Segment Forecasts To 2024; 2016.
influences of biolubricant base stocks and additives on tribological [10] Sethuramiah A, Kumar R. Modeling of chemical wear. Oxford: Elsevier; 2016. p.
185–211.
performance are summarized in Table 3. It is hoped that the tribological [11] Srivastava SP. Advances in lubricant additives and tribology. New Delhi: CRC
data presented in this review may assist interested parties in preparing Press; 2010.
and formulating biolubricants which fulfill basic lubrication require- [12] Bart JCJ, Gucciardi E, Cavallaro S. Biolubricants: science and technology. Oxford:
Woodhead Publishing; 2013. p. 351–95.
ments. [13] Chiong Ing T, Rafiq AKM, Azli Y, Syahrullail S. Tribological behaviour of refined
bleached and deodorized palm olein in different loads using a four-ball tribotester.
Sci Iran 2012;19:1487–92.
Acknowledgement [14] Syahrullail S, Kamitani S, Shakirin A. Performance of vegetable oil as lubricant in
extreme pressure condition. Procedia Eng 2013;68:172–7.
The author thanks the Director-General of MPOB for permission to [15] Quinchia LA, Delgado MA, Reddyhoff T, Gallegos C, Spikes HA. Tribological stu-
dies of potential vegetable oil-based lubricants containing environmentally
publish this work. friendly viscosity modifiers. Tribol Int 2014;69:110–7.
[16] Reeves CJ, Menezes PL, Jen T-C, Lovell MR. The influence of fatty acids on tri-
bological and thermal properties of natural oils as sustainable biolubricants. Tribol
Declarations of interest Int 2015;90:123–34.
[17] Rani S, Joy ML, Nair KP. Evaluation of physiochemical and tribological properties
none of rice bran oil - biodegradable and potential base stoke for industrial lubricants.

155
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

Ind Crops Prod 2015;65:328–33. multifunctional additive for blends of vegetable oil-based lubricants. J Clean Prod
[18] Kalam MA, Masjuki HH, Cho HM, Mosarof MH, Mahmud MI, Chowdhury MA, 2017;151:1–9.
Zulkifli NWM. Influences of thermal stability, and lubrication performance of [51] Singh RK, Pandey S, Saxena RC, Thakre GD, Atray N, Ray SS. Study of cystine
biodegradable oil as an engine oil for improving the efficiency of heavy duty diesel schiff base esters as new environmentally benign multifunctional biolubricant
engine. Fuel 2017;196:36–46. additives. J Ind Eng Chem 2015;26:149–56.
[19] Ruggiero A, D’Amato R, Merola M, Valašek P, Müller M. Tribological character- [52] Gulzar M, Masjuki HH, Varman M, Kalam MA, Mufti RA, Zulkifli NWM, et al.
ization of vegetal lubricants: comparative experimental investigation on Jatropha Improving the AW/EP ability of chemically modified palm oil by adding CuO and
curcas L. oil, Rapeseed Methyl Ester oil, Hydrotreated Rapeseed oil. Tribol Int MoS2 nanoparticles. Tribol Int 2015;88:271–9.
2017;109:529–40. [53] Zulkifli NWM, Kalam MA, Masjuki HH, Yunus R. Experimental analysis of tribo-
[20] Fox NJ, Tyrer B, Stachowiak GW. Boundary lubrication performance of free fatty logical properties of biolubricant with nanoparticle additive. Procedia Eng
acids in sunflower oil. Tribol Lett 2004;16:275–81. 2013;68:152–7.
[21] Kalin M, Vižintin J. A comparison of the tribological behaviour of steel/steel, [54] Lovell MR, Kabir MA, Menezes PL, Higgs CF. Influence of boric acid additive size
steel/DLC and DLC/DLC contacts when lubricated with mineral and biodegradable on green lubricant performance. Philos Trans A Math Phys Eng
oils. Wear 2006;261:22–31. 2010;368:4851–68.
[22] Willing A. Lubricants based on renewable resources – an environmentally com- [55] Tan K-H, Awala H, Mukti RR, Wong K-L, Ling TC, Mintova S, et al. Zeolite na-
patible alternative to mineral oil products. Chemosphere 2001;43:89–98. noparticles as effective antioxidant additive for the preservation of palm oil-based
[23] Salimon J, Salih N, Yousif E. Improvement of pour point and oxidative stability of lubricant. J Taiwan Inst Chem E 2016;58:565–71.
synthetic ester basestocks for biolubricant applications. Arab J Chem [56] Guzman Borda FL, Ribeiro de Oliveira SJ, Seabra Monteiro Lazaro LM, Kalab
2012;5:193–200. Leiróz AJ. Experimental investigation of the tribological behavior of lubricants
[24] Erhan SZ, Asadauskas S. Lubricant basestocks from vegetable oils. Ind Crops Prod with additive containing copper nanoparticles. Tribol Int 2018;117:52–8.
2000;11:277–82. [57] Reeves CJ, Menezes PL, Lovell MR, Jen T-C. The influence of surface roughness
[25] Erhan SZ, Sharma BK, Perez JM. Oxidation and low temperature stability of ve- and particulate size on the tribological performance of bio-based multi-functional
getable oil-based lubricants. Ind Crops Prod 2006;24:292–9. hybrid lubricants. Tribol Int 2015;88:40–55.
[26] Silva MS, Foletto EL, Alves SM, de Castro Dantas TN, Dantas Neto AA. New hy- [58] Zhou Y, Dahl J, Carlson R, Liang H. Effects of molecular structures of carbon-based
draulic biolubricants based on passion fruit and moringa oils and their epoxy. Ind molecules on bio-lubrication. Carbon 2015;86:132–8.
Crops Prod 2015;69:362–70. [59] Hu X. On the size effect of molybdenum disulfide particles on tribological per-
[27] Salih N, Salimon J, Yousif E. Synthetic biolubricant basestocks based on en- formance. Ind Lubr Tribol 2005;57:255–9.
vironmentally friendly raw materials. J King Saud Univ Sci 2012;24:221–6. [60] Talib N, Nasir RM, Rahim EA. Tribological behaviour of modified jatropha oil by
[28] Campanella A, Rustoy E, Baldessari A, Baltanás MA. Lubricants from chemically mixing hexagonal boron nitride nanoparticles as a bio-based lubricant for ma-
modified vegetable oils. Bioresour Technol 2010;101:245–54. chining processes. J Clean Prod 2017;147:360–78.
[29] Salih N, Salimon J, Yousif E. The physicochemical and tribological properties of [61] Wang A, Chen L, Jiang D, Zeng H, Yan Z. Vegetable oil-based ionic liquid micro-
oleic acid based triester biolubricants. Ind Crops Prod 2011;34:1089–96. emulsion biolubricants: effect of integrated surfactants. Ind Crops Prod
[30] Madankar CS, Dalai AK, Naik SN. Green synthesis of biolubricant base stock from 2014;62:515–21.
canola oil. Ind Crops Prod 2013;44:139–44. [62] Zhu L, Zhao Q, Wu X, Zhao G, Wang X. A novel phosphate ionic liquid plays dual
[31] Kulkarni RD, Deshpande PS, Mahajan SU, Mahulikar PP. Epoxidation of mustard role in synthetic ester oil: from synthetic catalyst to anti-wear additive. Tribol Int
oil and ring opening with 2-ethylhexanol for biolubricants with enhanced thermo- 2016;97:192–9.
oxidative and cold flow characteristics. Ind Crops Prod 2013;49:586–92. [63] Wang A, Chen L, Jiang D, Yan Z. Vegetable oil-based ionic liquid microemulsions
[32] Isbell TA, Cermak SC. Synthesis of triglyceride estolides from lesquerella and and their potential as alternative renewable biolubricant basestocks. Ind Crops
castor oils. J Am Oil Chem Soc 2002;79:1227–33. Prod 2013;51:425–9.
[33] Isbell TA, Lowery BA, DeKeyser SS, Winchell ML, Cermak SC. Physical properties [64] Nagendramma P, Khatri PK, Thakre GD, Jain SL. Lubrication capabilities of amino
of triglyceride estolides from lesquerella and castor oils. Ind Crops Prod acid based ionic liquids as green bio-lubricant additives. J Mol Liq
2006;23:256–63. 2017;244:219–25.
[34] Cermak SC, Isbell TA. Synthesis and physical properties of mono-estolides with [65] Blau PJ. The significance and use of the friction coefficient. Tribol Int
varying chain lengths. Ind Crops Prod 2009;29:205–13. 2001;34:585–91.
[35] García-Zapateiro LA, Franco JM, Valencia C, Delgado MA, Gallegos C. Viscous, [66] Guezmil M, Bensalah W, Mezlini S. Effect of bio-lubrication on the tribological
thermal and tribological characterization of oleic and ricinoleic acids-derived es- behavior of UHMWPE against M30NW stainless steel. Tribol Int 2016;94:550–9.
tolides and their blends with vegetable oils. J Int Eng Chem 2013;19:1289–98. [67] Maru MM, Tanaka DK. Consideration of stribeck diagram parameters in the in-
[36] Aziz NAM, Yunus R, Rashid U, Zulkifli NWM. Temperature effect on tribological vestigation on wear and friction behavior in lubricated sliding. J Braz Soc Mech Sci
properties of polyol ester-based environmentally adapted lubricant. Tribol Int Eng 2007;29:55–62.
2016;93:43–9. [Part A]. [68] Zia KM, Noreen A, Zuber M, Tabasum S, Mujahid M. Recent developments and
[37] Zulkifli NWM, Azman SSN, Kalam MA, Masjuki HH, Yunus R, Gulzar M. Lubricity future prospects on bio-based polyesters derived from renewable resources: a re-
of bio-based lubricant derived from different chemically modified fatty acid me- view. Int J Biol Macromol 2016;82:1028–40.
thyl ester. Tribol Int 2016;93:555–62. [Part B]. [69] McNutt J, He Q. Development of biolubricants from vegetable oils via chemical
[38] Zulkifli NWM, Kalam MA, Masjuki HH, Shahabuddin M, Yunus R. Wear prevention modification. J Ind Eng Chem 2016;36:1–12.
characteristics of a palm oil-based TMP (trimethylolpropane) ester as an engine [70] Prerna Singh C, VKC. Non-edible oil as a source of bio-lubricant for industrial
lubricant. Energy 2013;54:167–73. applications: a review. Int J Eng Sci Innov Technol 2013;2:299–305.
[39] Åkerman CO, Gaber Y, Ghani NA, Lämsä M, Hatti-Kaul R. Clean synthesis of [71] Nagendramma P, Kaul S. Development of ecofriendly/biodegradable lubricants: an
biolubricants for low temperature applications using heterogeneous catalysts. J overview. Renew Sust Energ Rev 2012;16:764–74.
Mol Catal B Enzym 2011;72:263–9. [72] Lu C, Napier JA, Clemente TE, Cahoon EB. New frontiers in oilseed biotechnology:
[40] Sripada PK, Sharma RV, Dalai AK. Comparative study of tribological properties of meeting the global demand for vegetable oils for food, feed, biofuel, and industrial
trimethylolpropane-based biolubricants derived from methyl oleate and canola applications. Curr Opin Biotechnol 2011;22:252–9.
biodiesel. Ind Crops Prod 2013;50:95–103. [73] Schneider MP. Plant‐oil‐based lubricants and hydraulic fluids. J Sci Food Agric
[41] da Silva JAC, Soares VF, Fernandez- Lafuente R, Habert AC, Freire DMG. 2006;86:1769–80.
Enzymatic production and characterization of potential biolubricants from castor [74] Panchal TM, Patel A, Chauhan DD, Thomas M, Patel JV. A methodological review
bean biodiesel. J Mol Catal B Enzym 2015;122:323–9. on bio-lubricants from vegetable oil based resources. Renew Sust Energ Rev
[42] Dodos GS, Karonis D, Zannikos F, Lois E. Renewable fuels and lubricants from 2017;70:65–70.
Lunaria annua L. Ind Crops Prod 2015;75:43–50. [Part B]. [75] Syahir AZ, Zulkifli NWM, Masjuki HH, Kalam MA, Alabdulkarem A, Gulzar M,
[43] Sethuramiah A, Kumar R. Chapter 2 - Lubricants and their formulation. Modeling et al. A review on bio-based lubricants and their applications. J Clean Prod
of chemical wear. Oxford: Elsevier; 2016. p. 25–39. 2017;168:997–1016.
[44] Born M, Hipeaux JC, Marchand P, Parc G. The relationship between chemical [76] Zainal NA, Zulkifli NWM, Gulzar M, Masjuki HH. A review on the chemistry,
structure and effectiveness of some metallic dialkyl- and diaryl-dithiophosphates production, and technological potential of bio-based lubricants. Renew Sust Energ
in different lubricated mechanisms. Lubr Sci 1992;4:93–116. Rev 2018;82:80–102.
[45] Watkins RC. The antiwear mechanism of zddp's. Part II. Tribol Int 1982;15:13–5. [77] Mariprasath T, Kirubakaran V. A critical review on the characteristics of alter-
[46] Kim B, Jiang JC, Aswath PB. Mechanism of wear at extreme load and boundary nating liquid dielectrics and feasibility study on pongamia pinnata oil as liquid
conditions with ashless anti-wear additives: analysis of wear surfaces and wear dielectrics. Renew Sust Energ Rev 2016;65:784–99.
debris. Wear 2011;270:181–94. [78] Rafiq M, Lv YZ, Zhou Y, Ma KB, Wang W, Li CR, et al. Use of vegetable oils as
[47] Kim B, Mourhatch R, Aswath PB. Properties of tribofilms formed with ashless di- transformer oils – a review. Renew Sust Energ Rev 2015;52:308–24.
thiophosphate and zinc dialkyl dithiophosphate under extreme pressure condi- [79] Kania D, Yunus R, Omar R, Abdul Rashid S, Mohamad Jan B. A review of biolu-
tions. Wear 2010;268:579–91. bricants in drilling fluids: recent research, performance, and applications. J Pet Sci
[48] Quinchia LA, Delgado MA, Valencia C, Franco JM, Gallegos C. Viscosity mod- Eng 2015;135:177–84.
ification of different vegetable oils with EVA copolymer for lubricant applications. [80] Myant C, Cann P. On the matter of synovial fluid lubrication: implications for
Ind Crops Prod 2010;32:607–12. Metal-on-Metal hip tribology. J Mech Behav Biomed Mater 2014;34:338–48.
[49] Martín-Alfonso JE, Valencia C. Tribological, rheological, and microstructural [81] Mobarak HM, Niza Mohamad E, Masjuki HH, Kalam MA, Al Mahmud KAH,
characterization of oleogels based on EVA copolymer and vegetables oils for lu- Habibullah M, et al. The prospects of biolubricants as alternatives in automotive
bricant applications. Tribol Int 2015;90:426–34. applications. Renew Sust Energ Rev 2014;33:34–43.
[50] Delgado MA, Quinchia LA, Spikes HA, Gallegos C. Suitability of ethyl cellulose as [82] Debnath S, Reddy MM, Yi QS. Environmental friendly cutting fluids and cooling

156
C.-H. Chan et al. Renewable and Sustainable Energy Reviews 93 (2018) 145–157

techniques in machining: a review. J Clean Prod 2014;83:33–47. stability. J Saudi Chem Soc 2011;15:195–201.
[83] Berman D, Erdemir A, Sumant AV. Graphene: a new emerging lubricant. Mater [107] Cermak SC, Isbell TA. Synthesis and physical properties of cuphea-oleic estolides
Today 2014;17:31–42. and esters. J Am Oil Chem Soc 2004;81:297–303.
[84] Lawal SA, Choudhury IA, Nukman Y. Application of vegetable oil-based me- [108] Aziz NAM, Yunus R, Rashid U, Syam AM. Application of response surface meth-
talworking fluids in machining ferrous metals—A review. Int J Mach Tool Manuf odology (RSM) for optimizing the palm-based pentaerythritol ester synthesis. Ind
2012;52:1–12. Crops Prod 2014;62:305–12.
[85] Shashidhara YM, Jayaram SR. Vegetable oils as a potential cutting fluid?An evo- [109] Yunus R, Fakhru’L-Razi A, Ooi TL, Iyuke SE, Idris A. Development ofoptimum
lution. Tribol Int 2010;43:1073–81. synthesis method for transesterification of palm oil methyl esters andtrimethy-
[86] John J, Bhattacharya M, Raynor PC. Emulsions containing vegetable oils for cut- lolpropane to environmentally acceptable palm oil-based lubricant. J Oil Palm Res
ting fluid application. Colloids Surf A Physicochem Eng Asp 2004;237:141–50. 2003;15:35–41.
[87] Singh Y, Farooq A, Raza A, Mahmood MA, Jain S. Sustainability of a non-edible [110] Choo YM, Cheng SF, Ma AN, Yusof B Lubricant base oil of palm origin. 2012; US
vegetable oil based bio-lubricant for automotive applications: a review. Process Saf patent 8101560 B2.
Environ Prot 2017;111:701–13. [111] Yeong Sk, Ooi TL, Ahmad S. Lubricant base from palm oil and its by-products; US
[88] Wan S, Tieu AK, Xia Y, Zhu H, Tran BH, Cui S. An overview of inorganic polymer patent 7781384 B2; 2010.
as potential lubricant additive for high temperature tribology. Tribol Int [112] Adhvaryu A, Biresaw G, Sharma BK, Erhan SZ. Friction behavior of some seed oils:
2016;102:620–35. biobased lubricant applications. Ind Eng Chem Res 2006;45:3735–40.
[89] Shahnazar S, Bagheri S, Abd Hamid SB. Enhancing lubricant properties by nano- [113] Briscoe WH. Aqueous boundary lubrication: molecular mechanisms, design
particle additives. Int J Hydrog Energy 2016;41:3153–70. strategy, and terra incognita. Curr Opin Colloid Interface Sci 2017;27:1–8.
[90] Dai W, Kheireddin B, Gao H, Liang H. Roles of nanoparticles in oil lubrication. [114] Lundgren SM, Persson K, Mueller G, Kronberg B, Clarke J, Chtaib M, et al.
Tribol Int 2016;102:88–98. Unsaturated fatty acids in alkane solution: adsorption to steel surfaces. Langmuir
[91] Saidur R, Kazi SN, Hossain MS, Rahman MM, Mohammed HA. A review on the 2007;23:10598–602.
performance of nanoparticles suspended with refrigerants and lubricating oils in [115] Sahoo RR, Biswas SK. Frictional response of fatty acids on steel. J Colloid Interface
refrigeration systems. Renew Sust Energ Rev 2011;15:310–23. Sci 2009;333:707–18.
[92] Hudson LK, Eastoe J, Dowding PJ. Nanotechnology in action: overbased nanode- [116] Mathiesen T. Modified TMP esters as multifunctional additives in metalworking
tergents as lubricant oil additives. Adv Colloid Interface Sci fluids. J Synth Lubr 1998;14:381–90.
2006;123–126:425–31. [117] Yunus R, Fakhru'l-Razi A, Ooi TL, Iyuke SE, Perez JM. Lubrication properties of
[93] Maleque MA, Masjuki HH, Sapuan SM. Vegetable‐based biodegradable lubricating trimethylolpropane esters based on palm oil and palm kernel oils. Eur J Lipid Sci
oil additives. Ind Lubr Tribol 2003;55:137–43. Tech 2004;106:52–60.
[94] Darminesh SP, Sidik NAC, Najafi G, Mamat R, Ken TL, Asako Y. Recent develop- [118] Kamalakar K, Manoj GNVTS, Prasad RBN, Karuna MSL. Influence of structural
ment on biodegradable nanolubricant: a review. Int J Heat Mass Transf modification on lubricant properties of sal fat-based lubricant base stocks. Ind
2017;86:159–65. Crops Prod 2015;76:456–66.
[95] Xiao H, Liu S. 2D nanomaterials as lubricant additive: a review. Mater Des [119] Havet L, Blouet J, Robbe Valloire F, Brasseur E, Slomka D. Tribological char-
2017;135:319–32. acteristics of some environmentally friendly lubricants. Wear 2001;248:140–6.
[96] Amiri M, Khonsari MM. On the thermodynamics of friction and wear―a review. [120] Salih N, Salimon J, Abdullah BM, Yousif E. Thermo-oxidation, friction-reducing
Entropy 2010;12:1021. and physicochemical properties of ricinoleic acid based-diester biolubricants. Arab
[97] (a) Bruce RW. Handbook of lubrication and tribology: theory and design. 2nd ed. J Chem 2014.
London: Taylor & Francis; 2012; [121] Martins RC, Cardoso NFR, Bock H, Igartua A, Seabra JHO. Power loss performance
(b) Bruce RW. Section 1: theory and practice of lubrication and tribology. 2nd ed. of high pressure nitrided steel gears. Tribol Int 2009;42:1807–15.
London: Taylor & Francis; 2012. [122] Martins R, Seabra J, Brito A, Seyfert C, Luther R, Igartua A. Friction coefficient in
[98] (a) Stachowiak GW, Batchelor AW. Engineering tribology. 4rth ed. Boston: FZG gears lubricated with industrial gear oils: biodegradable ester vs. mineral oil.
Butterworth-Heinemann; 2013; Tribol Int 2006;39:512–21.
(b) Stachowiak GW, Batchelor AW. Chapter 8: boundary and extreme pressure [123] Cardoso NFR, Martins RC, Seabra JHO, Igartua A, Rodríguez JC, Luther R.
lubrication. 4rth ed. Boston: Butterworth-Heinemann; 2013; Micropitting performance of nitrided steel gears lubricated with mineral and ester
(c) Stachowiak GW, Batchelor AW. Chapter 11: abrasive, erosive and cavitation oils. Tribol Int 2009;42:77–87.
wear. 4rth ed. Boston: Butterworth-Heinemann; 2013; [124] Tan CP, Che Man YB. Differential scanning calorimetric analysis of edible oils:
(d) Stachowiak GW, Batchelor AW. Chapter 12: adhesion and adhesive wear. 4rth comparison of thermal properties and chemical composition. J Am Oil Chem Soc
ed. Boston: Butterworth-Heinemann; 2013; 2000;77:143–55.
(e) Stachowiak GW, Batchelor AW. Chapter 14: fatigue wear. 4rth ed. Boston: [125] Lundgren SM, Ruths M, Danerlöv K, Persson K. Effects of unsaturation on film
Butterworth-Heinemann; 2013. structure and friction of fatty acids in a model base oil. J Colloid Interface Sci
[99] Barriga J, Kalin M, Acker KV, Vercammen K, Ortega A, Leiaristi L. Tribological 2008;326:530–6.
performance of titanium doped and pure DLC coatings combined with a synthetic [126] Hu Z-S, Hsu SM, Wang PS. Tribochemical reaction of stearic acid on copper surface
bio-lubricant. Wear 2006;261:9–14. studied by surface enhanced raman spectroscopy. Tribol Trans 1992;35:417–22.
[100] Menezes PL, Kishore, Kailas SV, Lovell MR. Role of surface texture, roughness, and [127] Hewstone RK. Environmental health aspects of lubricant additives. Sci Total
hardness on friction during unidirectional sliding. Tribol Lett 2011;41:1–15. Environ 1994;156:243–54.
[101] Bayer RG. Mechanical wear prediction and prevention. New York: Marcel Dekker; [128] Alves SM, Barros BS, Trajano MF, Ribeiro KSB, Moura E. Tribological behavior of
1994. vegetable oil-based lubricants with nanoparticles of oxides in boundary lubrica-
[102] Hamrock BJ, Dowson D. Ball bearing lubrication: the elastohydrodynamics of el- tion conditions. Tribol Int 2013;65:28–36.
liptical contacts. New York: Wiley-Interscience; 1981. [129] Liu W, Ye C, Gong Q, Wang H, Wang P. Tribological performance of room-tem-
[103] Papay AG. Antiwear and extreme-pressure additives in lubricants. Lubr Sci perature ionic liquids as lubricant. Tribol Lett 2002;13:81–5.
1998;10:209–24. [130] Minami I. Ionic liquids in tribology. Molecules 2009;14:2286.
[104] Jayadas NH, Prabhakaran Nair K, GA. Tribological evaluation of coconut oil as an [131] Yao M, Fan M, Liang Y, Zhou F, Xia Y. Imidazolium hexafluorophosphate ionic
environment-friendly lubricant. Tribol Int 2007;40:350–4. liquids as high temperature lubricants for steel–steel contacts. Wear
[105] Lathi PS, Mattiasson B. Green approach for the preparation of biodegradable lu- 2010;268:67–71.
bricant base stock from epoxidized vegetable oil. Appl Catal B 2007;69:207–12. [132] Schneider A, Brenner J, Tomastik C, Franek F. Capacity of selected ionic liquids as
[106] Salimon J, Salih N, Yousif E. Chemically modified biolubricant basestocks from alternative EP/AW additive. Lubr Sci 2010;22:215–23.
epoxidized oleic acid: improved low temperature properties and oxidative

157

You might also like