Open access peer-reviewed chapter

Ionic Liquids as High-Performance Lubricants and Lubricant Additives

Written By

Hong Guo and Patricia Iglesias Victoria

Submitted: 20 December 2020 Reviewed: 05 February 2021 Published: 16 March 2021

DOI: 10.5772/intechopen.96428

From the Edited Volume

Ionic Liquids - Thermophysical Properties and Applications

Edited by S. M. Sohel Murshed

Chapter metrics overview

495 Chapter Downloads

View Full Metrics

Abstract

Taking into account the environmental awareness and ever-growing restrictive regulations over contamination, the study of new lubricants or lubricant additives with high performance and low toxicity over the traditional lubes to reduce the negative impact on the environment is needed. In this chapter, the current literature on the use of ionic liquids, particularly protic ionic liquids, as high-performance lubricants and lubricant additives to different types of base lubricants are reviewed and described. The relation between ionic liquids structures and their physicochemical properties, such as viscosity, thermal stability, corrosion behavior, biodegradability, and toxicity, is elaborated. Friction reduction and wear protection mechanisms of the ionic liquids are discussed with relation to their molecular structures and physicochemical properties.

Keywords

  • ionic liquids
  • friction
  • wear
  • tribofilm
  • additives

1. Introduction

Friction and wear are inescapable problems in mechanical and electromechanical systems, resulting in massive energy losses. Holmberg and Erdemir [1] have estimated that the energy consumption generated by contacting surfaces in mechanical elements is almost 23% of the total energy consumption in the world, where 20% is used to overcome friction and 3% is used to replace worn surfaces. However, energy losses could be reduced by up to 40% through new advances in lubrication, which can save 8.7% of world energy consumption. Especially, the losses by friction could be decreased by using high-performance lubricants, which cannot only result in economic savings but also in important environmental benefits. In addition, the increase in energy prices leads to high demand for improvement of energy efficiency.

Ionic liquids (ILs) are a class of salts composing of bulky organic cations and organic or inorganic anions. Some of the typical IL molecular structures are shown in Figure 1. The large molecular size of the ions and their possible delocalized charge contribute to the uncommonly low melting points of ILs, which are below 100 °C. The first IL, ethylammonium nitrate [(C2H5NH3)NO3], reported by Walden in 1914 is found to have a melting point of 12 °C. Since the 1970s, the research of ILs has become increasingly popular and now ILs have been used for various applications such as effective solvent, catalyst, electrolytes in batteries, and carbon and carbon dioxide capturing.

Figure 1.

Typical ionic liquids molecular structures.

In the tribological field, ILs have shown great potential as advanced lubricants and “tailor-made” lubricating additives since been explored for lubrication in 2001 [2]. ILs have excellent physicochemical properties including low melting point, low flammability, negligible vapor pressure, and high thermal stability that meet the demands of high-performance lubricants. One of the most important characteristics of ILs is that their properties can be tailored by varying the species of the cations and anions, giving rise to numerous families that can be used across different tribological systems. The superiority of ILs in lubrication can be attributed to their inherent polarity, which can make them form stable ordered layers in the liquid state on metal surfaces to prevent them against contact; and that some elements of ILs can react with the substrate materials to generate a tribofilm to protect the substrate from further wear.

ILs can be conventionally categorized into aprotic ionic liquids (AILs) and protic ionic liquids (PILs), based on the nature of the cation present in the combinations [3]. Since most of the studies in lubrication are focused on AILs, many literature reviews have summarized the research efforts of them. Therefore, this chapter covers more about the progress of PILs in lubrication. Firstly, some important physicochemical properties of ILs will be introduced. Secondly, ILs as neat lubricants, specifically as bulk lubricants, thin-film lubricants, and the surface interactions between ILs and contact surfaces will be discussed. The third part will be focused on the ILs as additives in different base lubricants.

Advertisement

2. Physicochemical properties of ionic liquids

ILs are highly tunable by changing the cation structures, anion structures or both to satisfy specific engineering and manufacturing requirements. Therefore, to understand the relationship between the chemical structures of ILs and their physicochemical properties, as well as the tribological properties, becomes crucial to the molecular design of the more effective ILs. Their physicochemical properties can be easily influenced by combining different types of cations and anions or varying their alkyl chain lengths.

2.1 Viscosity and thermal stability

The viscosity behavior will affect the load-carrying capacity of ILs, as well as their formation of boundary lubricating films. In addition, the thermal stability of an IL is also a prerequisite for being used in various tribological systems. Particularly, an outstanding thermal stability would contribute to its application in high-temperature environments. Since thermogravimetric analysis (TGA) has been employed in most of the studies to characterize the thermal stability of ILs, the viscosity and onset thermal decomposition temperature (Td) of some ILs obtained using this method have been summarized in Table 1. In general, the molecular structure modification of the cation or the anion will affect the IL’s viscosity and thermal stability. ILs having symmetric cations with long alkyl side chains are found to have high viscosity [16, 17, 18], which is attributed to the closer packing and enhanced van der Waals interactions between the long alkyl chains. Particularly, the branched ILs are reported to possess higher viscosity than the linear ones [19]. In addition, ILs having high molar mass and ion-interactions such as hydrogen bonds in their molecular structures will get high viscosity [13, 20]. For instance, an imidazolium-based IL with a hydroxyl group (-OH) grafted into the N-1 position of its cation obtained an increase in viscosity, which is attributed to the increased hydrogen bond interactions and the resulting higher molar mass. In the study of Guo et al. [15], the viscosity of the hydroxylammonium PIL is highly dominated by the hydrogen bond interactions among its molecules instead of its molar mass.

CationAnionViscosity (25 °C)Viscosity (40 °C)Viscosity (100 °C)Viscosity
Index
Thermal Stability Td (°C)Ref
[C4C1im]NTf260.750.146413.85[4]
[C4im]129.152.146.2382.85
N4441NTf2∼480∼190∼20360[5]
N8886BScB∼190Ψ∼110∼20[6]
N8888∼690Ψ∼370∼50
N88810∼200Ψ∼120∼20
N88812∼110Ψ∼80∼20
N8881C6:02752.9Ψ1284.956.592[7]
[8]
C8:02410.3Ψ1121.248.685175.2
C12:01475.8Ψ715.736.985175.1
C16:01188.3Ψ596.337.499183.3
C18:01033.2Ψ524.235.1102
C18:11234.7Ψ627.239.3101
P66614C10:0141800*16278*16.91*268.65[9]
P66614DEHP1031^418.449.2∼300[10]
P66614(iC8)2PO21204.1528.9355.06169300.73[11]
P66614BEHP1156528.0559181293.48
P66614NTf2277.26123.4916.15140417.3
P44414DBS4423.11355.262.3498333.01
P4442DEP451.39171.5314.8183304.1
aCitrate321691*65097*941.21*191.2[12]
aSuccinate32798*8123.5*167.1*178.6[13]
bFormiate16.13150[14]
Pentanoate1333.2122
bHexanoate#8943.8*2405.56*59.4*177.48[15]
cHexanoate#926.86*301.58*14.72*174.82
dHexanoate#61.74*27.74*3.71*141.44

Table 1.

Kinetic viscosity and thermal stability of some ILs.

2-ethylhexanoate.


Dynamic viscosity (cP).


kinetic viscosity at 30 °C; and the numbers in imidazolium CxCxim, ammonium Nx,x,x,x, and phosphonium Px,x,x,x represent the alkyl chain length.


Note: a- NH2((CH2)2OH)2; b- NH3((CH2)2OH); c- NH2CH3(CH2)2OH); d- NH(CH3)2(CH2)2OH).

^kinetic viscosity at 23 °C.

The thermal stability of an IL is also closely related to its cation and anion, as can be seen in Table 1. Generally, when pairing with the same anion, the imidazolium-based ILs have a higher thermal stability than the tetraalkylphosphonium-based and the tetraalkylammonium-based ILs [4, 5, 9]. And the imidazolium-based ILs are reported to have a higher thermal stability when their cations have a smaller alkyl chain [21]. In contrast to cations, the anions have more impacts on the thermal stability of ILs. For example, an alkylammonium PIL derived from a stronger acid tends to have a higher thermal stability [22]. Recently, Fadeeva et al. [23] reported that the thermal stability of the alkylimidazolium-based PILs is mainly determined by the anions nature instead of the cation structure. However, with the same triflate anion, the PILs having the cation with larger size and branched chain structure would get a higher thermal stability [24, 25]. In several studies [13, 14, 15], the hydroxylammonium PILs with carboxylate anions were found to have low thermal stability. The reason underlaying this phenomenon is related to the reversal proton transfer, leading to the presence of free acid and ethanolamine [15]. Considering the long-term practical applications of ILs in lubrication, the characterization of their long-term thermal stability, through isothermal TGA, should also be concerned [21].

2.2 Corrosion

Corrosion for most lubricants, such as water-based lubricants, is a complex problem that needed to be solved. The corrosivity of neat ILs or ILs additives are usually evaluated by means of immersion corrosion test, in which the testing specimen, such as copper [26, 27, 28, 29], steel [28, 30], or cast iron [29] will be immersed into (or cover the metal surface with) neat ILs or IL containing lubricants. In addition, electrochemical corrosion tests will also be conducted to examine the anticorrosion properties of ILs as well as study the corrosion mechanisms [29, 31]. Through the two methods, the anticorrosion performance of four hydroxylammonium phosphate PILs additives was investigated in [28]. Compared to the reference sample immersed in water, the use of neat PILs significantly improved the corrosion resistance of the copper and iron sheets. What is more, the results from the electrochemical test further demonstrated the excellent corrosion inhibition of PILs additives and their attributes of anodic corrosion inhibitors. A protective film generated by the adsorption of PIL molecules, particularly the hydrophilic functional group on steel surface is attributed to anticorrosion performance. Among the four PILs, 2-hydroxypropylammonium di-(2-ethylhexyl) phosphate (TEOAP6) was found to have the best corrosion resistance, and PIL’s corrosion inhibition efficiency was related to its functional groups.

In the study of Ma et al. [26], the immersion corrosion test was employed to evaluate the anticorrosion property of PAO when ILs additives were added in different concentrations. Along with ILs, an additive containing Sulfur element was also added to PAO in 1%. It is noted that the addition of only 0.1% ILs can greatly reduce the corrosion tendency of a lubricant, and IL concentration (0.25–0.1%) just slightly affected the corrosion-inhibiting performance. Recently, a PIL 2-hydroxyethylammonium oleate [32], was proved to be an effective corrosion inhibitor for aluminum 1100 in neutral sodium chloride solution. The PIL adsorption layer on the aluminum substrate surface was pointed out to inhibit the diffusion of chloride anions during the electrochemical measurement, where it can also provide a corrosion protection at high chloride concentration for 72 hours.

2.3 Biodegradability and toxicity

In terms of the prospective large-scale industrial applications of ILs, it is crucial to examine their biodegradability and toxicity to control the discharge of IL-involved solvents or lubricants, minimizing the environmental damage. To understand the correlation between the molecular structure and biodegradability and toxicity of ILs is essential to design green IL lubricants. Nowadays, many experimental studies [7, 33] have been conducted to evaluate the environmental impact of ILs. In addition, computational approaches [34, 35, 36] are also employed to assess their toxicity. In general, cations and anions do have an influence on the ILs toxicity particularly, cations have a greater impact than anions. ILs having longer alkyl chain length and more branched-chain groups on their cations tend to be more toxic [37]. However, some anions containing fluorine in their structure will cause an increase in toxicity of their corresponding ILs. For example, although the hydroxylammonium and imidazolium cations were evaluated to be less toxic, the toxicity of their ILs increased drastically once NTf2 was incorporated as the anion [38]. Regarding the biodegradability, the IL components 1–Butyl–3–methylimidazolium (Bmim) and bis(trifluoromethanesulfonyl) imide (NTf2) were reported to be non-biodegradable even at low concentration (10 mg/L), while the N,N,N–trimethylethanolammonium (Choline) and acetate (Ac) could be completely degraded with a concentration up to 50 mg/L [38].

In the study of Tzani et al. [39], the biodegradability of a series of carboxylate-PILs were examined and proved to be relevant to the alkyl chain length of the anions. PILs having anions with long alkyl chain length were found to get a decreased biodegradability, except for the one that had an alicyclic ring in its anion, showing an enhanced biodegradability. Lately, Viesca et al. [33] characterized the biodegradability and bacteria toxicity of six PILs derived from alkylhydroxylamine. Owing to the presence of benzenesulfonate aromatic group in anions, the sulfonate-PILs were reported to be less biodegradable compared to the hexanoate-PILs. Regarding the bacterial toxicity behavior, even the hexanoate-PILs exhibited a better environmental impact, all of them were mild toxic to Vibrio fischeri. Nevertheless, all these PILs were found to outperform the traditional lubricant additive ZDDP concerning the biodegradability and toxicity performance.

Advertisement

3. Ionic liquids as lubricants

The tribological behavior of ILs as lubricants have been typically evaluated through laboratory bench tests using various macroscopic tribometers, such as the Optimol SRV series tribometers, mini-traction machines, Microtest pin-on-disk tribometer, Plint TE77 high-frequency reciprocating rigs, etc. In addition, the atomic force microscope (AFM) and surface force apparatus (SFA) are usually applied to investigate the nanotribological performance of lubricants. The two main factors, coefficient of friction (COF), and wear volume (or wear rate) of the rubbing materials are normally used to evaluate and compare the lubricating ability and anti-wear performance of IL lubricants.

3.1 Ionic liquids as neat lubricants

Since Ye et al. [2] initiated the study of ILs in lubrication in 2001, the studies about ILs as neat lubricants for various contact systems such as steel-steel contact [15], steel-ceramic contact [40], and steel-aluminum contact [14] have received considerable attention. Table 2 summarizes some recent studies of ILs as neat lubricants. Compared to AILs, the use of PILs as neat lubricants has gained more attention than before, owing to their low cost and facile synthesis process.

CationAnionTribo-pairContact modeLoad (N)COFRef
bHexanoate#Steel/steelBall-on-flat30.038[15]
c0.032
d0.58
bFormiateSteel/AlBall-on-plate0.50.35 ± 0.12[14]
Pentanoate0.14 ± 0.026
P888HCaprylateSteel/steelFour-ball3920.038[41]
Oleate0.044
bOleateAlumina/AlBall-on-plate0.50.11–0.15[42]
c0.12–0.15
a0.12–0.076
[C4C1im]BF4Steel/steelBall-on-disk20∼0.066[43]
PF6∼0.0.8
BF4400.06
PF60.07
BF4600.06
PF6∼0.075
aSuccinateSapphire/steelPin-on-disk0.980.119[40]
NHHH10OleateSteel/steelBall-on-plate40.048δ/0.054λ[27]
IL-TO0.069δ/0.056λ
a0.063δ/0.066λ
P66614(iC8)2PO20.069δ/0.093λ

Table 2.

Tribological results of some ILs as neat lubricants (2017–2020).

2-ethylhexanoate.


coefficient of friction at 30 °C.


coefficient of friction at 80 °C; and the numbers in imidazolium CxCxim, ammonium NHHHx, and phosphonium Px,x,x,x represent the alkyl chain length.


Note: a- NH2((CH2)2OH)2; b- NH3((CH2)2OH); c- NH2CH3(CH2)2OH); d- NH(CH3)2(CH2)2OH).

In Khan et al.’s research, two phosphonium-based PILs with different alkyl chain length in the anions were tested as neat lubricants under steel-steel contact, and a synthetic oil PEG 200 was used as a reference [41]. Since the fatty acid anions of PILs have a better affinity to steel surfaces, the use of PILs showed a significant friction reduction with respect to PEG 200. The tribological performance of the two PILs were found to be determined by the alkyl chain length of their anions and their viscosity, where a PIL with a shorter anion chain length and lower viscosity led to a lower friction coefficient but more material loss. While the results may be inverse once the experiment conditions are changed or other PILs are used. In the study of Vega et al. [14], the effect of anion chain length on the friction and wear behavior of ammonium-based PILs was investigated under steel-aluminum contact. The results revealed that increasing the anion chain length will improve the lubricating ability of PIL with a low friction coefficient. From another study of Vega et al. [42], three oleic-acid derived ammonium-based PILs were evaluated as lubricants in alumina-aluminum contact. In addition to the low friction coefficient, the use of PILs yielded an important wear reduction (98%) compared to the dry condition. Lately, the hexanoate-based PILs were also found to greatly reduce the wear of steel with respect to mineral oil as well as a commercial oil [15]. Tribofilms were detected on the worn steel disks when PILs were used to protect the steel against severe wear.

In addition to the above-mentioned bulk lubricants, ILs can also be employed in the form of thin layers for lubricating micro/nano electromechanical systems (MEMS/NEMS). For example, in Bermúdez’s group [40], a PIL - di[bis(2-hydroxyethyl)ammonium] succinate thin layer was created on a steel substrate surface by evaporating water from the PIL + Water mixture, where the PIL thin layer extremely reduced the wear rate of steel compared to the bulk neat PIL.

3.2 Surface interactions

As shown in Figure 2, it has been widely accepted that when neat ILs or IL additives are introduced between the contacting work pairs, the IL molecules tend to adsorb onto the workpiece surfaces physically or/and chemically and form an ordered boundary lubricating film to protect the moving components from direct contact, leading to low friction. During the sliding frictional process, a protective tribofilm will be subsequently generated on top of the substrate by means of the tribochemical reactions between ILs or their decomposition products and the contacting metal surfaces to reduce mechanical wear.

Figure 2.

Schematic diagram of (A) ILs boundary lubricating film, and (B) IL-induced tribofilm on the metal surface.

Although the process of forming the adsorbed boundary lubricating film is still not clear, the IL-adsorption film has been verified through electrical contact resistance (ECR) measurement by Viesca et al. [44]. The results showed that the IL-additive ([C6C1im][BF4]) outpaced the base oil to form a boundary film on the metal surface. The generation of the IL-tribofilm has been demonstrated on various material surfaces [45, 46, 47]. But the results from most of the work are relied on the post-analysis of the worn surfaces by employing Scanning Electron Microscopy (SEM), Transmission Electron Microscopy (TEM), Energy-dispersive X-ray Spectroscopy (EDS), Auger Electron Spectroscopy (AES), X-ray Photoelectron Spectroscopy (XPS), Raman Spectroscopy, etc. From the previous research [48], IL decomposition has been demonstrated during the sliding process, but only the anion was found to react with or adsorbed on the steel surface. Particularly, the IL undergoing facile decomposition would interact rapidly with the sliding surface, leading to a low friction coefficient. So the thermal stability of IL can be considered as an index for evaluating the tribo-decomposition behavior on nascent substrate surfaces [49].

Up to now, the characterization of the IL-induced tribofilm thickness, composition, and structure have been intensively investigated. For instance, when phosphonium-phosphate ILs were introduced to the base oils with a small amount (1.04 wt.%), an amorphous-nanocrystalline tribofilm with a 10–200 nm-thick was probed on the worn cast iron surface by TEM, EDS, and electron diffraction [17]. Furthermore, the participation of wear debris in the IL-tribofilm growth was proposed and demonstrated recently by Qu et al. [45, 46] through Atom Probe Tomography (APT) and Scanning Transmission Electron Microscopy (STEM) characterization.

In addition, tribofilm mechanical properties, such as hardness and resistance-to-plastic-deformation (P/S2), have also been investigated through nanoindentation measurements [50, 51]. The results revealed that only P/S2 had a correlation with the friction and wear performance, in which a small P/S2 value corresponded to a low friction and wear.

Regarding the growth mechanism of IL-induced tribofilm, a more precise in situ characterization is highly desirable in spite of many characterization approaches and spectroscopy techniques have been employed so far. At the same time, the application of the computational methods, such as molecular dynamic simulation, would help to elucidate the generation process of the boundary lubricating film.

Advertisement

4. Ionic liquids as lubricant additives

Limited to the high cost of being used as neat lubricants (particularly when AILs are used), ILs as additives have gained more and more research attention in recent years. Their highly tunable molecular structures and physicochemical properties make ILs suitable to be added to base lubricants with different nature (polar or nonpolar), such as ester, polyethylene glycol (PEG), PAO, mineral oils (MO), grease, and water-based lubricants.

Until now, ILs have been tested as friction-reducing additives, anti-wear additives, or extreme-pressure additives in many research articles. The tribological performance of IL as additives to non-polar, polar, and water-based lubricants have been summarized in Tables 35, respectively. Different from the traditional friction modifiers, ILs can be strongly adsorbed to the sliding surfaces and generate a resilient boundary lubricating film, leading to important reduction of friction and wear. Some active-elements containing ILs are easy to chemically react with the rubbing surfaces and create an effective tribofilm on top of the workpieces to prevent it against wear or extreme pressure.

CationAnionBase OilContent (wt.%)Tribo-pairContact modeLoad (N)COFRef
N121212PDOSSPAO101–3Steel/steelBall-on-disk500.1–0.12[52]
Laurate
DOSS2000.09–0.11
Laurate
N888HDEHPPAO40.87Steel/
bronze
Ball-on-flat200.09[53]
aSuccinatePAO401Steel/steelBall-on-flat30.085[13]
Citrate0.065
Succinate40.075
Citrate0.059
aCitrateMO1Steel/steelBall-on-flat20.1[12]
bHexanoate#MO1Steel/steelBall-on-flat30.084[80]
c0.085
d0.086
aSteel/Al0.101
b0.061
c0.046
P8888DEHPGTL41.04Steel/ironBall-on-flat1000.115[46]
P66614(iC8)2PO2MO1Steel/steelBall-on-flat20.09[54]
NTf20.127
P66614
DEHPPAO41Steel/
OD-Ti64
Ball-on-flat1000.05[55]
Stearate0.99100∼0.08
BTMPP0.99100∼0.07
N888HDEHP0.87100∼0.06
P88816DOSS500SN1–4Steel/AlBall-on-disk1000.11–0.125[47]
DOSSSteel/Mg1000.1–0.115
P888PDOSSSteel/Al1000.11–0.125
DOSSSteel/Mg1000.1–0.115
P66614(iC8)2PO2PAO40.5Steel/steelBall-on-plate600.015–0.02[56]
BEHP∼0.017
(iC8)2PO21∼0.015
BEHP∼0.014
TTAOADEHPPAO100.25–1Steel/steelBall-on-disk2000.107–0.11[26]
TTADO

Table 3.

Tribological results of ILs as additives for non-polar oils (2017–2020).

2-ethylhexanoate; and the numbers in ammonium NX,X,X,X, and phosphonium PX,X,X,X represent the alkyl chain length.


Note: a- NH2((CH2)2OH)2; b- NH3((CH2)2OH); c- NH2CH3(CH2)2OH); d- NH(CH3)2(CH2)2OH).

CationAnionBase OilContent (wt.%)Tribo-pairContact modeLoad (N)COFRef
bHexanoate#BO1/2Steel/steelBall-on-flat30.07/0.05[57]
c0.052/0.035
d0.035/0.059
N1888NTf2Diester1.25–5Steel/steelBall-on-disk400.069–0.071[58]
800.068–0.07
1200.068–0.07
aCitrateBO1Titanium/
ceramic
Ball-on-flat2∼0.12[59]
N6666OctanoatePOE0.5/2Steel/steelBall-on-plate300.062/0.059[60]
500.065/0.061
Palmitate300.066/0.062
500.066/0.063
P66614StearatePETO4Steel/steelBall-on-disk2000.105–0.12[61]
TMPTO0.12–0.16
OleatePETO0.105–0.15
TMPTO0.12–0.16
(C1)2S2PO2PETO0.075
TMPTO0.09
P66614(iC8)2PO2MJO1–10Steel/steelFour-ball3920.05–0.075[62]
N1888NTf2
P66614SiSOSQL1Steel/steelBall-on-disk6000.04–0.07[63]
P66614(iC8)2PO2BO1Steel/steelPin-on-disk4.90.125–0.22[64]
NTf20.12–0.28
N4444SulphatePEG
200
1.5Steel/steelFour-ball3920.12[65]
N88880.1
P44440.13
P88880.13

Table 4.

Tribological results of ILs as additives for polar oils (2017–2020).

2-ethylhexanoate; and the numbers in ammonium NX,X,X,X, and phosphonium PX,X,X,X represent the alkyl chain length.


Note: a- NH2((CH2)2OH)2; b- NH3((CH2)2OH); c- NH2CH3(CH2)2OH); d- NH(CH3)2(CH2)2OH).

CationAnionContent (wt.%)Tribo-pairContact modeLoad (N)COFRef
C6C1imIbu2Steel/steelBall-on-disk1000.135[29]
C8C1imIbu2Steel/steelBall-on-disk1000.127
N16111βP10.4Steel/steelFour-ball3920.075[66]
N16111βP20.4Steel/steelFour-ball3920.085
N4444BTA0.03εSteel/steelBall-on-plate1000.125[67]
P4444BTA0.03εSteel/steelBall-on-plate1000.125
NP1611χGAS0.5Steel/steelBall-on-disk1000.110[68]
NP1611χAK0.5Steel/steelBall-on-disk1000.095
NP1611ϕ1Br0.5Steel/steelBall-on-disk1000.075[69]
NP1611ϕ2Br0.5Steel/steelBall-on-disk1000.100
bStearate1Steel/sapphirePin-on-disk10.129[70]
a0.117
aPalmitate0.107
bφRicinoleate1Steel/steelBall-on-disk1250.14–0.15[71]
MBT0.18–0.2
aCitrate1Al/tungsten carbidePin-on-disk2.94∼0.45[9]
P66614NTf2∼0.57
P66614Decanoate∼0.18
aOleate1Steel/steelBall-on-plate20.078[72]
d0.066
aAlumina/steel20.079
d0.09
aSteel/steel40.091
d0.076
aAlumina/steel40.084
d0.094
cφRicinoleate1Steel-steelBall-on-disk125∼0.12[73]
Phosphate∼0.2

Table 5.

Tribological results of ILs as additives for water-based lubricants (2017–2020).

water-diethylene glycol.


water-sodium D-gluconate.


water-triethanolamine.


water-glycerol.


mol/L.


Note: The base lubricant of β is water-glycol.

a- NH2((CH2)2OH)2; b- NH3((CH2)2OH); c- NH2CH3(CH2)2OH); d- NH3(CH3)2CCH2C(CH3)3; and the numbers in imidazolium CxCxim; ammonium NX,X,X,X, and phosphonium PX,X,X,X represent the alkyl chain length.

4.1 Ionic liquids as oil additives

Due to the inherent polarity, the solubility of ILs in oils is a complicated issue. Most imidazolium-ILs are insoluble in the non-polar synthetic oils and mineral oils. So they are always used as lubricant additives in in very low concentrations or in oil-IL emulsions. In 2012, the fully oil-soluble phosphonium-based ILs [P6,6,6,14][DEHP] and [P6,6,6,14][BTMPP] were explored [74, 75]. These three-dimensional ILs have quaternary structures for both the cations and anions with long alkyl chains, giving rise to a high steric hindrance to screen the ions charge. Inspired by this, ILs having quaternary ammonium and phosphonium cations and halogen-free anions, such as phosphate, sulfonate, orthoborate, and carboxylate have been synthesized and tested as additives to the base oils [76]. Generally, larger cation sizes lead to higher solubilities of IL in nonpolar oils. In addition, ILs having symmetric cations would outperform the ones with asymmetric cations in wear reduction, and the symmetric-cation ILs are hypothesized to have a better mobility in the base oil to interact with metal surfaces and form protective boundary lubricating film [17]. In addition, some phosphonium-based ILs have been examined to show synergistic interactions with traditional additives ZDDP in hydrocarbon oils [77], or GTL base oil [78] to improve the wear resistance of oils.

In contrast to nonpolar oils, ILs have much better solubility in some polar oils, such as PEG200, in which [C6C1C1im][NTf2] can be dissolved up to 40 wt.%. Taher et al. [79] studied the lubricating properties of halogen-free ILs pyrrolidinium bis(mandelato)borate (hf-BILs) as additives to PEG200 in steel-steel contact. The addition of 3 wt.% of hf-BILs in the base oil reduced friction and wear significantly compared to PEG200 and 5 W40 engine oil. It is noted that shorten the length of the longest alkyl chain in this IL cation will improve the friction reduction and wear resistance of the IL-blends under same working conditions.

Recently, Guo et al. [57, 80] examined the tribological properties of three hydroxylammonium hexanoate PIL additives to a nonpolar mineral oil and a polar biodegradable oil. The impact of PILs ionicity and hydrogen bonding on the friction and wear performance was discussed. The results revealed that all PILs improved the lubricity and wear resistance of the biodegradable oil under steel-steel, particularly, the one with the lowest ionicity obtained the least material loss. While, when used as additives to the mineral oil, the three PILs behaved slightly different between steel-steel and steel-aluminum contact. The use of any PIL improved the mineral oil lubricity and wear resistance under both contacts, but PILs had quite different friction behaviors in steel-Al that the one with the highest ionicity presented the best friction.

4.2 Ionic liquids in water-based lubricant

Water or water-based lubricants can effectively reduce the temperature and clean the contaminants from surface contacts, which leads to a better working conditions and increase the machine lifetime. Since the high volatile characteristic and high freezing point of water-based lubricants, they are preferable in some specific industrial applications such as cutting and machining. Recent studies about IL additives in water are summarized in Table 5.

In the study of Wang et al. [81], N-(3-(diethoxyphosphoryl)propyl)-N,N-dimethyloctadecan-1-ammonium bromide (NP) was investigated as water additive in a steel-steel contact. A lower friction and wear rate, and excellent extreme-pressure and abrasion resistance were obtained compared to an oil-based lubricant. The superior tribological property was attributed to the physical adsorption of ILs on the steel surfaces and the formation of a protective film due to the tribo-chemical reactions between NP and sliding surfaces.

Bermudez’s team [82] reported that water containing 1 wt.% PIL (2-hydroxyethylammonium) succinate (MSu) could reduce the running-in period when lubricating the sapphire-stainless steel contact. It is also noted that a thin PIL boundary film was found on the steel surface once the base water evaporated, leading to an extremely low minimum friction coefficient of 0.0001. In addition, another PIL additive, di[bis(2-hydroxyethyl)ammonium] succinate (DSu) was also investigated under sapphire-stainless steel [40]. The results showed that although the use of 1 wt.% DSu + Water caused a higher running-in friction coefficient compared to that of neat DSu, PIL-mixture received a comparable anti-wear behavior with regards to the neat Dsu, and even got a slightly smaller wear rate of 1.83 × 10−5 mm3/m.

4.3 Ionic liquids and nanoscale additives

Nanomaterials, such as nanoparticles (NPs), graphene, and carbon nanotubes (CNTs), have been regarded as attractive solid lubricants which can be applied as lubricant additives and components for coatings to achieve good lubricity or super-lubricity. In [83], the magnesium silicate hydroxide–based nanoparticles have been studied and proved to be effective anti-wear additives, where the excellent tribological properties can be generally ascribed to the grinding, rolling, filling effects and the tribofilm formation.

However, the poor dispersion and low solubility of nanomaterials in the base lubricants limit their long-term practical applications. Therefore, the nanomaterial surface functionalization becomes necessary to their lubrication performance. The use of an oil-soluble PIL with long-alkyl-chain to incorporate the copper oxide nanoparticles as additives to a base oil PAO was firstly reported in [84]. In this study, the PIL was employed to improve the dispersion of the copper oxide NPs, where the hybrid PIL-NPs additives exhibited an enhanced oil-load capacity and a better anti-wear performance compared to that just using copper nanoparticles as additives. Recently, the friction behavior and wear performance of diamond and ZnO NPs stabilized by trihexyltetradecylphosphonium bis (2, 4, 4-trimethylpentyl) phosphinate were investigated in a steel-ceramic contact [85]. It was found that nanoparticles mixed with IL caused a higher friction coefficient with respect to only IL was used as additive to the gear base oil, where the nanoparticles were regarded as to wear out the film formed by the IL. While the use of diamond/ZnO nanoparticles with IL obtained a smaller wear volume of the ceramic ball compared to that of IL. Particularly, both the hybrid IL-nanoadditives showed effective anti-scuffing properties which revealed their potential to be extreme pressure additives.

Advertisement

5. Conclusions

As the aforementioned excellent physicochemical properties and friction and wear performance, ILs not only can be used as neat lubricants, friction-reducing additives, anti-wear additives, extreme pressure additives, but can also be used as corrosion inhibitors. Although IL corrosion inhibitors have been evaluated on many ferrous metals and alloys, their study on non-ferrous metals, such as aluminum is extremely limited, which is worthwhile to discuss. Meanwhile, the relationship between the outstanding corrosion inhibition and high performance of lubrication should be explored, when ILs are used as lubricants and lubricant additives.

Additionally, enormous literature has revealed that the adsorption of the ILs on the metallic surfaces and the tribo-chemical reactions between the active elements of ILs and the surfaces effectively improved the tribological performances of different contacts. However, the adsorption mechanism and tribofilm growth mechanism of ILs are still not clear, and the application of ILs in the tribology field, especially for PILs, should be further explored owing to its efficiency and green nature.

Advertisement

Acknowledgments

Hong Guo wants to express her gratitude to the Gleason Doctoral Fellowship from the Gleason Corporation.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Acronyms and abbreviations

[NTf2]

bis(trifluoromethylsulfonyl)amide

BScB

bis(salicylato)borates

C6:0

hexanoate

C8:0

octanoate

C10:0

decanoate

C12:0

laurate

C16:0

palmitate

C18:0

stearate

C18:1

oleate

(iC8)2PO2/ BTMPP

bis(2,4,4-trimethylpentyl)phosphinate

BEHP/DEHP

bis(2-ethylhexyl)phosphate

DBS

dodecylbenzenesulfonate

DEP

diethylphosphate

BF4

tetrafluoroborate

PF6

hexafluorophosphate

IL-TO

t-octhylammonium

DOSS

dioctyl sulfosuccinate

TTAOA

4(or 5)-methyl-benzotriazole-1-ylmethyl)-octadec-9-enyl-ammonium

TTADO

4(or 5)-methyl-benzotriazole-1-ylmethyl)-dioctyl-ammonium

(C1)2S2PO2 - O,O′

diethyldithiophosphate

SiSO – 3

(trimethylsily)propane-1-sulfonate

Ibu

ibuprofen

P1

phosphate

P2

phosphite

BTA

benzotriazole

[NP1611][GAS]

N-(3-(Diethoxyphosphoryl)propyl)-N,N-dimethyloctadecan-1-aminium-2,3,4,5,6-pentahydroxyhexanoate

[NP1611][AK]

N-(3-(Diethoxyphosphoryl)propyl)-N,N-dimethyloctadecan-1-aminium-6-methyl-4-oxo-4H-1,2,3-oxathiazin-3-ide-2,2-dioxide

Br

bromide

MBT

2-mercaptoben- zothiazole

POE

polyol ester

GTL4

gas-to-liquid 4 cSt

PETO

pentaerythritol oleate

TMPTO

trimethylolpropyl trioleate

MJO

modified Jatropha oil

SQL

squalane

PEG200

polyethylene glycol

References

  1. 1. Holmberg, K., Erdemir, A.: Influence of tribology on global energy consumption, costs and emissions. Friction. 5, 263–284 (2017). doi:10.1007/s40544-017-0183-5
  2. 2. Ye, C., Liu, W., Chen, Y., Yu, L.: Room-temperature ionic liquids: a novel versatile lubricant. Chem. Commun. (Camb). 2244–2245 (2001). doi:10.1039/B106935G
  3. 3. Sharma, G., Singh, D., Gardas, R.L.: Effect of Fluorinated Anion on the Physicochemical, Rheological and Solvatochromic Properties of Protic and Aprotic Ionic Liquids: Experimental and Computational Study. ChemistrySelect. 2, 11653–11658 (2017). doi:10.1002/slct.201701985
  4. 4. Shang, D., Zhang, X., Zeng, S., Jiang, K., Gao, H., Dong, H., Yang, Q., Zhang, S.: Protic ionic liquid [Bim][NTf2] with strong hydrogen bond donating ability for highly efficient ammonia absorption. Green Chem. 19, 937–945 (2017). doi:10.1039/c6gc03026b
  5. 5. González, R., Ramos, D., Blanco, D., Fernández-González, A., Viesca, J.L., Hadfield, M., Hernández Battez, A.: Tribological performance of tributylmethylammonium bis(trifluoromethylsulfonyl)amide as neat lubricant and as an additive in a polar oil. Friction. 7, 282–288 (2019). doi:10.1007/s40544-018-0231-9
  6. 6. Gusain, R., Bakshi, P.S., Panda, S., Sharma, O.P., Gardas, R., Khatri, O.P.: Physicochemical and tribophysical properties of trioctylalkylammonium bis(salicylato)borate (N888n-BScB) ionic liquids: Effect of alkyl chain length. Phys. Chem. Chem. Phys. 19, 6433–6442 (2017). doi:10.1039/c6cp05990b
  7. 7. Oulego, P., Faes, J., González, R., Viesca, J.L., Blanco, D., Battez, A.H.: Relationships between the physical properties and biodegradability and bacteria toxicity of fatty acid-based ionic liquids. J. Mol. Liq. 292, (2019). doi:10.1016/j.molliq.2019.111451
  8. 8. Battez, A.H., Rivera, N., Blanco, D., Oulego, P., Viesca, J.L., González, R.: Physicochemical, traction and tribofilm formation properties of three octanoate-, laurate- and palmitate-anion based ionic liquids. J. Mol. Liq. 284, 639–646 (2019). doi:10.1016/j.molliq.2019.04.050
  9. 9. Del Sol, I., Gámez, A.J., Rivero, A., Iglesias, P.: Tribological performance of ionic liquids as additives of water-based cutting fluids. Wear. 426–427, 845–852 (2019). doi:10.1016/j.wear.2019.01.109
  10. 10. Qu, J., Luo, H., Chi, M., Ma, C., Blau, P.J., Dai, S., Viola, M.B.: Comparison of an oil-miscible ionic liquid and ZDDP as a lubricant anti-wear additive. Tribol. Int. 71, 88–97 (2014). doi:10.1016/j.triboint.2013.11.010
  11. 11. Hernández Battez, A., Bartolomé, M., Blanco, D., Viesca, J.L., Fernández-González, A., González, R.: Phosphonium cation-based ionic liquids as neat lubricants: Physicochemical and tribological performance. Tribol. Int. 95, 118–131 (2016). doi:10.1016/j.triboint.2015.11.015
  12. 12. Guo, H., Iglesias, P.: Tribological behavior of ammonium-based protic ionic liquid as lubricant additive. Friction. 9, 169–178 (2021). doi:10.1007/s40544-020-0378-z
  13. 13. Guo, H., Iglesias, P.: Tribological Properties of Ammonium Protic Ionic Liquids As Additives in Polyalphaolefin for Steel-Steel Contact. In: Volume 12: Advanced Materials: Design, Processing, Characterization, and Applications. American Society of Mechanical Engineers (2019)
  14. 14. Vega, M.R.O., Parise, K., Ramos, L.B., Boff, U., Mattedi, S., Schaeffer, L., Malfatti, C.F.: Protic ionic liquids used as metal-forming green lubricants for aluminum: Effect of anion chain length. Mater. Res. 20, 675–687 (2017). doi:10.1590/1980-5373-MR-2016-0626
  15. 15. Guo, H., Smith, T.W., Iglesias, P.: The study of hexanoate-based protic ionic liquids used as lubricants in steel-steel contact. J. Mol. Liq. 299, 112208 (2019). doi:10.1016/j.molliq.2019.112208
  16. 16. Rocha, M.A.A., Neves, C.M.S.S., Freire, M.G., Russina, O., Triolo, A., Coutinho, J.A.P., Santos, L.M.N.B.F.: Alkylimidazolium based ionic liquids: Impact of cation symmetry on their nanoscale structural organization. J. Phys. Chem. B. 117, 10889–10897 (2013). doi:10.1021/jp406374a
  17. 17. Barnhill, W.C., Qu, J., Luo, H., Meyer, H.M., Ma, C., Chi, M., Papke, B.L.: Phosphonium-organophosphate ionic liquids as lubricant additives: Effects of cation structure on physicochemical and tribological characteristics. ACS Appl. Mater. Interfaces. 6, 22585–22593 (2014). doi:10.1021/am506702u
  18. 18. Barnhill, W.C., Luo, H., Meyer, H.M., Ma, C., Chi, M., Papke, B.L., Qu, J.: Tertiary and Quaternary Ammonium-Phosphate Ionic Liquids as Lubricant Additives. Tribol. Lett. 63, 1–11 (2016). doi:10.1007/s11249-016-0707-6
  19. 19. Xue, L., Gurung, E., Tamas, G., Koh, Y.P., Shadeck, M., Simon, S.L., Maroncelli, M., Quitevis, E.L.: Effect of Alkyl Chain Branching on Physicochemical Properties of Imidazolium-Based Ionic Liquids. J. Chem. Eng. Data. 61, 1078–1091 (2016). doi:10.1021/acs.jced.5b00658
  20. 20. Xia, Y., Wang, Z., Song, Y.: Influence of hydroxyl group functionalization and alkyl chain length on physicochemical and antiwear properties of hexafluorophosphate imidazolium ionic liquids. Ind. Lubr. Tribol. 66, 443–451 (2014). doi:10.1108/ilt-11-2011-0089
  21. 21. De La Hoz, A.T., Brauer, U.G., Miller, K.M.: Physicochemical and thermal properties for a series of 1-alkyl-4-methyl-1, 2,4-triazolium bis(trifluoromethylsulfonyl)imide ionic liquids. J. Phys. Chem. B. 118, 9944–9951 (2014). doi:10.1021/jp505592t
  22. 22. Er, H., Xu, Y., Zhao, H.: Properties of mono-protic ionic liquids composed of hexylammonium and hexylethylenediaminium cations with trifluoroacetate and bis (trifluoromethylsulfonyl) imide anions. J. Mol. Liq. 276, 379–384 (2019). doi:10.1016/j.molliq.2018.11.132
  23. 23. Fadeeva, Y.A., Gruzdev, M.S., Kudryakova, N.O., Shmukler, L.E., Safonova, L.P.: Physico-chemical characterization of alkyl-imidazolium protic ionic liquids. J. Mol. Liq. 297, 111305 (2020). doi:10.1016/j.molliq.2019.111305
  24. 24. Xiang, J., Chen, R., Wu, F., Li, L., Chen, S., Zou, Q.: Physicochemical properties of new amide-based protic ionic liquids and their use as materials for anhydrous proton conductors. Electrochim. Acta. 56, 7503–7509 (2011). doi:10.1016/j.electacta.2011.06.103
  25. 25. Xue, Z., Qin, L., Jiang, J., Mu, T., Gao, G.: Thermal, electrochemical and radiolytic stabilities of ionic liquids. Phys. Chem. Chem. Phys. 20, 8382–8402 (2018). doi:10.1039/c7cp07483b
  26. 26. Ma, R., Zhao, Q., Zhang, E., Zheng, D., Li, W., Wang, X.: Synthesis and evaluation of oil-soluble ionic liquids as multifunctional lubricant additives. Tribol. Int. 151, 106446 (2020). doi:10.1016/j.triboint.2020.106446
  27. 27. Kreivaitis, R., Gumbytė, M., Kupčinskas, A., Kazancev, K., Makarevičienė, V.: Investigating the tribological properties of PILs derived from different ammonium cations and long chain carboxylic acid anion. Tribol. Int. 141, 105905 (2020). doi:10.1016/j.triboint.2019.105905
  28. 28. Dong, R., Yu, Q., Bai, Y., Wu, Y., Ma, Z., Zhang, J., Zhang, C., Yu, B., Zhou, F., Liu, W., Cai, M.: Towards superior lubricity and anticorrosion performances of proton-type ionic liquids additives for water-based lubricating fluids. Chem. Eng. J. 383, (2020). doi:10.1016/j.cej.2019.123201
  29. 29. Wang, Y., Yu, Q., Cai, M., Shi, L., Zhou, F., Liu, W.: Ibuprofen-Based Ionic Liquids as Additives for Enhancing the Lubricity and Antiwear of Water–Ethylene Glycol Liquid. Tribol. Lett. 65, 1–13 (2017). doi:10.1007/s11249-017-0840-x
  30. 30. Huang, G., Fan, S., Ba, Z., Cai, M., Qiao, D.: Insight into the lubricating mechanism for alkylimidazolium phosphate ionic liquids with different alkyl chain length. Tribol. Int. 140, 105886 (2019). doi:10.1016/j.triboint.2019.105886
  31. 31. Su, T., Song, G., Zheng, D., Ju, C., Zhao, Q.: Facile synthesis of protic ionic liquids hybrid for improving antiwear and anticorrosion properties of water-glycol. Tribol. Int. 153, 106660 (2021). doi:10.1016/j.triboint.2020.106660
  32. 32. Ortega Vega, M.R., Mattedi, S., Schroeder, R.M., de Fraga Malfatti, C.: 2-Hydroxyethilammonium Oleate Protic Ionic Liquid As Corrosion Inhibitor for Aluminum in Neutral Medium. Mater. Corros. (2020). doi:10.1002/maco.202011847
  33. 33. Viesca, J.L., Oulego, P., González, R., Guo, H., Battez, A.H., Iglesias, P.: Miscibility, corrosion and environmental properties of six hexanoate- and sulfonate-based protic ionic liquids. J. Mol. Liq. 114561 (2020). doi:10.1016/j.molliq.2020.114561
  34. 34. Ghanem, O. Ben, Mutalib, M.I.A., Lévêque, J.-M., El-Harbawi, M.: Development of QSAR model to predict the ecotoxicity of Vibrio fischeri using COSMO-RS descriptors. Chemosphere. 170, 242–250 (2017). doi:10.1016/j.chemosphere.2016.12.003
  35. 35. Hodyna, D., Kovalishyn, V., Semenyuta, I., Blagodatnyi, V., Rogalsky, S., Metelytsia, L.: Imidazolium ionic liquids as effective antiseptics and disinfectants against drug resistant S. aureus: In silico and in vitro studies. Comput. Biol. Chem. 73, 127–138 (2018). doi:10.1016/j.compbiolchem.2018.01.012
  36. 36. Cao, L., Zhu, P., Zhao, Y., Zhao, J.: Using machine learning and quantum chemistry descriptors to predict the toxicity of ionic liquids. J. Hazard. Mater. 352, 17–26 (2018). doi:10.1016/j.jhazmat.2018.03.025
  37. 37. Abramenko, N., Kustov, L., Metelytsia, L., Kovalishyn, V., Tetko, I., Peijnenburg, W.: A review of recent advances towards the development of QSAR models for toxicity assessment of ionic liquids. J. Hazard. Mater. 384, 121429 (2020). doi:10.1016/j.jhazmat.2019.121429
  38. 38. Mena, I.F., Diaz, E., Palomar, J., Rodriguez, J.J., Mohedano, A.F.: Cation and anion effect on the biodegradability and toxicity of imidazolium– and choline–based ionic liquids. Chemosphere. 240, 124947 (2020). doi:10.1016/j.chemosphere.2019.124947
  39. 39. Tzani, A., Elmaloglou, M., Kyriazis, C., Aravopoulou, D., Kleidas, I., Papadopoulos, A., Ioannou, E., Kyritsis, A., Voutsas, E., Detsi, A.: Synthesis and structure-properties relationship studies of biodegradable hydroxylammonium-based protic ionic liquids. J. Mol. Liq. 224, 366–376 (2016). doi:10.1016/j.molliq.2016.09.086
  40. 40. Avilés, M.D., Carrión-Vilches, F.J., Sanes, J., Bermúdez, M.D.: Diprotic Ammonium Succinate Ionic Liquid in Thin Film Aqueous Lubrication and in Graphene Nanolubricant. Tribol. Lett. 67, 1–10 (2019). doi:10.1007/s11249-019-1138-y
  41. 41. Khan, A., Gusain, R., Sahai, M., Khatri, O.P.: Fatty acids-derived protic ionic liquids as lubricant additive to synthetic lube base oil for enhancement of tribological properties. J. Mol. Liq. 293, 111444 (2019). doi:10.1016/j.molliq.2019.111444
  42. 42. Ortega Vega, M.R., Ercolani, J., Mattedi, S., Aguzzoli, C., Ferreira, C.A., Rocha, A.S., Malfatti, C.F.: Oleate-Based Protic Ionic Liquids As Lubricants for Aluminum 1100. Ind. Eng. Chem. Res. 57, 12386–12396 (2018). doi:10.1021/acs.iecr.8b02426
  43. 43. Yao, Y., Xu, Y., Fan, X., Zhu, M., Liu, G.: Tribological properties of spherical and mesoporous NiAl particles as ionic liquid additives. Friction. 8, 384–395 (2020). doi:10.1007/s40544-019-0266-6
  44. 44. Viesca, J.L., Battez, A.H., González, R., Reddyhoff, T., Pérez, A.T., Spikes, H.A.: Assessing boundary film formation of lubricant additivised with 1-hexyl-3-methylimidazolium tetrafluoroborate using ECR as qualitative indicator. Wear. 269, 112–117 (2010). doi:10.1016/j.wear.2010.03.014
  45. 45. Zhou, Y., Leonard, D.N., Guo, W., Qu, J.: Understanding Tribofilm Formation Mechanisms in Ionic Liquid Lubrication. Sci. Rep. 7, 1–8 (2017). doi:10.1038/s41598-017-09029-z
  46. 46. Guo, W., Zhou, Y., Sang, X., Leonard, D.N., Qu, J., Poplawsky, J.D.: Atom Probe Tomography Unveils Formation Mechanisms of Wear-Protective Tribofilms by ZDDP, Ionic Liquid, and Their Combination. ACS Appl. Mater. Interfaces. 9, 23152–23163 (2017). doi:10.1021/acsami.7b04719
  47. 47. Yu, Q., Wang, Y., Huang, G., Ma, Z., Shi, Y., Cai, M., Zhou, F., Liu, W.: Task-Specific Oil-Miscible Ionic Liquids Lubricate Steel/Light Metal Alloy: A Tribochemistry Study. Adv. Mater. Interfaces. 5, 1–12 (2018). doi:10.1002/admi.201800791
  48. 48. Kawada, S., Watanabe, S., Kondo, Y., Tsuboi, R., Sasaki, S.: Tribochemical Reactions of Ionic Liquids Under Vacuum Conditions. Tribol. Lett. 54, 309–315 (2014). doi:10.1007/s11249-014-0342-z
  49. 49. Kawada, S., Watanabe, S., Tadokoro, C., Tsuboi, R., Sasaki, S.: Lubricating mechanism of cyano-based ionic liquids on nascent steel surface. Tribol. Int. 119, 474–480 (2018). doi:10.1016/j.triboint.2017.11.019
  50. 50. Sharma, V., Doerr, N., Aswath, P.B.: Chemical–mechanical properties of tribofilms and their relationship to ionic liquid chemistry. RSC Adv. 6, 22341–22356 (2016). doi:10.1039/C6RA01915C
  51. 51. Landauer, A.K., Barnhill, W.C., Qu, J.: Correlating mechanical properties and anti-wear performance of tribofilms formed by ionic liquids, ZDDP and their combinations. Wear. 354–355, 78–82 (2016). doi:10.1016/j.wear.2016.03.003
  52. 52. Huang, G., Yu, Q., Ma, Z., Cai, M., liu, W.: Probing the lubricating mechanism of oil-soluble ionic liquids additives. Tribol. Int. 107, 152–162 (2017). doi:10.1016/j.triboint.2016.08.027
  53. 53. Li, W., Kumara, C., Meyer, H.M., Luo, H., Qu, J.: Compatibility between Various Ionic Liquids and an Organic Friction Modifier as Lubricant Additives. Langmuir. 34, 10711–10720 (2018). doi:10.1021/acs.langmuir.8b02482
  54. 54. Guo, H., Liu, R., Fuentes-Aznar, A., Iglesias Victoria, P.: Friction and Wear Properties of Halogen-Free and Halogen-Containing Ionic Liquids Used As Neat Lubricants, Lubricant Additives and Thin Lubricant Layers. In: Volume 10: 2017 ASME International Power Transmission and Gearing Conference. pp. 1–5. American Society of Mechanical Engineers (2017)
  55. 55. Duan, H., Li, W., Kumara, C., Jin, Y., Meyer, H.M., Luo, H., Qu, J.: Ionic liquids as oil additives for lubricating oxygen-diffusion case-hardened titanium. Tribol. Int. 136, 342–348 (2019). doi:10.1016/j.triboint.2019.03.069
  56. 56. González, R., Viesca, J.L., Battez, A.H., Hadfield, M., Fernández-González, A., Bartolomé, M.: Two phosphonium cation-based ionic liquids as lubricant additive to a polyalphaolefin base oil. J. Mol. Liq. 293, 26–30 (2019). doi:10.1016/j.molliq.2019.111536
  57. 57. Guo, H., Adukure, A.R., Iglesias, P.: Effect of Ionicity of Three Protic Ionic Liquids as Neat Lubricants and Lubricant Additives to a Biolubricant. Coatings. 9, 713 (2019). doi:10.3390/coatings9110713
  58. 58. Viesca, J.L., Mallada, M.T., Blanco, D., Fernández-González, A., Espina-Casado, J., González, R., Hernández Battez, A.: Lubrication performance of an ammonium cation-based ionic liquid used as an additive in a polar oil. Tribol. Int. 116, 422–430 (2017). doi:10.1016/j.triboint.2017.08.004
  59. 59. Magar, S.A., Guo, H., Iglesias, P.: Ionic liquid as cutting fluid additive using minimum quantity lubricant (MQL) in titanium-ceramic contact. ASME Int. Mech. Eng. Congr. Expo. Proc. 12, 1–7 (2019). doi:10.1115/IMECE2019-10647
  60. 60. Sernaglia, M., Blanco, D., Hernández Battez, A., González, R., Fernández-González, A., Bartolomé, M.: Two fatty acid anion-based ionic liquids - part II: Effectiveness as an additive to a polyol ester. J. Mol. Liq. 310, (2020). doi:10.1016/j.molliq.2020.113158
  61. 61. Zhu, L., Zhao, G., Wang, X.: Investigation on three oil-miscible ionic liquids as antiwear additives for polyol esters at elevated temperature. Tribol. Int. 109, 336–345 (2017). doi:10.1016/j.triboint.2016.10.032
  62. 62. Amiril, S.A.S., Rahim, E.A., Embong, Z., Syahrullail, S.: Tribological investigations on the application of oil-miscible ionic liquids additives in modified Jatropha-based metalworking fluid. Tribol. Int. 120, 520–534 (2018). doi:10.1016/j.triboint.2018.01.030
  63. 63. Hansen, J., Björling, M., Minami, I., Larsson, R.: Performance and mechanisms of silicate tribofilm in heavily loaded rolling/sliding non-conformal contacts. Tribol. Int. 123, 130–141 (2018). doi:10.1016/j.triboint.2018.03.006
  64. 64. Cigno, E., Magagnoli, C., Pierce, M.S., Iglesias, P.: Lubricating ability of two phosphonium-based ionic liquids as additives of a bio-oil for use in wind turbines gearboxes. Wear. 376–377, 756–765 (2017). doi:10.1016/j.wear.2017.01.010
  65. 65. Khan, A., Yasa, S.R., Gusain, R., Khatri, O.P.: Oil-miscible, halogen-free, and surface-active lauryl sulphate-derived ionic liquids for enhancement of tribological properties. J. Mol. Liq. 318, 114005 (2020). doi:10.1016/j.molliq.2020.114005
  66. 66. Zheng, G., Zhang, G., Ding, T., Xiang, X., Li, F., Ren, T., Liu, S., Zheng, L.: Tribological properties and surface interaction of novel water-soluble ionic liquid in water-glycol. Tribol. Int. 116, 440–448 (2017). doi:10.1016/j.triboint.2017.08.001
  67. 67. Fan, M., Du, X., Ma, L., Wen, P., Zhang, S., Dong, R., Sun, W., Yang, D., Zhou, F., Liu, W.: In situ preparation of multifunctional additives in water. Tribol. Int. 130, 317–323 (2019). doi:10.1016/j.triboint.2018.09.020
  68. 68. Wang, Y., Yu, Q., Cai, M., Zhou, F., Liu, W.: Halide-free PN ionic liquids surfactants as additives for enhancing tribological performance of water-based liquid. Tribol. Int. 128, 190–196 (2018). doi:10.1016/j.triboint.2018.07.018
  69. 69. Wang, Y., Yu, Q., Ma, Z., Huang, G., Cai, M., Zhou, F., Liu, W.: Significant enhancement of anti-friction capability of cationic surfactant by phosphonate functionality as additive in water. Tribol. Int. 112, 86–93 (2017). doi:10.1016/j.triboint.2017.03.034
  70. 70. Avilés, M.D., Carrión, F.J., Sanes, J., Bermúdez, M.D.: Effects of protic ionic liquid crystal additives on the water-lubricated sliding wear and friction of sapphire against stainless steel. Wear. 408–409, 56–64 (2018). doi:10.1016/j.wear.2018.04.015
  71. 71. Zheng, D., Wang, X., Zhang, M., Ju, C.: Synergistic Effects Between the Two Choline-Based Ionic Liquids as Lubricant Additives in Glycerol Aqueous Solution. Tribol. Lett. 67, 1–13 (2019). doi:10.1007/s11249-019-1161-z
  72. 72. Kreivaitis, R., Gumbytė, M., Kupčinskas, A., Kazancev, K., Ta, T.N., Horng, J.H.: Investigation of tribological properties of two protic ionic liquids as additives in water for steel–steel and alumina–steel contacts. Wear. 456–457, (2020). doi:10.1016/j.wear.2020.203390
  73. 73. Zheng, D., Wang, X., Liu, Z., Ju, C., Xu, Z., Xu, J., Yang, C.: Synergy between two protic ionic liquids for improving the antiwear property of glycerol aqueous solution. Tribol. Int. 141, 105731 (2020). doi:10.1016/j.triboint.2019.04.015
  74. 74. Qu, J., Bansal, D.G., Yu, B., Howe, J.Y., Luo, H., Dai, S., Li, H., Blau, P.J., Bunting, B.G., Mordukhovich, G., Smolenski, D.J.: Antiwear performance and mechanism of an oil-miscible ionic liquid as a lubricant additive. ACS Appl. Mater. Interfaces. 4, 997–1002 (2012). doi:10.1021/am201646k
  75. 75. Yu, B., Bansal, D.G., Qu, J., Sun, X., Luo, H., Dai, S., Blau, P.J., Bunting, B.G., Mordukhovich, G., Smolenski, D.J.: Oil-miscible and non-corrosive phosphonium-based ionic liquids as candidate lubricant additives. Wear. 289, 58–64 (2012). doi:10.1016/j.wear.2012.04.015
  76. 76. Cowie, S., Cooper, P.K., Atkin, R., Li, H.: Nanotribology of Ionic Liquids as Lubricant Additives for Alumina Surfaces. J. Phys. Chem. C. 121, 28348–28353 (2017). doi:10.1021/acs.jpcc.7b09879
  77. 77. Zhang, Y., Cai, T., Shang, W., Sun, L., Liu, D., Tong, D., Liu, S.: Environmental friendly polyisobutylene-based ionic liquid containing chelated orthoborate as lubricant additive: Synthesis, tribological properties and synergistic interactions with ZDDP in hydrocarbon oils. Tribol. Int. 115, 297–306 (2017). doi:10.1016/j.triboint.2017.05.038
  78. 78. Qu, J., Barnhill, W.C., Luo, H., Meyer, H.M., Leonard, D.N., Landauer, A.K., Kheireddin, B., Gao, H., Papke, B.L., Dai, S.: Synergistic Effects between Phosphonium-Alkylphosphate Ionic Liquids and Zinc Dialkyldithiophosphate (ZDDP) as Lubricant Additives. Adv. Mater. 27, 4767–4774 (2015). doi:10.1002/adma.201502037
  79. 79. Taher, M., Shah, F.U., Filippov, A., De Baets, P., Glavatskih, S., Antzutkin, O.N.: Halogen-free pyrrolidinium bis(mandelato)borate ionic liquids: Some physicochemical properties and lubrication performance as additives to polyethylene glycol. RSC Adv. 4, 30617–30623 (2014). doi:10.1039/c4ra02551b
  80. 80. Guo, H., Pang, J., Adukure, A.R., Iglesias, P.: Influence of Hydrogen Bonding and Ionicity of Protic Ionic Liquids on Lubricating Steel–Steel and Steel–Aluminum Contacts: Potential Ecofriendly Lubricants and Additives. Tribol. Lett. 68, 1–10 (2020). doi:10.1007/s11249-020-01354-1
  81. 81. Wang, Y., Yu, Q., Ma, Z., Huang, G., Cai, M.: Tribology International Signi fi cant enhancement of anti-friction capability of cationic surfactant by phosphonate functionality as additive in water. Tribiology Int. 112, 86–93 (2017). doi:10.1016/j.triboint.2017.03.034
  82. 82. Espinosa, T., Jimenez, M., Sanes, J., Jimenez, A.E., Iglesias, M., Bermudez, M.D.: Ultra-low friction with a protic ionic liquid boundary film at the water-lubricated sapphire-stainless steel interface. Tribol. Lett. 53, 1–9 (2014). doi:10.1007/s11249-013-0238-3
  83. 83. Guo, H., Chen, F., Liu, R., Iglesias, P.: Lubricating Ability of Magnesium Silicate Hydroxide–Based Nanopowder as Lubricant Additive in Steel–Steel and Ceramic–Steel Contacts. Tribol. Trans. 63, 585–596 (2020). doi:10.1080/10402004.2019.1710312
  84. 84. Taaber, T., Enok, A.E., Joost, U., Oras, S., Järvekülg, M., Lõhmus, R., Mäeorg, U., Saal, K.: Tribological properties of protic ionic liquid and functionalized copper oxide nanoparticles as additives to base oil. Mechanika. 21, 148–153 (2015). doi:10.5755/j01.mech.21.2.11723
  85. 85. Han, X., Zhang, Z., Thrush, S.J., Barber, G.C., Qu, H.: Ionic liquid stabilized nanoparticle additive in a steel-ceramic contact for extreme pressure application. Wear. 452–453, 203264 (2020). doi:10.1016/j.wear.2020.203264

Written By

Hong Guo and Patricia Iglesias Victoria

Submitted: 20 December 2020 Reviewed: 05 February 2021 Published: 16 March 2021