You are on page 1of 9

Journal of Molecular Liquids 339 (2021) 116829

Contents lists available at ScienceDirect

Journal of Molecular Liquids


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

Aminoguanidine-based deep eutectic solvents as


environmentally-friendly and high-performance lubricant additives
Amzad Khan a,b, Raghuvir Singh a, Piyush Gupta a, Kanika Gupta a,b, Om P. Khatri a,b,⇑
a
Chemical and Material Sciences Division, CSIR-Indian Institute of Petroleum, Dehradun 248005, India
b
Academcy of Scientific and Innovative Research, Ghaziabad 201002, India

a r t i c l e i n f o a b s t r a c t

Article history: Halogen-free, environmentally-friendly, and economically viable lubricant additives, which can furnish
Received 9 May 2021 excellent miscibility with mineral lube base oils and enhance the tribological properties, are gaining
Revised 22 June 2021 increasing interest for the formulation of new generation lubricants. Herein, aminoguanidine salt-
Accepted 24 June 2021
based deep eutectic solvents (DESs) having variable mole ratios of fatty acid (as a hydrogen bond donor;
Available online 29 June 2021
HBD) are synthesized through a single-step facile approach. The detailed chemical characterizations
based on FTIR and NMR (1H and 13C) confirmed the synthesis of DESs via a hydrogen-bonding network.
Keywords:
A linear correlation between the shear stress and shear rate implied Newtonian fluidic behaviour of these
Deep eutectic solvents
Hydrogen bonding
DESs. An increasing mole ratio of octanoic acid decreased the viscosity and thermal stability of resultant
Lubricant additives DESs. The long alkyl chain of HBD furnished good compatibility of these DESs with SN 150 mineral lube
Viscosity base oil. The DESs as additives to SN 150 mineral lube oil improved the viscosity index by 12% and
Fatty acid enhanced the tribological properties for steel tribopair by decreasing the wear volume (94%) and coeffi-
Friction cient of friction (30%). The detailed analyses of worn scars revealed the formation of DESs-derived tribo-
chemical thin film. The low shear strength of DESs-derived lubricious and protective thin film between
steel tribopair facilitated the sliding and avoided the direct contact between the steel tribopair, conse-
quently, significant enhancement of tribological properties.
Ó 2021 Elsevier B.V. All rights reserved.

1. Introduction systems. The use of low-viscosity lube base oil and variable
functional additives, which can furnish adequate lubrication to a
Lubricants are indispensable substances for engineering sur- diversified range of engineering surfaces and minimize the envi-
faces to minimize friction and wear, dissipate the heat from ronmental footprints, are gaining immense interest. In this context,
tribo-surfaces, keep the moving parts apart, transmit the forces, nanostructured materials, ionic liquids, organometallic, and
disperse the foreign particles, and improve the efficiency of engi- organic compounds have been explored as (a) novel additives to
nes and machines where two or more bodies are in relative motion. liquids lubricants and greases, (b) fillers to polymeric and metal-
The automotive/motor oils dominate the liquid lubricants market, matrix composites, and thin films/coatings to tribo surfaces for
followed by their applications in turbines, hydraulic systems, com- minimizing the friction and wear [3–8].
pressors, bearings, gears, metal-working, and so on [1]. The mas- Ionic liquids, the poorly coordinated salts, exhibit a diffused
sive wastage of energy and rising level of CO2 emission can be electrostatic interaction between the bulkier organic cation and
reduced by redefining the roles of lubricants and tribology in inorganic/organic anion. Ionic liquids are considered innovative
industrial machinery, devices, and automotive engines. The new fluids because of their unique and tunable physicochemical prop-
developments in tribological practices and lubricant formulations erties. The negligible volatility, inherent polarity, high thermal sta-
can save ~11% of the energy used by the transportation, industrial, bility, wide liquid range, good conductivity, confined layering
and utility sectors [2]. Over the recent past, several efforts have structure, and excellent affinity towards engineering surfaces
been made to develop new generation lubricants and additives to make them promising candidates for enhancing the tribological
increase the fuel/energy efficiency and durability of engineering properties [5,9–12]. However, the high cost of precursors, multiple
reaction steps for their preparation, purification processes, and cor-
rosiveness of halogenated ionic liquids have been major bottle-
⇑ Corresponding author at: Chemical and Material Sciences Division, CSIR-Indian
necks and hinder their applicability for several industrial
Institute of Petroleum, Dehradun 248005, India.
E-mail address: opkhatri@iip.res.in (O.P. Khatri). applications, including lubrication [13–15].

https://doi.org/10.1016/j.molliq.2021.116829
0167-7322/Ó 2021 Elsevier B.V. All rights reserved.
A. Khan, R. Singh, P. Gupta et al. Journal of Molecular Liquids 339 (2021) 116829

Deep eutectic solvents (DESs), also known as green analogous to The present work addresses the aminoguanidine salt-octanoic
ionic liquids, exhibit several similar physicochemical properties to acid (Ag-C8)-based DESs as an additive to SN 150 lube oil for the
ionic liquids. The DESs are prepared by synergistic blending of the enhancement of tribological properties of steel tribopair. The Ag-
hydrogen bond donors (HBD; amines, alcohols, fatty acids, etc.) C8 DESs showed good miscibility and improved the viscosity index
and the hydrogen bond acceptors (HBA; ammonium, phospho- of SN 150 mineral lube base oil. The surface analyses of worn steel
nium, choline, alkyl phosphate, etc.). The charge delocalization balls are carried out to understand the role of Ag-C8 DESs in the
via hydrogen bonding between the HBD and HBA reduces the melt- formation of tribochemical thin film and reduction of friction and
ing point of resultant DESs. The DESs have shown immense interest wear. The ease of preparation, low cost of precursors, and environ-
because of their low volatility, moderate conductivity, cost- mental compatibility make DESs greener alternatives to ionic liq-
effectiveness, and solvation. DESs are acknowledged over the ionic uids for the development of new generation lubricants.
liquids for several applications because of their environmental and
economic viability, low cytotoxicity and phytotoxicity, good 2. Experimental section
biodegradability, and ease of preparation [16–18]. The DESs are
still in their infancy stage, particularly for tribological applications. 2.1. Chemicals and materials
The polarity of DESs enhances their affinity towards the engineer-
ing surfaces and makes them potential candidates as alternatives Aminoguanidine bicarbonate (98%, Fluka) and octanoic acid
to ionic liquids for the improvement of tribological properties. (caprylic acid, 99%, Loba Chemie) were used without further purifi-
Abbott et al. demonstrated the choline-based DESs for lubrica- cation for the preparation of Ag-C8 DESs.
tion of steel tribopair. The choline chloride-urea and choline
chloride-ethylene glycol DESs formed a thin film, which enhanced 2.2. Synthesis of DESs
the lubrication of tribo interfaces by reducing friction and wear.
The quality of the lubricious thin film is governed by the composi- Ag-C8 DESs having variable molar ratios of aminoguanidine
tion of DESs [19]. The choline chloride-based DESs having urea, bicarbonate salt as an HBA and caprylic acid as an HBD (1:2, 1:4,
maleic acid, and ethylene glycol as HBD showed a significantly and 1:8) were synthesized by their blending at 60 °C under the
lower friction coefficient for steel tribopair than PAO6 lube oil uninterrupted stirring as shown in Fig. 1. The synthesized homoge-
[20]. The water-miscible choline chloride-based DESs with urea, nous liquid phase of Ag-C8 DES was kept in the vacuum oven for
ethylene glycol, oxalic acid, and glycerol displayed low corrosion 24 h. These DESs having mole ratios of 1:2, 1:4, and 1:8 of
rates for steel, aluminium, and nickel. Moreover, these DESs fur- aminoguanidine bicarbonate salt and caprylic acid are coined as
nished a low coefficient of friction for iron-based materials com- AG-C8-(1:2), AG-C8-(1:4), and AG-C8-(1:8) throughout the manu-
pared to mineral lube base oil [21]. The DESs of sulfur-containing script. The syntheses of these DESs were confirmed by NMR and
ionic liquids and polyethylene glycol formed the tribo thin film, FTIR spectroscopic measurements. The chemical structure of AG-
which extended the lubrication of silicon substrate by decreasing C8 DES and the plausible hydrogen linkages are shown in Fig. 1.
the friction and protecting the tribo-surfaces against undesirable
wear [22]. The ultrathin film of choline chloride-ethylene glycol 2.3. Chemical and physicochemical characterizations of DESs
DES between two mica surfaces furnished the structured layers
of quantized frictional properties, which are highly sensitive to Vibrational spectra of all samples were collected using an FTIR
water content. The quantized layer of DES offered the superlubric spectroscope (Spectrum-Two, Perkin Elmer) in a transmittance
(l < 0.01) behaviour under the load of 0 to 50 lN [23]. These find- mode. The NMR (1H and 13C) spectra of DESs were measured using
ings suggest the potential of DESs for lubrication enhancement. the Bruker Avance III 500 MHz spectroscope. A blend of 20% DES
However, not much work has been carried out, particularly with sample in the deuterated solvent, i.e., CDCl3, was used for collect-
halogen-free DESs for tribological applications. ing the NMR spectra. The thermal stability of these DESs was

Fig. 1. Synthesis of Ag-C8 DESs using aminoguanidine bicarbonate salt and octanoic acid as HBA and HBD, respectively.

2
A. Khan, R. Singh, P. Gupta et al. Journal of Molecular Liquids 339 (2021) 116829

measured by a thermogravimetric analyzer (TGA; Model: Dia- 3. Results and discussion


mond, PerkinElmer) under the N2 flow in a temperature range of
30–300 °C. The kinematic viscosity of DESs and their blend in Fig. 2 displays vibrational spectra of aminoguanidine bicarbon-
SN-150 mineral lube base oil was measured using the Stabinger ate salt and octanoic acid. The FTIR spectrum (Fig. 2a) of
viscometer (SVM 3000; Anton Paar). The viscosity index of 1.5% aminoguanidine bicarbonate exhibits multiple peaks in the spec-
DESs in SN-150 lube oil was estimated based on viscosity at 40 tral range of 3410–3150 cm 1, revealing the N-H stretches due to
and 100 °C as per the ASTM D2270 standard method. The flow primary and secondary amine functionalities in the aminoguani-
characteristics of DESs were measured using a modular compact dine. The N-H stretches due to protonated amine group in
rheometer (Anton Paar GmbH MCR 102) equipped with CP40 aminoguanidine salt are vibrated at lower wavenumbers (3050–
cone-plate geometry. 2900 cm 1) [24]. The C@O stretch (1690 cm 1) along with asym-
metric (1499 cm 1) and symmetric (1353 cm 1) stretches of car-
2.4. Tribo-evaluation of DESs bonate group revealed the presence of bicarbonate in
aminoguanidine salt. A bending mode at 1633 cm 1 along with a
The AG-C8 DESs were used as additives to SN 150 mineral base shoulder toward lower wavenumber (1611 cm 1) signified the
oil and evaluated the tribo-performance for steel tribopair. All tri- NAH based anime functionalities in aminoguanidine bicarbonate
bological tests were carried out using the four-ball tribo-test salt [25,26]. A strong and broad vibrational peak spanning from
machine (M/s. Ducom India). The steel balls (/ = 12.7 mm) were 3400 to 2800 cm 1 due to OAH stretch (Fig. 2b) signified the inter-
thoroughly cleaned with the aid of ultrasound using n-hexane sol- molecular hydrogen-bonded carboxylic group in octanoic acid [27].
vent. Three steel balls were fixed in a sample pot. The lubricant Moreover, it is overlapped with CAH stretching modes of methy-
sample was filled in a sample pot, and a fourth steel ball was lene and methyl units of octyl chains. An intense vibrational mode
rotated over the three stationary balls under the desired load due to C@O stretch at 1712 cm 1 further confirmed the carboxylic
(392 N), temperature (75 °C), and speed (1200 rpm). Each tribolog- group in the octanoic acid. As demonstrated in the molecular struc-
ical test was carried out for one hour as per the ASTM D4172B stan- ture of AG-C8 DESs (Fig. 1), the AOAH and AC@O linkages of a car-
dard test method. The coefficient of friction data was collected boxylic group formed the hydrogen bonding network with NAH
with a function of time. After completing the tribo-test, the three linkages of aminoguanidine; consequently, the vibrational features
stationary balls were removed from the sample holder and cleaned due to OAH and NAH became very intense and broad in a range of
with n-hexane. The worn scar developed on the stationary balls 3400–2800 cm 1 (Fig. 2d-f). The overlapping of intense and broad
was measured by an optical microscope and reported values are vibrational peaks of NAH and OAH stretches due to the hydrogen
based on the wear scar diameter (WSD) of six balls (from two inde- bonding network in AG-C8 DESs makes it very difficult to reveal
pendent tests of each sample). The microscopic features of worn their vibrational shift precisely. Interestingly the C@O stretch
scars of steel balls were probed using a field-emission scanning (1712 cm 1) in octanoic acid red-shifted to 1700–1706 cm 1 in
electron microscope (FESEM, FEI Quanta 200F). An energy- the AG-C8 DESs having variable molar ratios of octanoic acid. Like-
dispersive X-ray spectroscope (EDS) integrated with the FESEM wise, the N-H bending mode in aminoguanidine salt is red shifted
was used for analyzing the elemental distribution over the worn to 1535–1540 cm 1 in the AG-C8 DESs because of their participa-
scars. tion in hydrogen bonding. The OAH bending mode of the

Fig. 2. FTIR spectra of (a) aminoguanidine bicarbonate, (b) octanoic acid, (c) AG-C8-(1:2), (d) AG-C8-(1:8), (e) AG-C8-(1:4), and AG-C8-(1:2) DESs.

3
A. Khan, R. Singh, P. Gupta et al. Journal of Molecular Liquids 339 (2021) 116829

carboxylic group in octanoic acid at 1415 cm 1 is downshifted to 100


1402–1409 cm 1 in the AG-C8 DESs. The NAH, C@O, OAH groups
of HBA and HBD participated in hydrogen bonding (NAH. . ...OA),
(C@O. . ..HA), (OAH. . ..NA); consequently, the bond lengths of
75
these linkages increased, and their vibrational peaks are red
shifted, signifying the preparation of AG-C8 DESs [28–32]. The FTIR

Weight, %
peaks in a range of 2800–3000 cm 1 are assigned to asymmetric
and symmetric stretching modes of CH2 and CH3 units of octyl 50
chain in AG-C8 DESs. Moreover, the bending mode of C-H at
1465 cm 1 confirmed the alkyl chain in AG-C8 DESs.
1
H and 13C NMR spectra further confirmed the chemical struc-
25 AG-C8-(1:2)
ture and synthesis of AG-C8 DESs. Table 1 presents the chemical AG-C8-(1:4)
structure of AG-C8 DESs along with the assignment of chemical AG-C8-(1:8)
shifts associated with 1H NMR spectra of AG-C8 DESs. The protons AG
of terminal CH3 (He) in octanoic acid exhibited a chemical shift at a 0
shielded range of 0.6–0.9 ppm with respective 13C shift at 12– 60 120 180 240 300
15 ppm (Figs. S1-S3; Electronic Supplementary Information). The Temperature, °C
1
H and 13C chemical shifts at 1.2–1.4 ppm and 25–35 ppm, respec-
tively, are assigned to methylene units (Hd) of octanoic acid in the Fig. 3. Thermal decomposition patterns of AG-C8-(1:2), AG-C8-(1:4), AG-C8-(1:8)
AG-C8 DESs. The CH2 unit adjacent to the carboxylic group of fatty DESs, and the aminoguanidine bicarbonate salt under the N2 flow.
acid (Hc) showed chemical shifts at 2.3–2.5 ppm (1H) and the cor-
responding 13C shifts at 55 ppm. The unsaturated ammonium and octanoic acid, and the AG-C8-(1:2) showed the highest thermal
carboxylic protons (Hb) are highly deshielded and showed chemi- stability among all AG-C8 DESs.
cal shifts at 9.9–13.1 ppm with a corresponding 13C shifts at The viscosity of DESs plays a vital role in lubricant formulation.
160–175 ppm. The protons of the amine groups in AG-C8 DESs The viscosity and thin film formation capability of liquid lubricants
are appeared at the downfield region because of the higher elec- are governed by the chemical composition of lube base oil and the
tronegativity of nitrogen and exhibited chemical shifts at 6.6– blended additives. The size, charge, and composition of HBA and
8.6 ppm. The chemical shifts based on 1H and 13C confirmed the HBD, along with the interactive forces viz., hydrogen linkages,
preparation of AG-C8 DESs. van der Waals, dipole–dipole interactions, etc., control the viscos-
Thermal stability of DESs is governed by chemical composition, ity of DESs [36,37]. Fig. 4 shows the viscosity of octanoic acid
size, and structure of HBA and HBD, hydrogen bonding sites and and its DESs with a function of temperature. The viscosity of all
interactions, charge delocalization in the salt, polarity of HBA/ samples gradually decreased with the rising of temperature, and
HBD, and so on [28,33–35]. Fig. 3 presents the thermal degradation it is attributed to the increasing of thermal kinetic energy, which
patterns of aminoguanidine bicarbonate salt and its DESs with loosens the interactive forces. The viscosity of AG-C8-(1:2) DES is
octanoic acid, i.e., AG-C8-(1:2), AG-C8-(1:4), and AG-C8-(1:8). noted to be significantly higher, and it gradually decreased with
The DESs showed significantly higher thermal stability compared the increasing mole ratio of octanoic acid. These results suggest a
to aminoguanidine bicarbonate salt. The electronegative elements higher degree of hydrogen-based interactions when AG-C8-(1:2)
(N and O) based ample functionalities (NAH, OAH, C@O) make the DES is prepared using 1 mol of aminoguanidine bicarbonate salt
aminoguanidine bicarbonate salt and octanoic acid to be net- with 2 mol octanoic acid. The octanoic acid contains a long alky
worked with each other via hydrogen-based interaction as shown chain. The higher mole ratios of octanoic acids in AG-C8-(1:4)
in Fig. 1; consequently, the AG-C8 DESs showed higher thermal and AG-C8-(1:8) DESs is understood to furnish a larger degree of
stability than aminoguanidine bicarbonate salt. The higher thermal steric hindrance by longer alky chains of octanoic acid, which
stability of AG-C8-(1:2) DES could be attributed to a higher degree might reduce the H-based interaction with aminoguanidine
of hydrogen interaction between the HBA and HBD. The higher bicarbonate salt; consequently, the viscosity of the AG-C8
mole ratios of octanoic acid (HBD) in AG-C8-(1:4) and AG-C8- DESs decreased with increasing of mole ratio of HBD, i.e., octanoic
(1:8) DESs plausibly lead to steric hindrance induced by a higher acid [38].
degree of longer alkyl chain (octanoic acid) and decrease the Rheological measurements examine the fluidic characteristics
hydrogen-based interaction. Therefore, the thermal stability of of aminoguanidine bicarbonate salt-based DESs. Fig. 5a shows no
AG-C8 DESs gradually reduced with an increasing mole ratio of measurable changes in viscosity of AG-C8-(1:4) and AG-C8-(1:8)

Table 1
Assignment of 1H chemical shifts of aminoguanidine based-DESs.

Symbol Protons Chemical Shift, ppm

AG-C8-(1:2) AG-C8-(1:4) AG-C8-(1:8)


He ACH3 0.8–0.9, d, 6H 0.6–0.9, d, 12H 0.7–0.9, d, 24H
Hd ACH2 1.2–1.4, m, 16H 1.2–1.4, m, 32H 1.2–1.4, m, 64H
Hf H2CACAC@O 1.4–1.8, m, 4H 1.4–1.8, m, 8H 1.4–1.8, m, 16H
Hc H2CAC@O 2.3–2.5, m, 4H 2.3–2.5, m, 8H 2.3–2.5, m, 16H
Hb ACOOH/+NAH 9.9–13.1, s/m, 5H 9.9–13.1, s/m, 7H 9.9–13.1, s/m, 11H
Ha ANAH 6.5–8.6, m, 5H 6.5–8.6, m, 5H 6.5–8.6, m, 5H

4
A. Khan, R. Singh, P. Gupta et al. Journal of Molecular Liquids 339 (2021) 116829

800 1650

Kinematic Viscosity, mm2.s-1


(a)
-1

100 AG-C8-(1:2) 1600 AG-C8-(1:2)


700
Kinematic Viscosity, mm .s

AG-C8-(1:4)
2

1550
600 75 AG-C8-(1:8)

Viscosity, mPa.S
C8 1500
500 50

90
400 25 AG-C8-(1:4)
80
300 0
30 40 50 60 70 80 90 100
Temperature, °C
70
200
AG-C8-(1:8)
60
100
50
0
0 100 200 300 400 500 600
30 40 50 60 70 80 90 100
-1
Shear Rate, S
Temperature, °C

Fig. 4. Kinematic viscosity of AG-C8-(1:2), AG-C8-(1:4), AG-C8-(1:8) DESs, and 50


octanoic acid (HBD). (b)
1000 40

Shear Stress, Pa
30

DESs, whereas the viscosity of AG-C8-(1:2) DES marginally


800

Shear Stress, Pa
20
increased (<5%) over the shear rate of 0–600 s 1. These results sug-
gest that hydrogen-driven interaction between the aminoguani- 10

dine bicarbonate salt (HBA) and the octanoic acid (HBD) 600 0
0 100 200 300 400 500 600
remained intact even at the higher shear rate, consequently no Shear Rate, S-1
AG-C8-(1:2)
measurable changes in viscosity of AG-C8 DESs. The linear correla- AG-C8-(1:4)
400 AG-C8-(1:8)
tion between the shear stress and shear rate (Fig. 5b) further signi-
fied the Newtonian fluidic nature of aminoguanidine bicarbonate
salt-based DESs. The slope of linear correlation between the shear 200
stress and shear rate is governed by hydrogen-based interactive
forces, driven by ratios of aminoguanidine bicarbonate salt and 0
octanoic acid in these DESs. 0 100 200 300 400 500 600
The lube oil thinning effect at higher temperatures leads to -1
Shear Rate, S
direct contact between the tribopair as the thin lube oil is easily
squeezed out from the contact interfaces, particularly under the Fig. 5. Changes in (a) viscosity and (b) shear stress as a function of shear rate for
boundary and mixed lubrication regimes. The viscosity index pre- AG-C8-(1:2), AG-C8-(1:4), and AG-C8-(1:8) DESs.
sents the change in viscosity relative to temperature for lube oils.
Higher viscosity index means low thinning of lube oil at higher selected as a representative DES for dose optimization to SN 150
temperatures. The viscosity index improvers are used as additive lube base oil. Fig. 6 presents the changes in average WSD and coef-
to lube oils for minimizing the thinning of lube oil at higher tem- ficient of friction for steel tribopair while lubricated with variable
peratures. The lube oil with low viscosity index indicates the thin- doses of AG-C8-(1:4) DES. The steel balls lubricated with SN 150
ning effect at the higher temperature. In general, lube oil should lube oil showed a high average coefficient of friction and WSD,
have a viscosity index of more than 100 for its good performance, i.e., 0.11 and 743 lm. The non-polar hydrocarbon contents of SN
and it favors the lubrication performance at higher temperatures 150 lube oil show poor affinity towards steel surfaces, which could
by inhibiting the lube thinning effect. The blending of 1.5% be accountable for the lack of lubricious tribo-film between the
aminoguanidine bicarbonate salt-based DESs in SN 150 mineral steel balls under the tribo-stress. The AG-C8-(1:4) DES as an addi-
lube oil increased the viscosity index of SN 150 lube base oil by tive to SN 150 lube oil improved the tribological properties by
12% (Table 2). The improved viscosity index of SN 150 lube oil by reducing the average coefficient of friction and WSD. The 0.5% dose
1.5% AG-C8 DESs promises the applicability of resultant lubricants of AG-C8-(1:4) DES decreased the friction and WSD of SN 150 lube
at higher temperatures and the possibility to develop lubricious oil by 12% and 28%, respectively. The increasing quantity of AG-C8-
tribo thin film between the tribopair for the enhancement of lubri- (1:4) DES showed a gradual enhancement of tribological perfor-
cation effect. mance. A 1.5% AG-C8-(1:4) DES as an additive showed maximum
Tribological properties of AG-C8 DESs as additives to SN 150 reduction in the coefficient of friction (30%) and WSD (40%).
mineral lube base oil are probed using four-ball contact geometry Further increasing of dose showed no measurable change in wear.
as per the ASTM D4172B standard test method. The AG-C8-(1:4) is Therefore, a 1.5% dose of AG-C8-(1:4) DES is considered an

Table 2
Kinematic viscosity and viscosity index of SN 150 lube base oil and 1.5% blend of each DES in the SN 150 lube oil.

No. Sample Detail Kinematic viscosity, mm2.s 1


Viscosity Index
At 40 °C At 100 °C
1 SN 150 Lube oil 30.55 5.28 104
2 1.5% AG-C8-(1:2) in SN 150 Lube oil 31.69 5.60 116
3 1.5% AG-C8-(1:4) in SN 150 Lube oil 30.67 5.54 113
4 1.5% AG-C8-(1:8) in SN 150 Lube oil 30.23 5.40 114

5
A. Khan, R. Singh, P. Gupta et al. Journal of Molecular Liquids 339 (2021) 116829

Average Wear Scar Diameter, µm


0.12 800
Average Coefficient of Friction

50

Wear Volume x 10-4, mm3


0.11 700
40

600
0.10 30

71.5%
500
0.09 20

90.1%
400

94.0%
0.08 10

300
0.0 0.5 1.0 1.5 2.0 0
Doping of AG-C8 (1:4) in SN-150 Lube Base Oil, % w/v 150 8) 4) 2)
SN- -(1: -(1: -(1:
G-C8 G -C8 G-C8
A A A
Fig. 6. Changes in the average coefficient of friction and wear scar diameter as a 1.5% 1.5% 1.5%
function of variable dose of AG-C8-(1:4) DES, blended in SN 150 lube oil. Load:
392 N, temperature: 75 °C, speed: 1200 rpm, test duration: 1 h. Fig. 8. Changes in wear volume of steel tribopair lubricated with SN 150 lube oil
Average Wear Scar Diameter, µm and its blend with 1.5% AG-D8 DESs having a variable mole ratio of octanoic acid.
Average Coefficient of Friction

0.12 800
(a)
0.11 700

0.10 600

0.09 500

0.08 400

0.07
300
- 1 50 :8) :4) (1:2
)
SN -C 8 (1 -C 8 (1 -C8
-AG - AG -AG
1. 5 % 1. 5 % 1.5%
Coefficient of Friction

0.12 (b)

0.10 SN-150

0.08
1.5% AG-C8-(1:4) in SN-150
0.06

0.04
600 1200 1800 2400 3000 3600
Run Time, Seconds

Fig. 7. (a) Changes in average coefficient of friction and WSD for steel tribopair
lubricated with SN 150 lube oil and its blend with 1.5% of AG-C8 DESs having a
variable mole ratio of octanoic acid. (b) Coefficient of friction between steel
tribopair in the presence of SN-150 lube oil and its blend with 1.5% AG-C8-(1:4)
DES. All tribological tests are conducted as per the ASTM D4172B standard test
method.

optimized concentration for tribo-performance evaluation. The


polar functionalities in DESs are believed to facilitate their interac-
tions with contact interfaces of steel tribopair and improve the tri-
bological performance of SN 150 lube oil [12,39].
Fig. 7a shows the average coefficient of friction and WSD of
steel balls lubricated with SN 150 lube oil and its blend with
1.5% of AG-C8 DESs having a variable mole ratio of octanoic acid.
Fig. 9. FESEM images of worn scars developed on steel balls, which were lubricated
All DESs exhibited a significantly low coefficient of friction and with (a-b) SN 150 lube oil and its blend with1.5% (c-d) AG-C8-(1:8), (e-f) AG-C8-
wear compared to SN 150 lube oil. The degrees of friction and wear (1:4), and (g-h) AG-C8-(1:2) DESs.

6
A. Khan, R. Singh, P. Gupta et al. Journal of Molecular Liquids 339 (2021) 116829

reduction increased with decreasing mole ratio of octanoic acid in increased with increasing the mole ratio of octanoic acid in the
the AG-C8 DESs. Among all DESs, the 1.5% AG-C8-(1:2) showed a AG-C8 DESs-based lubricant system. The EDS measurements of
minimum coefficient of friction and wear, i.e., 31 and 46% lower worn scars of lubricated steel balls probed the role of DESs in tribo
than the SN 150 lube oil. The higher viscosity and polar compo- thin film formation.
nents in AG-C8-(1:2) DES are believed to furnish the good quality Fig. 10 displays microscopic images of worn scars and elemental
lubricous thin film over the contact interfaces of steel tribopair distribution of the corresponding area developed on steel balls
for lowering the friction and wear [40,41]. The DESs yielded the during the lubrication test using SN 150 lube oil and its blend with
tribo-induced thin film of low shear strength, which facilitated 1.5% AG-C8-(1:2), AG-C8-(1:4), and AG-C8-(1:8) DESs. The regular
the shearing under the tribo-stress and avoided the direct contact distribution of carbon and nitrogen over the worn scars lubricated
between the steel tribopair. The good affinity of these DESs with the DESs-based system as deduced from their elemental map-
towards the steel tribo surfaces and the formation of tribochemical ping revealed the role of DESs for developing the tribochemical
thin film is further supported by a low and stable friction profile of thin film. Nitrogen is a characteristic element of the AG-C8-based
1.5% AG-C8(1:4) DES in SN 150 lube oil (Fig. 7b). lubricant system. Its regular distribution and intensification of car-
Lubricous and protective tribo thin film by DESs over the con- bon signified the DESs-based tribochemical thin film formation,
tact interfaces minimizes the wear. Fig. 8 displays the average wear which is accountable for lowering the coefficient of friction and
volume of steel balls lubricated with SN 150 lube oil and its blend preventing the direct contact between tribo-surfaces to minimize
with 1.5% AG-D8 DESs having a variable mole ratio of octanoic acid. the wear. Fig. 11 displays the atomic% of elements detected over
The 1.5% AG-D8 DESs as a lubricious additive to SN 150 lube oil the worn scars based on EDS measurement. The steel balls lubri-
decreased the wear volume of steel balls by 94%. The AG-C8- cated with SN 150 lube base oil showed no nitrogen signature,
(1:2) DES showed minimum wear volume, and it was attributed whereas 36.5 at% carbon is detected. The lubrication of steel balls
to the higher polar components and viscosity of AG-C8-(1:2), with 1.5% AG-C8 DESs increased the carbon at% to 60.1, and the
which formed the good quality protective thin film for minimizing nitrogen has emerged as a new element with 6.5 at%. The enhanced
the wear events. carbon at% and emergence of nitrogen could be attributed to the
Fig. 9 displays representative electron microscopic images of participation of octanoic acid (HBD) and aminoguanidine salt
worn scars developed on steel balls during the tribo tests using (HBA) of DESs in the tribochemical thin film formation. The fatty
SN 150 lube oil and 1.5% AG-C8 DESs. The steel balls lubricated acids, having a long alkyl chain, show adequate lubrication because
with SN 150 lube oil developed a larger wear scar and showed dee- of the carboxylic group-driven interaction with the steel surface.
per scratches and furrows (Fig. 9a-b) in the sliding direction. It They adhere to the tribo surfaces under the mixed and boundary
could be associated with the lack of the lubricous thin film by lubrication regime and form the lubricious thin film [42,43].
non-polar hydrocarbon contents of SN 150 lube oil. The steel balls The affinity of DESs and ionic liquids towards the engineering
lubricated with 1.5% AG-C8 DESs decreased the WSD to a signifi- surfaces and their nanoscopic layering structure play an important
cant extent. The mole ratio of fatty acid in the DESs governed the role in governing the tribological properties [23,38,44]. The struc-
size of WSD. The 1.5% AG-C8-(1:2), having a minimum mole ratio ture of nanoscopic layers of DESs over the engineering surfaces is
of octanoic acid and highest viscosity among all samples, devel- controlled by charge distribution in the salt (HBA), hydrogen inter-
oped the smallest WSD (Fig. 9g-h). The wear scar gradually action, the chemical composition of HBA and HBD, and the charge

Fig. 10. Microscopic images of worn scars and elemental distribution over the corresponding area of steel balls lubricated with (a) SN 150 and 1.5% blend of (b) AG-C8-(1:2),
(c) AG-C8-(1:4), and (d) AG-C8-(1:8) DESs in SN 150 lube oil.

7
A. Khan, R. Singh, P. Gupta et al. Journal of Molecular Liquids 339 (2021) 116829

80 4. Conclusion

60.1%
70 SN-150 Oil Aminoguanidine salt-based DESs were synthesized by a single-
1.5% AG-C8 DESs step facile approach using variable mole ratios of octanoic acid as
60 an HBD. The AG-C8-(1:2) DES having a two-mole ratio of octanoic
45.5%

acid showed the highest viscosity and thermal stability among all
50
Atomic %

36.5%
samples. The increasing mole ratio of octanoic acids in AG-C8-
(1:4) and AG-C8-(1:8) DESs decreased the magnitude of
40
hydrogen-based interaction because of steric hindrance by a higher
degree of a longer alkyl chain; as a result, these DESs showed lower

17.8%
30 16.6%
viscosity and thermal stability. The good miscibility of these DESs
15.3%

20 in SN 150 lube base oil and thin film formation capability make

6.5%
them potential candidates for new generation lubricant formula-
10 1.0% tion. The AG-C8 DESs as additives to SN 150 mineral lube base
0.3% oil improved the tribological properties of steel by reducing the
0 coefficient of friction (30%) and wear volume (94%). The degrees
Iron Carbon Oxygen Chromium Nitrogen of friction and wear reductions are noted to be governed by the
Fig. 11. Quantification of tribochemical thin film over the steel balls lubricated
composition of DESs. The AG-C8-(1:2)-based lubricant system with
with SN 150 lube base oil and its blend with 1.5% AG-C8 DESs. Atomic% of all a higher degree of hydrogen linkages yielded a good quality tribo
elements in the thin films over the worn scars is quantified based on multiple EDS thin film; consequently, low and stable coefficient of friction. The
spectra. affinity of amino-rich aminoguanidine and octanoic acid towards
iron surfaces under the tribo-stress promoted the tribochemical
thin film formation, supported by spectroscopic measurement.
on engineering surfaces [23]. Herein, the AG-C8-(1:2) DES exhibits Being halogen-free, economically viable, prepared by a facile
the highest degree of hydrogen interaction and maximum viscosity approach, good miscibility with mineral lube base oils, and signif-
among all studied samples. The higher degree of interaction might icant improvement in viscosity index, friction, and wear make
allow the AG-C8-(1:2) DES to generate good quality quantized lay- these DESs potential candidates for new generation lubricant
ering structure on the tribo interfaces, which facilitated the sliding formulation.
events and reduced the friction. The atomic force microscopic stud-
ies demonstrated the molecular layering of confined ester and CRediT authorship contribution statement
polyalphaolefins on steel surface in the nanoscale range [45].
Moreover, the carboxylic group of fatty acids and the highly polar Amzad Khan: Conceptualization, Data curation, Formal analy-
nature of electron-rich nitrogen in the aminoguanidine salt facili- sis, Investigation, Methodology, Validation, Writing - original draft.
tated their interaction with steel surfaces, particularly under con- Raghuvir Singh: Formal analysis. Piyush Gupta: Formal analysis.
tact stress, and formed a good quality tribochemical film for the Kanika Gupta: Investigation, Methodology. Om P Khatri: Concep-
protection against the undesirable events of wear [46,47]. The tualization, Data curation, Funding acquisition, Investigation, Pro-
AG-C8-(1:4) and AG-C8-(1:8) DESs having a higher mole ratio of ject administration, Supervision, Visualization, Writing - review &
octanoic acid probably decreased hydrogen-based interaction editing.
because of the steric hindrance by a higher degree of longer alkyl
chain; as a result, these DESs showed lower viscosity (Fig. 4). Plau- Declaration of Competing Interest
sibly, the lower viscosity minimized the layering structure of such
a system close to engineering interfaces; as a result, AG-C8-(1:4) The authors declare that they have no known competing finan-
and AG-C8-(1:8) DESs showed a higher coefficient of friction than cial interests or personal relationships that could have appeared
the AG-C8-(1:2)-based lubricant system. The adsorption kinetic of to influence the work reported in this paper.
the electron-rich amine group towards the steel surface governs by
steric hindrance. The faster adsorption of amine molecules forms
Acknowledgment
an effective lubricious thin film over the worn scar and dictates
the friction [46]. The wearing of steel tribo-interfaces exposes the
The authors are grateful to the Director, CSIR-Indian Institute of
highly active (pristine) surfaces for the adsorption of surface-
Petroleum, Dehradun, for his kind permission to publish these
active additives under contact stress. The layering of DESs on the
results. Authors are thankful to CSIR, India, for financial support
freshly prepared tribo-surfaces of steel and high contact stress
(OLP-884) and ASD, UWRD, and CMSD of CSIR-IIP, Dehradun, for
advance the tribochemical thin film formation [23,48]. The AG-
analytical support. AK and KG are grateful to CSIR and UGC for
C8-(1:2)-based lubricant system having a good layering structure
the research fellowship support.
via hydrogen linkages between the aminoguanidine bicarbonate
salt (HBA) and the octanoic acid (HBD) furnished good quality tribo
Appendix A. Supplementary material
thin film driven by the interaction of amine-rich cation in
aminoguanidine salt. Moreover, the good chemical affinity of octa-
Supplementary data to this article can be found online at
noic acid also facilitated the thin-film formation and yielded the
https://doi.org/10.1016/j.molliq.2021.116829.
tribochemical thin film of low shear strength for enhancement of
tribological properties. The steric hindrance by a higher degree of
References
octyl chain in AG-C8-(1:4) and AG-C8-(1:8) DESs-based lubricant
system compromised the interfacial layering and decreased the [1] C.J. Reeves, P.L. Menezes, Advancements in eco-friendly lubricants for
adsorption kinetics, consequently, poor quality tribochemical thin tribological applications: past, present, and future, in: J. Davim (Ed.),
film formation. It is further supported by a gradual decline in both Ecotribology, Springer, 2016, pp. 41–61.
[2] R.W. Carpick, A. Jackson, W.G. Sawyer, N. Argibay, P. Lee, A. Pachon, R.M.
wear and friction coefficient in the AG-C8 DESs-based lubricant Gresham, The tribology opportunities study: can tribology save a quad?, Tribol.
system when the molar ratio of octanoic acid decreased. Lubr. Techno. 72 (2016) 44.

8
A. Khan, R. Singh, P. Gupta et al. Journal of Molecular Liquids 339 (2021) 116829

[3] W. Dai, B. Kheireddin, H. Gao, H. Liang, Roles of nanoparticles in oil lubrication, [27] D.P. Arcon, F.C. Franco Jr, All-fatty acid hydrophobic deep eutectic solvents
Tribol. Int. 102 (2016) 88–98. towards a simple and efficient microextraction method of toxic industrial
[4] A. Chouhan, H.P. Mungse, O.P. Khatri, Surface chemistry of graphene and dyes, J. Mol. Liq. 318 (2020) 114220.
graphene oxide: a versatile route for their dispersion and tribological [28] H. Ghaedi, M. Ayoub, S. Sufian, B. Lal, Y. Uemura, Thermal stability and FT-IR
applications, Adv. Colloid Interface Sci. 283 (2020) 102215. analysis of Phosphonium-based deep eutectic solvents with different
[5] Y. Zhou, J. Qu, Ionic liquids as lubricant additives: a review, ACS Appl. Mater. hydrogen bond donors, J. Mol. Liq. 242 (2017) 395–403.
Interfaces 9 (2017) 3209–3222. [29] D.L. Pavia, G.M. Lampman, G.S. Kriz, J.A. Vyvyan, Introduction to Spectroscopy,
[6] Z. Tang, S. Li, A review of recent developments of friction modifiers for liquid Nelson Education, 2014.
lubricants (2007–present), Curr. Opin. Solid State Mater. Sci. 18 (2014) 119– [30] H. Wang, S. Liu, Y. Zhao, J. Wang, Z. Yu, Insights into the hydrogen bond
139. interactions in deep eutectic solvents composed of choline chloride and
[7] S.S. Rawat, A. Harsha, O.P. Khatri, R. Wäsche, Pristine, reduced, and alkylated polyols, ACS Sustain. Chem. Eng. 7 (2019) 7760–7767.
graphene oxide as additives to paraffin grease for enhancement of tribological [31] K.-I. Takei, R. Takahashi, T. Noguchi, Correlation between the hydrogen-bond
properties, J. Tribol. 143 (2021) 021903. structures and the C=O stretching frequencies of carboxylic acids as studied by
[8] H. Nautiyal, S. Kumari, R. Tyagi, U. Rao, O.P. Khatri, Evaluation of tribological density functional theory calculations: theoretical basis for interpretation of
performance of copper-based composites containing nano-structural 2D infrared bands of carboxylic groups in proteins, J. Phys. Chem. B 112 (2008)
materials and their hybrid, Tribol. Int. 153 (2021) 106645. 6725–6731.
[9] T. Torimoto, T. Tsuda, K.i. Okazaki, S. Kuwabata, New frontiers in materials [32] R. Gautam, N. Kumar, J.G. Lynam, Theoretical and experimental study of
science opened by ionic liquids, Adv. Mater. 22 (2010) 1196–1221. choline chloride-carboxylic acid deep eutectic solvents and their hydrogen
[10] S. Perkin, Ionic liquids in confined geometries, Phys. Chem. Chem. Phys. 14 bonds, J. Mol. Struct. 1222 (2020) 128849.
(2012) 5052–5062. [33] Q. Abbas, L. Binder, Synthesis and characterization of choline chloride based
[11] X. Gong, L. Li, Nanometer-thick ionic liquids as boundary lubricants, Adv. Eng. binary mixtures, ECS Trans. 33 (2010) 49.
Mater. 20 (2018) 1700617. [34] F.U. Shah, I.A. Khan, P. Johansson, Comparing the thermal and electrochemical
[12] R. Gusain, P. Gupta, S. Saran, O.P. Khatri, Halogen-free bis (imidazolium)/bis stabilities of two structurally similar ionic liquids, Molecules 25 (2020) 2388.
(ammonium)-di [bis (salicylato) borate] ionic liquids as energy-efficient and [35] P.J. Carvalho, S.P. Ventura, M.L. Batista, B. Schröder, F. Gonçalves, J. Esperança,
environmentally friendly lubricant additives, ACS Appl. Mater. Interfaces 6 F. Mutelet, J.A. Coutinho, Understanding the impact of the central atom on the
(2014) 15318–15328. ionic liquid behavior: phosphonium vs ammonium cations, J. Chem. Phys. 140
[13] A.I. Siriwardana, Industrial applications of ionic liquids, Electrochemistry in (2014) 064505.
Ionic Liquids, Springer, 2015, pp. 563–603. [36] R.L. Gardas, J.A. Coutinho, A group contribution method for viscosity
[14] R. Gusain, S. Dhingra, O.P. Khatri, Fatty-acid-constituted halogen-free ionic estimation of ionic liquids, Fluid Phase Equilib. 266 (2008) 195–201.
liquids as renewable, environmentally friendly, and high-performance [37] G. Yu, D. Zhao, L. Wen, S. Yang, X. Chen, Viscosity of ionic liquids: database,
lubricant additives, Ind. Eng. Chem. Res. 55 (2016) 856–865. observation, and quantitative structure-property relationship analysis, AIChE
[15] M. Amde, J.-F. Liu, L. Pang, Environmental application, fate, effects, and J. 58 (2012) 2885–2899.
concerns of ionic liquids: a review, Environ. Sci. Technol. 49 (2015) 12611– [38] A. Khan, R. Gusain, M. Sahai, O.P. Khatri, Fatty acids-derived protic ionic liquids
12627. as lubricant additive to synthetic lube base oil for enhancement of tribological
[16] E.L. Smith, A.P. Abbott, K.S. Ryder, Deep eutectic solvents (DESs) and their properties, J. Mol. Liq. 293 (2019) 111444.
applications, Chem. Rev. 114 (2014) 11060–11082. [39] A. Khan, S.R. Yasa, R. Gusain, O.P. Khatri, Oil-miscible, halogen-free, and
[17] M.T. Donato, R. Colaço, L.C. Branco, B. Saramago, A review on alternative surface-active lauryl sulphate-derived ionic liquids for enhancement of
lubricants: ionic liquids as additives and deep eutectic solvents, J. Mol. Liq. 333 tribological properties, J. Mol. Liq. 318 (2020) 114005.
(2021) 116004. [40] A. Westerholt, M. Weschta, A. Bosmann, S. Tremmel, Y. Korth, M. Wolf, E.
[18] K. Radošević, M.C. Bubalo, V.G. Srček, D. Grgas, T.L. Dragičević, I.R. Schlücker, N. Wehrum, A. Lennert, M. Uerdingen, Halide-free synthesis and
Redovniković, Evaluation of toxicity and biodegradability of choline chloride tribological performance of oil-miscible ammonium and phosphonium-based
based deep eutectic solvents, Ecotoxicol. Environ. Saf. 112 (2015) 46–53. ionic liquids, ACS Sustain. Chem. Eng. 3 (2015) 797–808.
[19] S. Lawes, S. Hainsworth, P. Blake, K. Ryder, A. Abbott, Lubrication of steel/steel [41] G. Huang, Q. Yu, Z. Ma, M. Cai, Probing the lubricating mechanism of oil-
contacts by choline chloride ionic liquids, Tribol. Lett. 37 (2010) 103–110. soluble ionic liquids additives, Tribol. Int. 107 (2017) 152–162.
[20] I. Garcia, S. Guerra, J. de Damborenea, A. Conde, Reduction of the coefficient of [42] C.J. Reeves, P.L. Menezes, T.-C. Jen, M.R. Lovell, The influence of fatty acids on
friction of steel-steel tribological contacts by novel graphene-deep eutectic tribological and thermal properties of natural oils as sustainable biolubricants,
solvents (DESs) lubricants, Lubricants 7 (2019) 37. Tribol. Int. 90 (2015) 123–134.
[21] A.P. Abbott, E.I. Ahmed, R.C. Harris, K.S. Ryder, Evaluating water miscible deep [43] R.R. Sahoo, S. Biswas, Frictional response of fatty acids on steel, J. Colloid
eutectic solvents (DESs) and ionic liquids as potential lubricants, Green Chem. Interface Sci. 333 (2009) 707–718.
16 (2014) 4156–4161. [44] R. Espinosa-Marzal, A. Arcifa, A. Rossi, N. Spencer, Microslips to ‘‘avalanches”
[22] M. Antunes, A.-S. Campinhas, M. de Sá Freire, F. Caetano, H.P. Diogo, R. Colaço, in confined, molecular layers of ionic liquids, J. Phys. Chem. Lett. 5 (2014) 179–
L.C. Branco, B. Saramago, Deep eutectic solvents (DES) based on sulfur as 184.
alternative lubricants for silicon surfaces, J. Mol. Liq. 295 (2019) 111728. [45] M.-D. Krass, G. Krämer, U. Dellwo, R. Bennewitz, Molecular layering in
[23] J.E. Hallett, H.J. Hayler, S. Perkin, Nanolubrication in deep eutectic solvents, nanometer-confined lubricants, Tribol. Lett. 66 (2018) 1–10.
Phys. Chem. Chem. Phys. 22 (2020) 20253–20264. [46] P.C. Nalam, A. Pham, R.V. Castillo, R.M. Espinosa-Marzal, Adsorption behavior
[24] M. Slaný, L’. Jankovič, J. Madejová, Structural characterization of organo- and nanotribology of amine-based friction modifiers on steel surfaces, J. Phys.
montmorillonites prepared from a series of primary alkylamines salts: Mid-IR Chem. 123 (2019) 13672–13680.
and near-IR study, Appl. Clay Sci. 176 (2019) 11–20. [47] R. Gusain, O.P. Khatri, Fatty acid ionic liquids as environmentally friendly
[25] G. Richner, G. Puxty, Assessing the chemical speciation during CO2 absorption lubricants for low friction and wear, RSC Adv. 6 (2016) 3462–3469.
by aqueous amines using in situ FTIR, Ind. Eng. Chem. Res. 51 (2012) 14317– [48] N. Gosvami, J. Bares, F. Mangolini, A. Konicek, D. Yablon, R. Carpick,
14324. Mechanisms of antiwear tribofilm growth revealed in situ by single-asperity
[26] C. Karthikeyan, S. Rajeswari, S. Maruthamuthu, K. Subramanian, G. Rajagopal, sliding contacts, Science 348 (2015) 102–106.
Biogenic ammonia for CO2 capturing and electrochemical conversion into
bicarbonate and formate, J. CO2 Util. 6 (2014) 53–61.

You might also like