Subscriber access provided by Bibliothèque de l'Université Paris-Sud
Combustion
Impact of diesel engine oil additives-soot interactions on physiochemical, oxidation and wear characteristics of soot Kimaya Vyavhare, Sujay Bagi, Mihir Patel, and Pranesh B Aswath Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.8b03841 • Publication Date (Web): 04 Apr 2019 Downloaded from http://pubs.acs.org on April 4, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
1
Impact of diesel engine oil additives-soot interactions on
2
physiochemical, oxidation and wear characteristics of soot
3 4
Kimaya Vyavhare†, Sujay Bagi†1, Mihir Patel†2 and Pranesh B. Aswath†*
5 6
† Department of Materials Science and Engineering, University of Texas at Arlington, Arlington,
7
TX 76019, USA
8
†1 Department of Mechanical Engineering, Massachusetts Institute of Technology, 77
9
Massachusetts Avenue, Cambridge, Massachusetts, 02139, United States.
10
†2 Product Development, Chevron Lubricants, San Ramon, California, 94583, United States.
11
*Corresponding author:
[email protected] 12 13 14 15 16 17 18 19 20
1 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 2 of 54
21
ABSTRACT
22
Carbonaceous soot accumulated in crankcase oil is known to have adverse effect on diesel engine
23
performance, durability and fuel efficiency. The current study is focused on determining the
24
influence of engine oil additive package and soot interactions on crankcase soot chemistry,
25
structure and oxidation. Soot was extracted from the crankcase oil of Mack T-12 dynamometer
26
diesel engine tests and characterized using X-ray absorption near edge structure (XANES)
27
spectroscopy, high-resolution transmission electron microscopy (HR-TEM), high-temperature X-
28
ray diffraction (HT-XRD) and energy dispersive spectroscopy (EDS). Additionally, four-ball wear
29
bench tests were conducted to study the effect of several interactions like antiwear additive-soot,
30
dispersant-soot on lubricity of formulated engine base oils. XANES and HR-TEM analysis suggest
31
presence of chemical compounds originating from engine oil chemistry as either
32
adsorbed/embedded into the turbostratic nano-structure of soot. HT-XRD method was employed
33
to map variations in interplanar spacing of basal plane (002) of turbostratic soot with temperature,
34
indicative of degree of disorder and ease of oxidation. Oxidative reactivity of crankcase soot is
35
found to be strongly dependent on the changes in physical and chemical makeup of carbonaceous
36
soot structure. Wear assessment proved that increase of soot concentration in formulated oils,
37
significantly increases wear of sliding surfaces. SEM-EDS worn surface analysis suggest that soot
38
induced wear occurs through abrasive wear mechanism, where soot antagonistically interact with
39
protective antiwear additive formed tribofilms and exacerbate wear of engine components.
40
Keywords: Diesel Soot; Soot structure; Soot Chemistry; Soot Induced Wear; XANES; Ionic
41
liquids; Ashless Antiwear Additives.
42 43 2 ACS Paragon Plus Environment
Page 3 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
44
1. Introduction
45
Soot is a nanoscopic carbonaceous material produced due to incomplete combustion of
46
hydrocarbon fuel in an internal combustion engine. Excessive soot formation occurs in diesel
47
engines as compared to the gasoline engines because of the difference in the way fuel/air mixture
48
is injected and ignited in them. Of the soot produced during in-cylinder fuel combustion, only 29%
49
reaches the atmosphere through the exhaust tail-pipe 2, with the remainder being deposited on the
50
cylinder walls and piston crown, which eventually ends up in the crankcase oil via blow-by
51
mechanisms 3-5. Rapid soot production with extended drain interval in a heavy-duty diesel engine
52
can lead to high soot levels of 8 or 10 vol% accumulating in crankcase engine oil 6. It has been
53
shown that the crankcase soot decreases engine efficiency by deteriorating physical and chemical
54
lubrication oil properties of the engine oil and inducing severe friction and wear of engine
55
components 7-11,12,13. However, a thorough understanding is required to apprehend the interaction
56
of crankcase soot and, lubrication oil additive chemistries and its potential effects on engine wear.
57
Soot is found to be in the form of necklace-like agglomerates, which are around 100 nm in size.
58
These agglomerates are composed of collections of smaller, spherical basic particle units called as
59
“primary soot particles”, with diameter around 10-80 nm and the cluster or chain-like soot
60
aggregates defined as “secondary particles”, which are composed of several tens to hundreds of
61
primary soot particles. The primary soot particle has a turbostratic structure wherein crystalline
62
graphite-like layers (known as platelets) are arranged in a concentric manner surrounding an inner
63
core. The factors like fuel composition, engine operating conditions, vehicle mileage can change
64
or modify the physical structure and bulk/surface chemistry of soot and are expected to play an
65
important role in controlling adverse effect of soot. Many studies have been directed to explore
66
the impact of these factors, Sharma et al. 12 used Raman spectroscopy, high resolution transmission
3 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 4 of 54
67
electron microscopy (HR-TEM), X-ray absorption near edge spectroscopy (XANES), X-ray
68
diffraction (XRD) techniques to study soot extracted from crankcase and exhaust and reported that
69
crankcase soot exhibited turbostratic structure with the presence of some nanocrystalline species
70
originating from lubricant chemistry, while exhaust soot showed perfectly ordered graphitic
71
crystalline structure. Uy et al. 14 compared the morphology and chemistry of soot produced through
72
different combustion modes (gasoline soot and diesel soot) using X-ray photoelectron
73
spectroscopy (XPS), Auger electron spectroscopy (AES), Raman spectroscopy, HR-TEM, and X-
74
ray fluorescence (XRF). They reported that gasoline soot contains more lubricant additive
75
elements and wear debris than the diesel soot. Ishiguro et al. 15 studied the structure of soot using
76
HR-TEM and found that primary particle of soot comprised of an inner core of 10nm which is
77
more amorphous and disordered and an outer shell which is more graphitic. Outer shell composed
78
of crystallites with polycyclic aromatic hydrocarbon layers oriented concentrically with the inner
79
core. The effect of fuel composition on the soot structure and physiochemical properties was
80
studied by Lu et al.
81
low-sulfur diesel, and ultra-low sulfur diesel fuel and reported that biodiesel soot has more
82
disordered nanostructure.
83
The oxidative reactivity of soot i.e. the ease of soot burning by downstream after-treatment
84
components (vis-à-vis DPF regeneration) closely depends on the structure and morphology of soot
85
12,21.
86
the catalytic oxygen groups present on primary soot particle generated from oxygenated fuel.
87
Yehliu et al.
88
carbonaceous nanostructure of soot. Extensive work of Vander Wal et al.
89
the effect of soot synthesis conditions (temperature, time and fuel content) on soot particle
16
They characterized soot extracted from engines operated with biodiesel,
Song et al. 22 and Lu et al. 16,23,24 showed that high oxidative reactivity of soot is attributed to
25
reported that soot oxidative reactivity depends on the degree of disorder of 26-28
on understanding
4 ACS Paragon Plus Environment
Page 5 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
90
nanostructure proved that structural changes in the graphene layer (like planar elongation of the
91
individual carbon layers) alters the ratio between basal plane and edge site carbon atoms which
92
increases the rate of soot oxidation. There is active research in the area of soot oxidation that
93
focuses on understanding structure-oxidative reactivity relationship of soot, which evokes us to
94
elucidate the oxidation behavior of crankcase soot. In this research work, the influence of
95
interaction of crankcase soot with oil additives on soot’s chemistry, structure, oxidation reactivity
96
and wear properties is examined which will provide benefit to engine manufacturers and catalyst
97
industry in resolving soot related issues.
98
Soot in the diesel engine oil can lead to wear and premature engine failure 29. Several researchers
99
have addressed the effect of soot on engine wear and various soot caused wear mechanisms have
100
been proposed such as 1) competition of soot with anti-wear additives for adsorption sites at metal
101
surfaces 30, 2) adsorption of active anti-wear component from the oil phase by soot 31,32, 3) abrasion
102
of protective anti-wear films by soot through three body wear mechanism where soot acts as the
103
intermediate body 33-37 and 4) accumulation of soot in the contact inlet, ultimately restricting oil
104
supply at the contact 38-41. Although various wear mechanisms have been suggested, there is still
105
no consensus on the mechanism contributing to high wear rates caused due to high soot loaded
106
lubricating oils, thereby a better understanding of the underlying soot caused wear mechanism is
107
required.
108
In light of the above considerations, in this research work, diesel engine soot extracted from Mack
109
T-12 engine dynamometer crankcase oil was characterized using XANES and HR-TEM. XANES
110
is used in determining the local coordination, valence and chemical bonding of individual
111
elements, coupled with HR-TEM it is possible to characterize the crystalline and amorphous
112
phases in diesel soot. In addition, the oxidation behavior of soot was evaluated using high 5 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 6 of 54
113
temperature X-ray diffraction (HT-XRD), Brunner-Emmett-Teller (BET), and energy dispersive
114
spectroscopy (EDS). The secondary objective of this study was to understand the effects of
115
crankcase soot induced engine wear. A series of experiments were designed to assess the
116
tribological performance of model test oil containing soot, dispersant and antiwear additives.
117
Detailed evaluation of the effect of diesel engine soot on the tribological performance of the three
118
different anti-wear additives, zinc dialkyl dithiophosphate, phosphonium based ionic liquid, and
119
metal free dithiophosphate was performed using four-ball tribological test. An in-depth
120
characterization of the wear surface, as well as chemical properties of formed tribofilms using
121
scanning electron microscopy (SEM) and energy dispersive spectroscopy (EDS), was performed
122
and various phenomenological models for wear mechanisms were proposed. The findings of this
123
paper will help engine oil formulators and lubricant additive manufacturers to develop improved
124
engine lubricants or and optimize existing lubrication oil additive technology capable of mitigating
125
the adverse effect of soot accumulation and enhancing diesel engine efficiency, durability, fuel
126
efficiency, and, simultaneously, reducing harmful emissions.
127
2. Experimental Methodology
128
2.1 Diesel Soot Extraction
129
Crankcase soot was extracted from the used diesel engine oil sample acquired from Mack-T12
130
lubrication qualification dynamometer engine test. At first, the collected sump oil was diluted with
131
hexane in a 50-50 wt% ratio followed by bath ultrasonication for 15 minutes. The oil-hexane
132
mixture was then centrifuged at 12,000 rpm for a period of 2 hours to collect soot. The supernatant
133
was discarded, and the remaining residue of soot was washed with hexane and ultrasonicated,
134
followed by centrifugal process. This process was repeated for three times until the oil-free thick
135
black residue of soot particles was obtained. To further ensure the removal of any trapped oil
6 ACS Paragon Plus Environment
Page 7 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
136
molecules a Soxhlet extraction method was used for 24 hours using hexane as the solvent. The
137
extracted black residue was then dried and ground using a mortar and pestle to break up the
138
agglomerates for further analysis.
139
2.2 Characterization of Diesel Soot – Chemistry, Structure and Oxidation Attributes
140
XANES was used to study the chemical composition of the extracted diesel engine oil soot. This
141
characterization technique helps to understand the local co-ordination of the elements present in
142
the soot structure from near surface to the bulk. The photo-absorption spectra were obtained using
143
a 2.9 GeV storage ring at the Canadian Light Source (CLS), Saskaton, Canada. Three beam lines,
144
soft X-ray micro-characterization (SXRM), variable line grating-plane grating monochromator
145
(VGM-PGM), and spherical grating monochromator (SGM) were used at CLS to acquire K and L
146
shell absorption edge spectra. The phosphorus K-edge (P K-edge), sulfur K-edge (S K-edge), and
147
calcium K-edge (Ca K-edge) spectra were recorded using the SXRM beam line with a photon
148
energy range of 1700-10,000 eV, resolution of 0.2 eV and beam spot size of 4mm × 300µm.
149
Phosphorus L-edge (P L-edge) and sulfur L-edge (S L-edge) spectra were collected using the
150
VGM-PGM beam line operating at the energy range of 5-250 eV with photon resolution of 0.2 eV
151
and beam spot size of 500 µm × 500 µm. Zinc L-edge spectra (Zn L-edge) was acquired at SGM
152
beam station covering energy range between 250 and 2000 eV with photon resolution of 0.2 eV.
153
A beam spot size of 100 µm × 100 µm was chosen to collect spectra at SGM beamline. XANES
154
sample preparation was carried out by placing soot particles using a spatula on the indium foil and
155
gently tapping on the rear of the foil to remove any loose particles. As a result, a very fine layer of
156
particles with a thickness of around 4-5 microns was left on the foil. These foils were then
157
immediately placed in the sample chamber for characterization. Spectra were acquired using total
158
electron yield (TEY) and fluorescence yield (FY) mode. TEY spectra give near-surface sensitive
7 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 8 of 54
159
information of the sample, whereas FY spectra provide information from the bulk of the sample.
160
In the case of P L-edge, the sampling depth of TEY and FY modes is 50 Å and up to 600 Å
161
respectively.
162
HR-TEM was used to characterize the structure of extracted carbonaceous soot particles. High
163
resolution lattice imaging was acquired using a Hitachi H-9500 microscope at an accelerating
164
voltage of 300 kV with a lattice resolution of 0.18 nm.
165
Oxidation characteristics of the soot particles were studied using HT-XRD. The diffraction pattern
166
of the sample was recorded using a Rigaku Smart Lab Diffractometer with CuKα radiation of
167
wavelength 1.5406 ˚A. The diffractometer was operated in Bragg-Brentano focusing geometry
168
with tube voltage at 40 kV, current at 44 mA and 2θ scan range at 10-90˚. The soot sample was
169
placed on a fused silica glass inserted in a heating furnace apparatus, co-planarity of the soot
170
particles surface with the sample holder surface and setting of the sample holder at the position of
171
symmetric reflection geometry was ensured. The heating unit of the furnace was set to raise
172
temperature from room temperature to 700˚C at a 30˚C/min. Phase or composition changes were
173
mapped for each X-ray scan taken at the temperature increments of 50˚C. After cooling the sample
174
from 700˚C, an additional scan at 50˚C was taken to study the final structure of the soot. Interplanar
175
spacing (d002) was calculated using Bragg’s law for each temperature step from the position of the
176
(002) peak. The weight percentage of the residue left behind after soot oxidation was calculated
177
by weighing initial and final weight of the sample. Phase identification of residue was carried out
178
by matching the peak positions and peak RIR (reference intensity ratio) values of model
179
compounds in the ICDD (international center for diffraction data) crystal structure database. In
180
addition, the elemental analysis of the incombustible residue obtained after HT-XRD studies was
8 ACS Paragon Plus Environment
Page 9 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
181
performed using a Vega 35B electron microscope with a fully integrated EDS microanalyzer from
182
TESCAN Instruments.
183
2.3 Bench Test Oil Formulations and Procedure
184
In this study, a total of sixteen test oil formulations were prepared to determine the effect of certain
185
interactions like soot - antiwear additive, soot - dispersant on the wear of tribological contact.
186
Table 1 details the test oil formulations and the respective code names that would be used further
187
as a reference. Oil formulations were composed of base oil, antiwear additives, dispersant, and
188
extracted soot particles. Dispersant – sorbitan monoleate was purchased from Sigma Aldrich.
189
Antiwear additives – zinc dialkyl dithiophosphate (ZDDP) and ashless neutral dialkyl
190
dithiophosphate (DDP) were acquired from the Chevron and BASF respectively. The phosphorus-
191
containing ionic liquid, tetra-n-butyl-phosphonium O,O-diethyl-dithiophosphate (DEDTP) was
192
provided from AC2T research GmbH. Oil formulations were prepared by adding all the antiwear
193
additives at 0.07 wt% phosphorus treat rate and dispersant at 5 wt%. A thorough mixing of all the
194
components in the oil formulations was ensured by using pulse ultrasonication for 15 min in de-
195
ionized water bath. The tribological tests were conducted for each oil formulation immediately
196
after mixing. The tribological performances of the formulated oil were evaluated using Plint TE92
197
four-ball Tribometer. ASTM D2266 was used for testing the antiwear properties of oil
198
formulations in continuous sliding mode under boundary lubrication regime. According to ASTM
199
standard, three AISI E52100 chrome steel balls of ½ inch diameter were clamped together at the
200
bottom and immersed in the formulation to be evaluated. A fourth steel ball held in ball holder was
201
pressed with force of 40 kg into the cavity formed by the three clamped balls. The temperature of
202
the test was regulated at 75˚C and the top ball was rotated at 1200 rpm for 60 min and 72000
203
cycles. All the tests were repeated three times to avoid any discrepancy in the data. After
9 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 10 of 54
Table 1. Overview of test oil formulations. Code Name
Test Oil Formulation
B. O
Group III Base oil
ZDDP_D
B.O + ZDDP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%)
ZDDP_D_Soot_2 wt%
B.O + ZDDP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%) + Soot (2 wt%)
ZDDP _D_Soot _5 wt%
B.O + ZDDP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%) + Soot (5 wt%)
ZDDP_D_Soot_10 wt%
B.O + ZDDP (0.0 7wt% of Phosphorus treat rate) + Dispersant (5 wt%) + Soot (10 wt%)
DEDTP_D
B.O + DEDTP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%)
DEDTP_D_Soot_2 wt%
B.O + DEDTP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%) + Soot (2 wt%)
DEDTP_D_Soot_5 wt%
B.O + DEDTP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%) + Soot (5 wt%)
DEDTP_D_Soot_10 wt%
B.O + DEDTP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%) + Soot (10 wt%)
DDP_D
B.O + DDP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%)
DDP_D_Soot_2 wt%
B.O + DDP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%) + Soot (2 wt%)
DDP_D_Soot_5 wt%
B.O + DDP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt%) + Soot (5 wt%)
DDP_D_Soot_10 wt%
B.O + DDP (0.07 wt% of Phosphorus treat rate) + Dispersant (5 wt %) + Soot (10 wt%)
D_Soot _2 wt%
B.O + Dispersant (5 wt%) + Soot (2 wt%)
D_Soot_5 wt%
B.O + Dispersant (5 wt%) + Soot (5 wt%)
204
D_Soot_10 wt% B.O + Dispersant (5 wt%) + Soot (10 wt%) termination of each test, three balls were removed and cleaned thoroughly using hexane to observe
205
wear scar.
206 207
2.4 Analysis of the Wear Surfaces and Formed Tribofilms 10 ACS Paragon Plus Environment
Page 11 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
208
Stereo-optical microscope (model type: Nikon SMZ 1500) was used to image the wear scars
209
formed on the three stationary balls after the four-ball tribological test. The optical images were
210
obtained at 100X magnification and were further analyzed using image J software to calculate the
211
Wear Scar Diameter (WSD). A precise measurement of the WSD for each test oil formulation
212
was carried out using an average of 12 readings on 6 ball samples from a total of 2 repeat runs.
213
Measured WSD for each ball is the average of longest horizontal striation and longest vertical
214
striation.
215
The nature of the wear occurring on the steel ball specimen surface was determined by carefully
216
examining the surface of the wear tracks using a SEM. A Hitachi S-3000N SEM equipped with an
217
energy dispersive X-ray spectrometer was used in the secondary electron mode to acquire the
218
images of the wear surface at highest magnification of upto 1000X with an accelerating voltage of
219
15-25 kV. Higher resolution images of the specific area of interest helped to understand the wear
220
mechanism in play on the surface. The corresponding elemental maps and composition data
221
collected by EDS micro-analyzer unit were used to study the tribochemistry of the films generated
222
on the rubbing surfaces.
223
3. Results and Discussion
224
3.1 XANES
225
3.1.1 Phosphorus Characterization
226
The P L-edge TEY spectra as shown in figure 1 provides information about local coordination of
227
phosphorus near the surface of the soot particles. To understand the interaction of the lubricating
228
oil additive chemistries with the soot, it is essential to compare the XANES spectra with the
229
different model compounds of the known local chemical environment representing the
230
characteristics of typical P and S based antiwear films as well as the decomposition products of
11 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 12 of 54
231
lubricating oil. As a result, the photo-absorption spectra of the soot are compared with iron
232
phosphate (FePO4), zinc phosphate (ZnPO4), calcium phosphate (Ca3(PO4)2) and calcium
233
pyrophosphate (Ca2P2O7). Phosphorus L-edge spectra is characterized by spin-orbit splitting of
234
phosphorus 2p electrons and excitation to the antibonding orbitals. Transition of spin orbit split of
235
2p electrons between 2p3/2 (L3 edge) and 2p1/2 (L2 edge) is shown by peaks a and b. Peak labelled
236
as c is assigned to transitions to 3p orbitals, which are sensitive to the presence of other elements
237
such as oxygen and cations like Fe or Zn. Peak d is the shape resonance peak characteristic of all
238
phosphates regardless of structure, whether crystalline or glassy 42-44. The presence of peak d in P
239
L-edge spectra of soot suggests that the phosphorous is present as phosphates in soot.
240
12 ACS Paragon Plus Environment
Page 13 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
241
Figure 1. (a) Phosphorus L-edge TEY (b) phosphorus K-edge FY spectra of crankcase soot
242
compared with reference compounds.
243
Comparison of the soot absorption spectra with the model compounds in TEY mode indicates that
244
the pre-edge absorption peak a closely matches with pre-edge of Fe, Zn and Ca phosphates. The
245
characteristic main absorption peak c of the soot aligns with that of Ca3(PO4)2 and rules out the
246
possibility of presence of Fe and Zn phosphates in the soot structure. The misalignment of the
247
main absorption edge peak c and d of the soot to that of undecomposed ZDDP eliminates the
248
possibility of any trapped oil in the soot structure.
249
The P K-edge FY spectra is presented and compared with the model compounds in figure 1. The
250
P K-edge spectra gives more adequate information of the local coordination of phosphorus in the
251
bulk of the soot because the sampling depth of FY (> 8000 Å) at P K-edge is greater than the
252
sampling depth of TEY (50 Å) at P L-edge 50. Excitation of electrons from phosphorus 1s orbital
253
to unoccupied 2p orbital results into a main absorption peak of P k-edge spectra as marked in figure
254
1. The energy at main absorption peak matches with main absorption edge of Zn3(PO4)2 and
255
Ca3(PO4)2, suggesting that Zn or Ca or a mixture of Zn and Ca phosphates is present. Zn3(PO4)2
256
is the decomposition product of lubrication anti-wear additive, ZDDP, while Ca3(PO4)2 arises from
257
the detergent chemistry. Possibility of FePO4 is not likely due to absence of pre-edge of FePO4 in
258
soot spectrum. Presence of short chain polyphosphate was identified by calculating a/c ratio
259
through phosphorus L-edge spectra. The method to calculate a/c ratio and estimate chain length of
260
polyphosphate glass is explained by various authors in the literature. 13,45-52,54.
261
The P K-edge and P L-edge XANES results indicate the variation of phosphorus local co-
262
ordination in the extracted soot at different depths. At near surface (P L-edge TEY), Ca phosphates
263
are present, whereas, in the bulk (P K-edge TEY and FY) P is primarily associated with Zn or Ca 13 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 14 of 54
264
phosphates. Based on these results, it can be suggested that as soot particles get trapped in the
265
crankcase oil, they interact with reactive degradation compounds or polar additive species present
266
in the engine oil additive package. This potential interaction results in the adsorption of inorganic
267
decomposition products like Ca phosphates from detergent chemistry and Zn phosphates from
268
decomposed ZDDP on the soot structure. In addition, the presence of Zn phosphates raises the
269
possibility of direct interaction of soot particles with the phosphorus based tribofilms formed by
270
ZDDP antiwear additives. During engine operation, soot particles can get adsorbed on the engine
271
part surfaces like cylinder walls and abrades the protective antiwear Zn phosphate tribofilms
272
formed on the engine parts. Such abrasive action along the surfaces in motion results in
273
incorporation of Zn phosphates in the soot structure.
274
3.1.2 Sulfur Characterization
275
Figure 2 shows the S K-edge TEY and FY spectra compared with sulfates of zinc, calcium and
276
iron and sulfides of zinc and iron. Sulfur has rich K-edge spectra with sharp linewidths and
277
chemical shift range (of nearly 14 eV) over its range of oxidation states of -2 to +6 54. The Sulfates
278
(+6) have absorption peak at higher energy clearly distinguishing them from sulfides (-2). ZnS has
279
characteristic low energy strong peak at 2470 eV while FeS has absorption edge at 2482 eV. FeSO4
280
has higher energy peak at 2482.3 eV which differentiate it from zinc and calcium sulfate. Zinc
281
sulfate and calcium sulfate have quite similar main absorption peak at ~2481 eV, however, a post
282
edge at 2484.8 eV of CaSO4 mark it off from ZnSO4.
283
Upon comparing S K-edge TEY and FY spectra of soot with reference compounds, main
284
absorption peak of soot matches with the characteristic peak of ZnSO4 and CaSO4. But the
285
possibility of CaSO4 in soot structure is nullified because of the absence of post-edge peak of
286
CaSO4 at 2484.8 eV. It is important to note that pre-edge and main absorption peak in TEY soot 14 ACS Paragon Plus Environment
Page 15 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
287
spectra slightly matches with the characteristic peaks of FeS, however, FY spectra of FeS has
288
distinctive features indicating the absence of FeS in the bulk of soot. Characteristic peak a in both
289
TEY and FY clearly indicates presence of ZnS in near-surface and bulk of soot. S K-edge spectra
290
can also be used to quantify the relative proportion of sulfur compounds in diesel soot. This
291
quantification was carried out by deconvoluting the sulfur K edge spectra in TEY mode. Analysis
292
of relative amount of sulfide and sulfate compound was done by calculating height ratio of sulfide
293
to sulfate peak. The sulfide/sulfate ratio for diesel soot is 0.37.
294
The presence of Zn sulfur compounds strongly indicates the interaction of the soot with the
295
antiwear tribofilms formed by ZDDP on the engine parts. In addition, the presence of CaSO4 and
296
FeS indicates the adhesion of decomposed Ca-based detergent compounds and wear debris with
297
the engine soot carried into the crankcase oil.
15 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 16 of 54
298 299
Figure 2. Sulfur K-edge (a) TEY spectra (b) FY spectra of crankcase soot and model compounds.
300
3.1.3 Calcium Characterization
301
The Ca K-edge TEY spectra for the extracted soot is presented in the figure 3. Spectra of soot is
302
compared with model compounds of CaSO4, Ca3(PO4)2, CaCO3 and CaO.
303
For the Ca K-edge, the pre-edge peak around 4042 eV is attributed to transition from 1s to bound
304
3d orbitals. The main absorption peak region consists of three peak structure, a shoulder on the
305
lower side (around 4047 eV) resulted from 1s to 4s quadrupole transition and the principal double
306
peaks attributed to 1s to 4p1/2 and 1s to 4p3/2 dipole transition. As the occurrence of dipole transition
307
is likelier than quadrupole, the K-edge spectra is primarily dominated by high intensity main peak
308
arising due 1s to 4p transition 55. CaO has a distinctive pre-edge at 4043 eV and strong peak at 16 ACS Paragon Plus Environment
Page 17 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
309
4049.8 eV. CaSO4 and Ca3(PO4)2 has an intense absorption edge at 4050 eV and 4049.4 eV
310
respectively. The CaSO4 has more fine structure compared to Ca3(PO4)2.
311 312
Figure 3. Calcium K-edge TEY spectra of crankcase soot and model compounds.
313
The main absorption peak of soot around 4049 eV is clearly matching with the characteristic edges
314
of CaSO4, Ca3(PO4)2 and CaO, but, the pre-edge (4043 eV) and pre-shoulder (4047 eV) peak of
315
CaO is not aligned with the absorption peak of soot. These results indicate that Ca is primarily
316
associated with Ca sulfate and phosphate compounds which corroborates the earlier observations
317
of the P, S L-edge and P, S K-edge results.
318
Additionally, P L-edge FY, P K-edge TEY, S L-edge TEY and FY and Zn L-edge TEY and FY
319
spectra were also recorded to gain detailed chemical composition analysis. These spectra can be
17 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 18 of 54
320
found in the supporting information section of this paper. Table 2 summarizes chemistry of the
321
soot identified through XANES detection modes.
Table 2. Overview of different compounds identified via XANES Chemical Compounds
XANES Measuring Modes PL TEY
PL FY
P K P K S L S L S K S K Zn L TEY FY TEY FY TEY FY TEY N/A N/A N/A N/A N/A
Zn L FY N/A
Ca K TEY
Ca K FY
Ca3(PO4)2
Zn3(PO4)2
-
-
N/A
N/A
N/A
N/A
N/A
N/A
ZnSO4
N/A
N/A
N/A
N/A
CaSO4
N/A
N/A
N/A N/A
N/A
N/A
N/A
N/A
FeS
N/A
N/A
N/A
N/A
FeSO4
N/A
N/A
N/A
N/A
ZnS
N/A
N/A
N/A
N/A
-
-
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
-
-
N/A
N/A
CaCO3
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
CaO
N/A
N/A
N/A
N/A
N/A
N/A
322 323
3.2 High Resolution Transmission Electron Microscopy (HR-TEM)
324
HR-TEM was employed to examine structure and morphology of soot particles. Figure 4 is a bright
325
field HR-TEM image illustrating a typical carbonaceous turbostratic structure of primary soot
326
particle comprised of both crystalline and amorphous regions.
327
A crystalline domain of primary soot particle can be identified in the figure 4 with several layers
328
stacked a top of each other in a circular fashion. Such a closed packed array suggests that the
329
crystalline domain has a turbostratic configuration. It has been reported in the literature that a 18 ACS Paragon Plus Environment
Page 19 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
330
turbostratic arrangement comprised of 4-5 graphene layers called platelets stacked in the circular
331
manner to form crystallite with interlayer distance of about 3.5 Å 12,56,57. Marked lattice fringe area
332
on the image can be identified as the nanocrystalline particles embedded in turbostratic structure
333
of soot. The size of these particles is in the range of 3-6 nm.
334 335
Figure 4. High-resolution bright field transmission electron micrograph of crankcase soot
336
showing nanocrystalline particles, turbostratic and amorphous regions.
337
Careful observation of the HR-TEM image of the soot imply that the crystalline region is present
338
on the outer shell of the primary soot particle and is tightly anchored. Nanocrystalline particles are
339
continuous with crystalline domain of a turbostratic soot structure. These observations indicate the
340
possibility of mechanical embedding or chemical bonding of nanocrystalline particles on surface
341
of diesel soot primary particles which cannot be removed by washing, ultra-sonication and
342
centrifuging the soot. Table 3. Interplanar spacing of the marked nanocrystalline particles. 19 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Marked
Page 20 of 54
Interplanar Lattice
Interplanar Lattice
Nanocrystalline
Spacing d (A˚)
Spacing d (A˚)
Particles
Standard Value
Measured Value
BD
2.81
2.823
BE
2.778/2.811
2.77
BB
2.778/2.811
2.789
BA
2.81
2.892
Ca5(PO4)3OH (112)
BC
2.81
2.852
Ca5(PO4)3OH (112)
BG
2.77
2.74
Ca5(PO4)3OH (112)
Miller indices of crystallographic plane (hkl) Ca5(PO4)3OH (112) Ca5(PO4)3OH (112)/ Ca5(PO4,CO3)3OH (211) Ca5(PO4)3OH (112)/ Ca5(PO4,CO3)3OH (211)
343 344
Careful observation of the HR-TEM image of the soot indicates that the crystalline region is
345
present on the outer shell of the primary soot particle and is tightly anchored. The crystalline phase
346
of nanocrystalline particles is continuous with crystalline domain of turbostratic soot structure.
347
These observations indicate the possibility of mechanical embedding or chemical bonding of
348
nanocrystalline particle on surface of diesel soot primary particles which are incapable of removal
349
by washing, ultra-sonication and centrifuging the soot.
350
To correlate XANES results with HR-TEM and gain detailed insight of embedded compounds,
351
interplanar spacing of nanocrystalline region is measured and detailed in table 3. Calculated values
352
were compared with similar d-spacing of compounds possible from XANES results. Based on the
353
XANES findings, it was speculated to detect various compounds such as ZnS, ZnSO4, Zn3(PO4)2,
354
CaO, CaSO4, Ca3(PO4)2, FeS, Ca(OH)2, but no apparent match was found. Matching of interplanar
355
spacing indicates the presence of only Ca based compounds, hydroxyapatite (Ca5(PO4)3OH) and
356
carbonate hydroxyl apatite (Ca5(PO4,CO3)3OH) in the soot structure. It appears that Zn phosphates,
357
sulfates or sulfides are present in amorphous form while crystalline particles are dominated by Ca 20 ACS Paragon Plus Environment
Page 21 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
358
based compounds. The origin of Ca compounds is from the detergents such as alkyl benzene
359
sulfonates wherein Ca occurs as CaCO3 used in engine oil formulations. These detergents are used
360
to suspend insoluble polar debris and neutralize acidic byproducts of combustion. Also, they
361
contribute to protective tribofilms formed on engine parts. Many researchers have studied
362
influence of Ca detergent chemistry on tribofilm formation and have reported the formation of
363
small chain calcium phosphate films on the rubbing surfaces
364
bonded or mechanically incorporated calcium phosphates in the turbostratic soot structure
365
indicates either possibility of cross-interaction of soot particles with detergent decomposition
366
products in engine oil or removal of Ca based tribofilms by trapped soot during engine operation.
367
One of the mechanism responsible for embedment of these nanocrystalline particles can be three
368
body wear which is a proposed wear model in several studies 36,37,62,63,93. It can be postulated that
369
soot particles/ agglomerates in crankcase film can get trapped between two surfaces in motion and
370
can experience three body wear condition at extreme pressure and temperature, leading to
371
incorporation of nanocrystalline particles originating from calcium detergent chemistry in soot
372
structure. These nanocrystalline particles had been reported to have higher hardness 13,51,58. Thus,
373
presence of such mechanically embedded hard nanocrystalline particles in soot structure will make
374
it more abrasive in nature and might promote soot induced abrasive engine wear.
375
3.3 Oxidation Characteristics of Soot (HT-XRD, BET surface analysis)
376
Several studies have investigated oxidation behavior of soot extracted from flame or engine
377
exhaust stream using thermogravimetric analysis; however less importance has been given to
378
understanding oxidation characteristics of soot extracted from engine oils 64-68. This research work
379
focuses on evaluating oxidation behavior of soot sampled from engine oil using a temperature
380
resolved X-ray diffractometry.
53, 59-61.
Detection of chemically
21 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 22 of 54
381
3.3.1 High Temperature X-ray Diffractometry (HT-XRD)
382
HT-XRD is crucial in analyzing phase transformations, crystal structure, incombustible residue
383
content and oxidative stability of materials with an increase in temperature. HT-XRD experimental
384
analysis of engine soot helps to assess the influence of physical-chemical properties of turbostratic
385
soot on its oxidation reactivity. Figure 5 and 6 displays the XRD spectra for soot and carbon black
386
respectively. Both figures are divided into two sections, left graphs represent the XRD data
387
recorded for the entire scan range of 10˚-90˚ 2θ and right graphs shows “d002 range” of 18˚-30˚ 2θ.
388
The peak originating at ~25˚ 2θ represents a characteristic peak arising from the basal plane (002)
389
of the disordered graphitic lattice in the soot structure 69,70. Therefore, the data in the range of 18˚-
390
30˚ was separately plotted for easier visualization of this peak on the figures. Spectra marked in
391
the blue color shows the XRD data acquired before complete oxidation of carbonaceous material,
392
while the spectra in orange color is from the oxidized sample. Even though XRD scans were
393
recorded at intervals of 50˚C increase in temperature starting from 50˚C to 700˚C, the data was
394
plotted for only those temperatures where changes in lattice spacing, phase transition or peaks
395
from crystalline species are noticeable. For this reason, XRD scans for carbon black are only
396
shown at 50ºC and 700ºC in figure 6.
22 ACS Paragon Plus Environment
Page 23 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
397 398
Figure 5. Variations in crankcase soot HT-XRD spectra with temperature for 10˚-90˚ 2θ scan
399
range (left) and 18˚-30˚ 2θ d002 scan range (right).
400 401
Figure 6. Variations in carbon black HT-XRD spectra with temperature for 10˚-90˚ 2θ scan range
402
(left) and 18˚-30˚ 2θ d002 scan range (right).
23 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 24 of 54
403
The data in the range of 18˚-30˚ provides information on diffraction peaks originating from
404
nanocrystallites embedded in the soot turbostratic structure. Result shows the changes in full width
405
half maxima, peak intensity and peak shifts with an increase in temperature. It is observed that the
406
d002 peak diminishes beyond 550˚C. A careful examination of d002 range of 18˚-30˚C 2θ in figure
407
5 indicates that with increase in temperature, intensity maxima (represented by red dotted lines) of
408
diffraction peaks gradually shifts towards lower 2θ values signifying increase in interplanar lattice
409
spacing. Expansion of lattice in [002] direction eases access of oxygen to soot core structure
410
helping soot particles to oxidize at relatively lower temperature.
411
Comparison of d002 range (18˚-30˚C 2θ) of soot and carbon black shows absence of any diffraction
412
peak originating from carbon black sample after oxidation. However, multiple diffraction peaks
413
are observed for soot sample at 700ºC possibly originating from polycrystalline material (i.e.
414
incombustible residue in the current sample) which may be adhered or embedded in the
415
carbonaceous turbostratic soot structure. These results imply that interaction of soot particles with
416
decomposition products of oil additives (metallic species from detergents, antiwear components
417
etc.) and wear debris in the engine oil, at extreme pressure and temperature conditions create higher
418
disorder in the soot graphitic structure, thereby leading to increase in oxidative reactivity. In
419
addition, the adsorbed wear debris in the form of Fe, Al oxides can catalyze the oxidation of
420
carbonaceous soot at higher temperature (above 300˚C) 71,72. Most recently, Bagi et al. 4,5,21,73 and
421
co-workers examined correlation between the oxidation behavior and chemical composition of
422
engine oil soot with changes in accumulated vehicle mileage. They proposed that with the engine
423
age, there was increase in the amount of wear debris, tribofilm compounds and engine oil species
424
embedded in the soot structure which would influence ease of oxidation of soot.
24 ACS Paragon Plus Environment
Page 25 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
425
Mapping changes in interplanar lattice spacing (d002) with temperature provides significant
426
information about degree of disorder, number of graphene layers per crystallite, crystallite height,
427
width, which can be correlated to fuel composition, engine operating conditions, combustion
428
mechanism, ease of oxidation, etc. Figure 7 shows variation in interplanar spacing (d002) with
429
increase in temperature for soot and carbon black.
430 431
Figure 7. Mapping variations in interplanar spacing (d002) with increase in temperature for
432
crankcase soot and carbon black.
433
To calculate d (002) values, baseline corrections and intensity maxima were applied to diffraction
434
spectra and then Bragg’s law was used. Interplanar lattice spacing increases in soot and carbon
435
black samples with increase in temperature. In both cases, carbonaceous soot structure remains
436
intact till 250 ˚C, beyond which the d (002) planes expands in the case of carbon black but remains
437
constant in the case of soot. The turbostratic structure of soot falls apart after 550˚C and that of
438
carbon black after 600˚C. Oxidation reaction rate can be attributed to induced higher disorder of 25 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 26 of 54
439
carbonaceous soot structure featuring increased number of graphene edges on the surfaces i.e. a
440
greater number of active sites for oxygen attack
441
changed soot turbostratic structure (HR-TEM results) and chemical composition (XANES results)
442
on soot oxidation characteristics.
443
3.3.2 Analysis of Incombustible Residue
444
Th residue was collected for both soot and carbon black sample after completion of their HT-XRD
445
studies at 700ºC. This residue has a yellowish to gray appearance as shown in figure 8. Quantitative
446
analysis of remaining residue was carried out by normalizing them as a percentage of their initial
447
weights and by using following formula.
71,93.
HT-XRD analysis ascribe the influence of
448
Residue (wt %) = {(Wsoot − Wresidue)/Wsoot} × (100)
449
“Wsoot” represents the weight of the soot sample before oxidation while “Wresidue” corresponds to
450
the weight of residue left behind after oxidation; they were measured on an electronic balance with
451
precision scale in milligrams. Soot and carbon black sample showed 64 wt% and 42 wt% of the
452
residue left at 700ºC respectively. The higher amount of incombustibles remained after combustion
453
of soot also supports postulate of possible interaction of crankcase soot particles with
454
wear/corrosion byproducts, decomposed additive compounds and polar species in engine oil,
455
leading to adsorption of inorganic compounds on soot particles. On the other hand, carbon black
456
experienced no interaction with engine oil and hence had less amount of residue.
457
Energy dispersive spectroscopy (EDS) was used to analyze the elemental composition of the
458
residue; EDS uses the ZAF method for quantification of elements in the sample. The acronym
459
'ZAF' describes a procedure in which corrections for atomic number effects (Z), absorption (A)
460
and fluorescence (F) are calculated separately. The atomic number (Z), absorption coefficient (A)
461
for every element is calculated. Figure 8 shows the EDS spectra for residue acquired from
26 ACS Paragon Plus Environment
Page 27 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
462
crankcase soot. Elements such as Zn, P, and S originate from antiwear additives, whereas Fe
463
originates from the wear of engine components. In addition, S can come from the fuel and
464
detergents like sulfonates and phenates. As mentioned before, source of Ca is from the calcium
465
sulfonate based detergents. To compliment EDS results and gain detailed knowledge on actual
466
composition and structure, XRD spectra of residue were acquired for soot and carbon black.
467
Figure 9 displays XRD spectra obtained from crankcase soot and carbon black samples after being
468
heated up to 700˚C. The residue left behind after oxidation of soot shows the presence of
469
polycrystalline material whereas carbon black spectra does not show any diffraction peaks. Even
470
though 42wt% of residue was left after oxidation of carbon black, compounds in the residue are
471
likely amorphous in nature, and hence were not detected in XRD spectra. Diffraction peak
472
positions and Miller indices are marked on the XRD spectra of crankcase soot along with the
473
reference XRD spectra of CaSO4 compound. Identification of possible polycrystalline compounds
474
were carried out using the search-match function in the ICDD database and majority of the peaks
475
were matched with CaSO4 spectra as shown in figure 9. CaSO4 in the diesel engine soot originates
476
from the interaction of soot with detergents dispersed in crankcase oil. EDS spectra of crankcase
477
soot (figure 8) shows strong peaks of Zn, P, S elements, however, these elements were not
478
identified in XRD spectra, thereby indicating the presence of amorphous zinc phosphates or
479
sulfates compounds which are well known to be a part of ZDDP based antiwear tribofilms
480
Presence of these compounds supports our analysis made from XANES of interaction of soot
481
particles with the protective antiwear tribofilms during three body wear condition where crankcase
482
oil soot gets trapped between two surfaces in motion and results in extreme wear and tear of
483
contacting surfaces 57. Thus, the results suggest the detrimental effect of interaction of soot with
78-80.
27 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 28 of 54
484
components of lubricant additive package on soot’s oxidation characteristics and the amount of
485
incombustible residue left behind.
486 487
Figure 8. EDS spectra of residue left behind after oxidation of crankcase soot.
488
489 490
Figure 9. XRD spectra of oxidized soot and phase identification 28 ACS Paragon Plus Environment
Page 29 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
491
3.3.3 Surface Area Analysis
492
The surface area of particles within the material can greatly influence its surface stability and
493
performance characteristics
494
determined by the BET nitrogen physisorption method described in the experimental section (2.2).
495
The BET physisorption studies were subjected to carbon black for baseline comparisons as it is
496
considered to be surrogate of diesel engine soot 35, having similar structure and morphology, but
497
different composition than diesel engine soot. The measured surface area of crankcase soot is 69
498
m2/g while carbon black exhibited higher surface area of 124 m2/g. Coupled with agglomeration,
499
the adsorption of lubricant additive compounds, decomposition products, wear debris and polar
500
impurities from engine oil on the soot particles results in lower surface area for crankcase soot.
501
The carbon black samples displayed higher surface area as they did not experience any interactions
502
with engine oil components. Patel et al. 13 also used the BET nitrogen adsorption method to study
503
the differences in surface area of the carbon black and diesel engine soot. Owing to similar primary
504
particle sizes between diesel soot and carbon black, it was proposed that size differences between
505
the two samples were not responsible for large variation in measured surface area. A previous
506
study by Bagi et al.
507
adhesion of wear debris and engine oil inorganic species with accumulated vehicle mileage. BET
508
results can be correlated with HT-XRD oxidation studies. It can be seen that the amount of residue
509
left after oxidation of soot and carbon black are inversely proportional to the available surface area
510
determined by BET analysis.
511
3.3.4 Oxidation Modes
512
The soot reactivity towards oxidation is strongly dominated by soot nanostructure and thus, the
513
data acquired through HR-TEM and HT-XRD should be used simultaneously to examine oxidation
21
25,64,81.
The surface area of the extracted crankcase soot was
reported a lower surface area of diesel engine soot due to increase in the
29 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 30 of 54
514
modes of soot. The primary particle of soot exhibits an inner core and an outer shell structure,
515
comprised of crystalline and amorphous domain 50,83-85. The HR-TEM study (Section 3.2) indicates
516
the presence of Ca based nano-crystalline particles embedded in the outer periphery (crystalline
517
domain) of the soot primary particle. Embedment of such particles in the outer shell possibly result
518
in smaller fringe length (measure of the physical extent of carbon layer planes)
519
fringe separation (mean distance between adjacent carbon layer planes, also known as d(002)) 70 of
520
soot crystallite domain. Smaller fringe length indicates smaller graphene layer dimensions and
521
more edge-site carbon atoms i.e. higher degree of carbon structural disorder
522
present at edge site have greater possibility to interact with oxygen than those located in basal
523
plane (002)
524
enhancing oxidation of the primary soot particle 88. The higher surface energy and lower activation
525
energy available due to structural disorder in the outer shell of the soot particle may possibly lead
526
to higher reactivity of outer shell. In addition, by referring to d(002) variations in figure 10, it can be
527
proposed that oxidation of primary soot particle initially starts with the slow external burning
528
(outside-in) of an outer shell till 550˚C. The inner core is considered to be highly reactive 70,71,89,
529
initial stage of combustion is followed by fast internal burning (inside-out) of inner core between
530
550˚C-700˚C. The HT-XRD results suggests likelihood of primary soot particles to follow “Dual
531
oxidation mode” of combustion. However, data from HR-TEM and BET obtained at different
532
temperatures need to be correlated with HT-XRD data to have a concrete understanding of
533
oxidation modes. In addition to the structural disorder factor, the possibility of adsorption of
534
catalytic metallic compounds on soot particles (that may catalyze soot oxidation) also advocates
535
the higher reactivity of outer shell and dual mode of primary soot particle oxidation.
536
3.4 Evaluation of Tribological Test Wear Results
87,89.
86,87
81,88.
and larger
Carbon atoms
Additionally, larger fringe separation can ease oxygen access to carbon layers
30 ACS Paragon Plus Environment
Page 31 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
537
XANES, HR-TEM, HT-XRD and BET studies confirm the interaction of trapped soot
538
particles/agglomerates with the protective tribofilms and decomposition products of P and S based
539
lubricant antiwear additives and Ca sulfonate based detergents; wear debris and combustion
540
byproducts in the engine oil used for MackT12 dynamometer test. Additionally, results indicate
541
the potential of crankcase soot to promote severe wear of engine components through wear
542
conditions like three body wear. To get comprehensive understanding of soot induced wear
543
mechanisms, four ball tribological bench tests were performed. Test oil formulations with different
544
concentrations of soot were prepared to examine the influence of various soot loading conditions
545
on wear mechanism, whereas, various antiwear additives were used to elucidate the effect of soot-
546
antiwear additive interaction on the wear rate of tribological contact. WSD was obtained for three
547
stationary balls using the method described in the section 2.3. Figure 10 is a comparative bar chart
548
detailing the WSD results for test oil formulations having three different antiwear additives and
549
soot in increment amount. Result for base oil was added for baseline comparison. WSD values are
550
reported in µm with error bars representing corresponding variation in the scar diameter values for
551
each test. In comparison with base oil, all three additives showed improvement in wear
552
performance of oil formulations.
553
ZDDP exhibited best antiwear performance with 618 µm WSD, while, DEDTP and DDP had
554
slightly higher wear losses. Addition of 2 wt% soot resulted in decrement of wear performance of
555
all oil formulations. Changes in antiwear behavior of the respective additives with an increase in
556
soot concentration from 2 wt% to 10 wt% are of prominent importance. ZDDP oil formulations
557
showed a stable WSD with increase in soot levels from 2 wt% to 5 wt%, however, a sudden
558
increase in WSD is observed at 10 wt% soot. DEDTP also displayed gradual decrease in wear
559
performance with a significant increase in WSD value by 870 µm from 0 to 10 wt% soot. In the 31 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 32 of 54
560
case of DDP formulations, notable rise in WSD is observed at 5 wt% as well as at of wt% soot,
561
unlike DEDTP and ZDDP formulations. These observations indicate, compared to ZDDP, both
562
DEDTP and DDP exhibit better wear performance at a higher soot concentration of 10 wt%.
563 564
Figure 10. Effect of increase in soot concentration on wear scar diameter of all oil formulations.
565
3.5 Analysis of Rubbed Surfaces (SEM-EDS)
566
Topographical and elemental composition analysis of worn surfaces of tribological test samples
567
were carried out using SEM-EDS. Figure 11 to 13 show low magnification (70X) and high
568
magnification (1000X) images of wear scars along with the corresponding surface elemental
569
composition spectra of wear scars acquired from test oil formulations containing different antiwear
570
additives and no soot. The boxes marked on low magnification images are the representative of
32 ACS Paragon Plus Environment
Page 33 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
571
the area under consideration, which is further magnified to 1000X and subjected to EDS spectrum
572
analysis.
573
SEM images at 70X show scratches in the direction of sliding which occurred during initial stage
574
of test, however, at higher magnification it is clearly seen that scratches are covered with tribofilm
575
patches (darker regions in the SEM images represent the presence of phosphate tribofilms). SEM
576
images demonstrate that the rubbed surfaces generated using ZDDP lubrication have wider
577
coverage of tribofilms than DEDTP and DDP, hence indicating better wear performance of ZDDP
578
(i.e. low WSD value) as compared to DEDTP and DDP. The corresponding EDS spectra
579
determines the type and amount of elements present on the worn surfaces. ZDDP is well known to
580
form antiwear tribofilms consisting of zinc phosphates, zinc sulfides/sulfates, iron phosphate and
581
iron sulfides/sulfates 83,84.
582
EDS spectra of ZDDP shown in figure 11.(A) confirms the presence of Zn, P, S, O, and Fe
583
elements. Zhang et al.
584
XANES and found that predominant cations were of Zn as compared to Fe. However, in this study
585
use of EDS can’t help to differentiate between Zn and Fe abundance. As the penetration depth of
586
the electron beam is deeper than the thickness of the tribofilm on the surface, the majority amount
587
of Fe as seen in the figure may have been originated from the substrate. The identification of P, S,
588
O, and Fe elements on the worn surfaces developed by DEDTP and DDP oil formulation indicates
589
that these tribofilms are comprised of iron phosphates and iron sulfides/sulfates. This is in
590
accordance with the study by Sharma et al. 43,54 and Kim et al. 47, who reported that phosphorus
591
and sulfur based ionic liquid and ashless dithiophosphate additives form pad like tribofilms very
592
similar to ZDDP with only one difference of presence of compounds coordinated by Fe cations
593
instead of Zn cations.
45,80
examined the low load and friction tribofilm formed by ZDDP using
33 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 34 of 54
594
34 ACS Paragon Plus Environment
Page 35 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
595
Figure 11. SEM-EDS results for test formulations having ZDDP as antiwear additive (A) ZDDP
596
(B) ZDDP_D_Soot_2wt% (C) ZDDP_D_Soot_5wt% (D) ZDDP_D_Soot_10wt%.
597
Figure 11.(B), 12.(B) and 13.(B) represents the SEM-EDS results for the oil formulations
598
containing 2 wt% soot. Abrasive wear along with some polishing wear tracks and smooth patches
599
of tribofilms can be observed in the direction of sliding for the rubbed surfaces generated by all
600
oil formulations containing 2wt% soot. Higher magnification SEM image of wear scar created
601
through ZDDP lubrication with 5wt% (Figure 11.C) soot display furrowed wear tracks with fine
602
cracks indicating significant amount of abrasive wear, while that of DEDTP at 5 wt% (Figure
603
12.C) soot show combination of adhesive and abrasive wear with some deep cuts. Worn surface
604
lubricated by DDP formulation having 5 wt% soot (Figure 13.C) has pitting and abrasive tracks,
605
indicating the severe wear of the substrate through corrosive aided abrasive mechanism. Various
606
studies in the literature have well-documented the combination of corrosive and abrasive processes
607
leading to high rates of wear 33,34,90, which justifies a significant increase in WSD (Figure 10) for
608
DDP formulations at 5 wt% as opposed to ZDDP and DEDTP at 5 wt%. The corresponding EDS
609
spectra shown in figure 12.(C) and 13.(C) show low intensity peak of tribofilm forming elements
610
like P and S which suggests the depletion of antiwear phosphate and sulfides tribofilms on rubbed
611
surfaces. Figure 11.(D), 12.(D) and 13.(D) corresponds to the SEM-EDS results of the wear
612
surfaces developed during tribological testing of formulations containing antiwear additives,
613
dispersant and 10 wt% soot. Representative SEM images of rubbed surfaces display larger wear
614
scars with deep ploughing grooves in the direction of sliding, which is sign of severe abrasive
615
wear. In addition to abrasive wear, there are evidences of scuffing, adhesive-abrasive wear, deep
616
metal cut and corrosive-abrasive wear. These observations suggest the possibility of adverse
617
engine wear induced due to high level of soot contamination of lubricating engine oil. 35 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 36 of 54
618
36 ACS Paragon Plus Environment
Page 37 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
619
Figure 12. SEM-EDS results for test formulations having DEDTP as antiwear additive (A)
620
DEDTP (B) DEDTP_D_Soot_2wt% (C) DEDTP_D_Soot_5wt% (D) DEDTP_D_Soot_10wt%.
621 37 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 38 of 54
622
Figure 13. SEM-EDS results for test formulations having DDP as antiwear additive (A) DDP (B)
623
DDP_D_Soot_2wt% (C) DDP_D_Soot_5wt% (D) DDP_D_Soot_10wt%.
624
38 ACS Paragon Plus Environment
Page 39 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
625
Figure 14. Correlation of soot content with WSD and concentration of P and S elements present
626
on tribosurfaces generated due to lubrication with formulations containing (A) ZDDP antiwear
627
additive (B) DEDTP antiwear additive (C) DDP antiwear additive.
628
Figure 14 summarizes the correlation between decrement of major elements corresponding to
629
tribofilm compounds (P, S) and wear scar diameter (WSD) of all the test formulations under study.
630
It can be observed in the graph 14(A), with the increase in soot concentration, wear scar diameter
631
increases and ZDDP formed tribofilm elements P and S decreases. Similar trend is also seen for
632
the formulations containing DEDTP and DDP as antiwear additives (figure 14(B) and 14(C)).
633
Therefore, the correlation between WSD and elements corresponding to antiwear tribofilms
634
indicates that the soot particles promotes high wear of surfaces in contact through removal of the
635
protective tribofilms formed by antiwear additives present in lubricating oil.
636
3.6 Antagonistic Interaction of Soot and Antiwear Additives
637
From the above discussion, negative impact of soot particles on the wear of tribological contacts
638
operating under boundary lubrication regime is evident. To understand the effect of antiwear
639
additives soot interaction on the wear of steel substrate, tests were conducted using base oil
640
containing only soot and dispersant.
641
Figure 15 shows comparison of wear results obtained with formulations containing only soot and
642
no antiwear additives to that of formulations having anti-wear additives and soot. Dispersant
643
concentration was kept constant at 5 wt% so as to directly correlate influence of presence of soot
644
and antiwear additives in oil with the wear outcomes.
645
For lubrication containing only soot and no antiwear additives, WSD increases from 756 μm to
646
776 μm as the soot concentration increases from 2 wt% to 5 wt%. In both cases, it can be speculated
647
that soot is acting as the third body abrasive particles generating local wear to the substrate through 39 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 40 of 54
648
sliding or rolling motion at the interface. For D_Soot_10 wt% sudden increase in the WSD is
649
observed. This can be explained with the loss in suspension of soot particles, wherein amount of
650
dispersant is not sufficient enough to suspend high soot concentration of 10 wt%. In such systems,
651
secondary or primary soot particles can coalesce to form agglomerates, large enough in creating
652
lubrication starvation at the interface and inducing direct two body wear.
653 654
Figure 15. Effect of soot-antiwear additive interaction on wear scar diameter.
655
In figure 15, it can be seen that all formulations containing anti-wear additives exhibits poor wear
656
performance (i.e. higher WSD values) as compared to the formulations containing only soot and
657
no anti-wear additives. Considering statistical significance, it is safe to say that this behavior is 40 ACS Paragon Plus Environment
Page 41 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
658
constant at all soot concentrations of 2, 5 and 10 wt%. There are two possible inferences from this
659
observation, ⅰ) soot particles are directly abrading off the formed tribofilms by mechanical action
660
or ⅱ) soot particles are disrupting the formation of tribofilms after initial removal by adsorbing
661
film forming species or accumulating on active steel substrate. This study clearly suggests the
662
prevailing wear mechanism as the direct removal of protective tribofilms by the abrasive soot
663
particles. Salehi et al. 33 also suggested corrosive-abrasive wear mechanism is responsible for high
664
wear observed in fully formulated oils containing carbon black as surrogate for diesel engine soot.
665
In their study, corrosive-abrasive wear was attributed to the removal of protective tribofilms
666
formed on steel substrate through abrasion by carbon black particles.
667
Generation of tribofilms in boundary lubrication regime is a complicated process. In our previous
668
work on ZDDP tribofilms, we have detailed two step tribofilm formation mechanism 78,79. Initially,
669
films (generally iron free) are formed on the steel surface due to deposition/ adsorption of
670
tribochemical reaction products produced by degradation /thermo-oxidative decomposition of
671
additive chemistries. During continued friction and wear process, this film can get disrupted and
672
the active elements (like P, S) present in lubricant chemically reacts with the nascent surface (Fe)
673
generating iron phosphates, iron sulfides/sulfates. These compounds are incorporated in the top
674
layer and are strongly bonded to the nascent steel surface; thus, a sacrificial protective surface film
675
is formed which is commonly known as a tribofilm. Tribofilms are reported to have less hardness
676
than the steel substrate 34,79,91 and hence, can get easily removed by abrasive soot particles. A new
677
tribofilm forms on the newly generated surface and is then again removed by soot particles. Rapid
678
removal of antiwear films can overcome the rate of film formation by antiwear additives and
679
results in severe wear of steel substrate. Therefore, the rate of wear can be said to be indirectly
680
dependent on the availability of a film formed by antiwear additives. 41 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 42 of 54
681
The proposed corrosive-abrasive wear mechanism induced due to presence of anti-wear additives
682
and soot particles is also explained with the help of phenomenological model shown in figure 16
683
(B), where soot particles are shown to create deep abrasive grooves (marked by red line) by
684
removal of tribofilms (marked by yellow line). Figure 16(A) shows the three body soot induced
685
wear taking place when soot particles and dispersant based lubrication is present at the interface.
686 687
Figure 16. Schematic diagram explaining soot induced wear mechanism with lubrication
688
containing (A) dispersant and soot; (B) anti-wear additives, dispersant and soot.
689
4. Conclusion
690
The variations in the carbonaceous structure and chemical composition of crankcase soot were
691
mapped using XANES and HR-TEM. XANES analysis demonstrated the presence of zinc
692
polyphosphate, calcium sulfate, calcium phosphate, and iron sulfate and phosphate. Identification
693
of short chain zinc polyphosphate suggested the interaction of soot with the antiwear tribofilm. 42 ACS Paragon Plus Environment
Page 43 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
694
HR-TEM results showed the nanocrystalline particles of hydroxyapatite and carbonate hydroxyl
695
apatite embedded in the periphery of the turbostratic carbon structure of the soot. It was concluded
696
that the mechanical embedment of the hard nanocrystalline particles during the three-body contact
697
of soot particles in the crankcase oil altered its structure and could make it more abrasive in nature.
698
In this study, a different approach using high temperature XRD was used to depict and correlate
699
the oxidation stability with interplanar lattice spacing changes of soot. Results from XANES, HR-
700
TEM, HT-XRD, BET techniques and EDS analysis of residue (left behind after oxidation)
701
indicated that interaction of lubricant additives and crankcase soot extracted from Mack-T12
702
dynamometer engine test influences its structure and chemistry, which in turn affects its ease of
703
oxidation.
704
This research work also focused on understanding the detrimental effect of increased levels of soot
705
in lubricating oil on antiwear properties of additives and elucidating the mechanism by which
706
abrasive soot promotes high wear in boundary lubrication conditions. The proposed dominant wear
707
mechanism is the abrasion wear due to rapid removal of the protective antiwear tribofilms by the
708
abrasive soot particles/agglomerates. ZDDP exhibited the worst performance at higher level of
709
soot while ionic liquid, DEDTP exhibited better anti-wear performance. This indicates that ionic
710
liquids may serve as a secondary antiwear protection for fully formulated diesel engine lubricants.
711 712
Acknowledgments
713
The authors appreciate Dr. Ewa Bardasz assistance in providing drained oil from a Mack T-12
714
engine test. We also appreciate Chevron, BASF, Sigma Aldrich and AC2T research GmbH
715
(Austrian COMET-Program K2 XTribology, project no. 849109) for providing with required
716
dispersants and anti-wear additives for this study. The XANES experiments were conducted at the 43 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 44 of 54
717
Canadian Light Source, which is supported by NSERC, NRC, CIHR, and the University of
718
Saskatchewan. Experimental facilities for HR-TEM, SEM and EDS provided by the Center of
719
characterization for materials and biology (CCMB), University of Texas at Arlington are gratefully
720
acknowledged.
721
Supporting Information
722
XANES P L-edge FY, P K-edge TEY, S L-edge TEY and FY and Zn L-edge TEY and FY spectra
723
are supplied as the supporting information. Additionally, EDS mapping results exhibiting removal
724
of antiwear tribofilms due to soot induced wear are available for reference.
725
References
726
[1] Xi J, Zhong BJ. Soot in Diesel Combustion Systems. Chem Eng Technol 2006;29:665-73.
727
[2] Daido S, Kodama Y, Inohara T, Ohyama N, Sugiyama T. Analysis of soot accumulation
728
inside diesel engines. JSAE Review 2000;21:303-8.
729
[3] Green D.A, Lewis R, Dwyer-Joyce R.S. Wear effects and mechanisms of soot-
730
contaminated automotive lubricants. Proceedings of the Institution of Mechanical
731
Engineers, Part J: Journal of Engineering Tribology 2006;220(3):159-69.
732
[4] Bagi S, Bowker R, Andrew R. Understanding chemical composition and phase transitions
733
of ash from field returned DPF units and their correlation with filter operating conditions.
734
SAE International Journal of Fuels and Lubricants 2016;9(1):239-59.
735
[5] Bagi S, Singh N, Andrew R. Investigation into Ash from Field Returned DPF Units:
736
Composition, Distribution, Cleaning Ability and DPF Performance Recovery. No. 2016-
737
01-0928. SAE Technical Paper, 2016.
738 739
[6] Chiñas-Castillo F, Spikes H. The behavior of diluted sooted oils in lubricated contacts. Tribol Lett 2004;16(4):317-22. 44 ACS Paragon Plus Environment
Page 45 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
740 741
Energy & Fuels
[7] Sharma V, Bagi S, Patel M, Aderniran O, Aswath P. B. Structure and chemistry of soot and its role in wear of diesel engines. Tribology Online 2016;11(5):551-5.
742
[8] Cadman W, Johnson JH. Study of the effect of exhaust gas recirculation on engine wear in
743
a heavy-duty diesel engine using analytical ferrography. SAE Technical paper
744
1986;860378.
745 746 747 748 749 750
[9] Colacicco P, Mazuyer D. Role of soot aggregation on the lubrication of diesel engines. Tribology Transactions 1995;38:959-65. [10] Gautam M, Durbha M, Chitoor K, Jaraiedi Mea. Contributions of Soot Contaminated Oils to Wear. SAE Technical Paper 1998. [11] Green D.A, Lewis R. The effects of soot-contaminated engine oil on wear and friction: A review. Proc Inst Mech Engineering Pt D: J Automobile Engineering 2008;222:1669-89.
751
[12] Sharma V, Uy D, Gangopadhyay A, O'Neill A, Paxton W, Sammut A, Ford M, Aswath P.
752
Structure and chemistry of crankcase and exhaust soot extracted from diesel engines.
753
Carbon 2016;103:327-38.
754 755
[13] Patel M, Aswath P. Structure and chemistry of crankcase and cylinder soot and tribofilms on piston rings from a Mack T-12 dynamometer engine test. Tribol Int 2014; 77:111-21.
756
[14] Uy D, Ford MA, Jayne DT, O’Neill AE, Haack LP, Hangas J. Characterization of gasoline
757
soot and comparison to diesel soot: Morphology, chemistry, and wear. Tribol Int
758
2014;80:198-209.
759
[15] Ishiguro T, Takatori Y, Akihama K. Microstructure of diesel soot particles probed by
760
electron microscopy first observation of inner core and outer shell. Combust Flame
761
1997;108:231-34.
45 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
762 763
Page 46 of 54
[16] Lu T, Cheung CS, Huang Z. Investigation on Particulate Oxidation from a DI Diesel Engine Fueled with Three Fuels. Aerosol Sci Technol 2012;46:1349-58.
764
[17] Timothy J, Dennis A, Zoran F. The Impact of Exhaust Gas Recirculation on Performance
765
and Emissions of a Heavy-Duty Diesel Engine. Society of automotive engineers 2003.
766
[18] Ajayi O, Erck R, Erdemir A, Fenske G, Goldblatt I. Effect of exhaust gas recirculation
767
(EGR) on diesel engine oil- Impact on wear. 14th DEER Conference, MI, 2008;4-7.
768
[19] Sasaki M, Kishi Y, Hyuga T, Okazaki K, Tanaka M, Kurihara I. Effect of EGR on diesel
769
engine oil, and its countermeasures. Proceedings of the 1997 International Spring Fuels &
770
Lubricants Meeting, May 5, 1997–May 8 1997;1271:39–44.
771 772
[20] Nagai I, Endo H, Nakamura H, Yano H. Soot and valve train wear in passenger car diesel engines. SAE Technical Paper 1983;831757.
773
[21] Bagi S, Sharma V, Patel M, Aswath P. Effects of diesel soot composition and accumulated
774
vehicle mileage on soot oxidation characteristics. Energy and Fuels 2016; 30(10):8479-90.
775
[22] Song J, Alam M, Boehman A, Kim U. Examination of the Oxidation Behavior of Biodiesel
776 777 778
Soot. Combust Flame 2006;146:589-604. [23] Lu T, Cheung CS, Huang Z. Effects of Engine Operating Conditions on the Size and Nanostructure of Diesel Engine Particles. J Aerosol Sci 2012;47:27-38.
779
[24] Lu T, Huang Z, Cheung CS, Ma J. Size Distribution of EC, OC And Particle-Phase PAHs
780
Emissions from a Diesel Engine Fueled with Three Fuels. Sci Total Environ 2012;438:33-
781
41.
782 783
[25] Yehliu K, Vander Wal RL, Armas O, Boehman AL. Impact of fuel formulation on the nanostructure and reactivity of diesel soot. Combust Flame 2012;159(12):3597-606.
46 ACS Paragon Plus Environment
Page 47 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
784 785 786 787
Energy & Fuels
[26] Vander Wal RL, Tomasek AJ. Soot Oxidation: Dependence Upon Initial Nanostructure. Combust Flame 2003;134(1):1–9. [27] Vander Wal RL, Tomasek AJ. Soot nanostructure: Dependence upon synthesis conditions. Combust Flame 2004;136:129-140.
788
[28] Vander Wal RL, Mueller CJ. Initial investigation of effects of fuel oxygenation on
789
nanostructure of soot from a direct-injection diesel engine. Energ Fuels 2006;20(6):2364-
790
69.
791 792 793 794 795 796
[29] Needelman W, Madhavan P. Review of lubricant contamination and diesel engine wear. SAE 1988;881827. [30] Berbezier, I., Martin, J.M., Kapsa, Ph. The role of carbon in lubricated mild wear. Tribol Int 1986; 19:115–22. [31] Rounds FG. Soots from used diesel engine oils - Their effects on wear as measured in 4ball wear tests. SAE Technical Paper 1981;810499.
797
[32] Rounds FG. Carbon: Cause of diesel engine wear? SAE Technical Paper 1977;770829.
798
[33] Salehi F, Khaemba D, Morina A, Neville A. Corrosive–Abrasive Wear Induced by Soot in
799 800 801 802 803 804 805
Boundary Lubrication Regime. Tribol Lett 2016; 63-19. [34] Olomolehin Y, Kapadia R, Spikes H. Antagonistic Interaction of Antiwear Additives and Carbon Black. Tribol Lett 2010; 37:49-58. [35] Uy D, O'Neill A. E, Simko S. J, Gangopadhyay, A. K. Soot–additive interactions in engine oils. Lubrication Science 2010; 22:19–36. [36] George S, Balla S, Gautam M. Effect of diesel soot contaminated oil on engine wear. Wear 2007; 262:1113-22.
47 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
806 807 808 809 810 811 812 813 814 815
Page 48 of 54
[37] Gautam M, Chitoor K, Durbha M, Summers J. Effect of diesel soot contaminated oil on engine wear-investigation of novel oil formulations. Tribol Int 1999;32:687–99. [38] Colacicco P, Mazuyer D. The role of soot aggregation on the lubrication of diesel engines. Tribology Transactions 1995; 38:959–65. [39] Yoshida, K. Effects of sliding speed and temperature on tribological behavior with oils containing a polymer additive or soot. Tribology Transactions 1990; 33:221–28. [40] Aldajah S, Ajayi OO, Fenske GR, Goldblatt IL. Effect of exhaust gas recirculation (EGR) contamination of diesel engine oil on wear. Wear 2007; 263:93-8. [41] George S, Balla S, Gautam V, Gautam M. Effect of diesel soot on lubricant oil viscosity. Tribol Int 2007; 40:809–18.
816
[42] Nicholls M, Najman MN, Zhang Z, Kasrai M, Norton PR, Gilbert PUPA. The contribution
817
of XANES spectroscopy to tribology. Canadian Journal of Chemistry 2007; 85:816-30.
818
[43] Sharma V, Erdemir A, Aswath P. An analytical study of tribofilms generated by the
819
interaction of ashless antiwear additives with ZDDP using XANES and nano-indentation.
820
Tribol Int 2015; 82:43-57.
821 822 823 824
[44] Patel M, Aswath PB. Chemical Characterization of Diesel Soot using XANES. Earth and Environmental Science, Activity Report 2009. [45] Zhang Z, Yamaguchi E, Kasrai M, Bancroft G, Liu X, Fleet M. Tribofilms generated from ZDDP and DDP on steel surfaces: Part 2, chemistry. Tribology Letters 2005; 19:221-29.
825
[46] Fuller M, Kasrai M, Michael Bancroft M, Fyfe K,Tan K. Solution decomposition of zinc
826
dialkyl dithiophosphate and its effect on antiwear and thermal film formation studied by X-
827
ray absorption spectroscopy. Tribology international 1998; 31:627-44.
48 ACS Paragon Plus Environment
Page 49 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
828
[47] Kim B, Mourhatch R, Aswath PB. Properties of tribofilms formed with ashless
829
dithiophosphate and zinc dialkyl dithiophosphate under extreme pressure conditions. Wear
830
2010; 268:579-91.
831
[48] Yin Z, Kasrai M, Fuller M, Bancroft G.M, Fyfe K, Tan K.H. Application of soft X-ray
832
absorption spectroscopy in chemical characterization of antiwear films generated by ZDDP
833
part I: the effects of physical parameters. Wear 1991; 202:172-191.
834
[49] Li Y, Pereira G, Lachenwitzer A, Kasrai M, Norton PR. X-ray absorption spectroscopy
835
and morphology study on antiwear films derived from ZDDP under different sliding
836
frequencies. Tribology Letters 2007;27:245-53.
837
[50] Li Y, Pereira G, Kasrai M, Norton PR. Studies on ZDDP anti-wear films formed under
838
different conditions by XANES spectroscopy, atomic force microscopy and 31P NMR.
839
Tribol Lett 2007;28:319-28.
840
[51] Patel M, Aswath P. Morphology, structure and chemistry of extracted diesel soot: Part II:
841
X-ray absorption near edge structure (XANES) spectroscopy and high-resolution
842
transmission electron microscopy. Tribol Int 2012;52:17-28.
843
[52] Sharma
V,
Timmons
R,
Erdemir
A,
Aswath
P.
Plasma-functionalized
844
polytetrafluoroethylene nanoparticles for improved wear in lubricated contact. ACS
845
Applied Materials and Interfaces 2017;9(30):25631-41.
846
[53] Yu G, Yamaguchi S, Kasrai M, Bancroft M. The chemical characterization of tribofilms
847
using XANES – interaction of nanosize calcium containing detergents with zinc dialkyl
848
dithiophosphate. Can J Chem 2007; 85:675–84.
849 850
[54] 1Sharma V, Gabler C, Doerr N, Aswath P. Mechanism of tribofilm formation with P and S containing ionic liquid. Tribology international 2015; 92:353-64.
49 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 50 of 54
851
[55] Hesse B, Salome M, Castillo-Michel H, Cotte M, Fayard B, Sahle C et al. Full-Field
852
Calcium K-Edge X-ray Absorption Near-Edge Structure Spectroscopy on Cortical Bone at
853
the Micron-Scale: Polarization Effects Reveal Mineral Orientation. Analytical Chemistry
854
2016; 88:3826-35.
855
[56] Patel M, Aswath P. Role of thermal, mechanical and oxidizing treatment on structure and
856
chemistry of carbon black and its impact on wear and friction Part I: Extreme pressure
857
condition. Tribology 2014.
858 859
[57] Patel M. Fundamental understanding of soot induced wear mechanism in heavy duty diesel engine [Ph.D. dissertation]. University of Texas, Arlington, TX; 2011.
860
[58] Patel M, Cristy Leonor Azanza Ricardo, Scardi P, Aswath P. Morphology, structure and
861
chemistry of extracted diesel soot—Part I: Transmission electron microscopy, Raman
862
spectroscopy, X-ray photoelectron spectroscopy and synchrotron X-ray diffraction study.
863
Tribol Int 2012;52:29-39.
864
[59] Najman M, Kasrai M, Michael-Bancroft G, Davidson R. Combination of ashless antiwear
865
additives with metallic detergents: interactions with neutral and overbased calcium
866
sulfonates. Tribology International 2006; 39:342–55.
867 868
[60] Ratoi M, Castle R, Bovington C, Spikes H.A. Influence of soot and dispersant on ZDDP film thickness and friction. Lubrication Science 2004; 17: 25–43.
869
[61] Morina A, Green J, Neville A. Surface and Tribological Characteristics of Tribofilms
870
Formed in the Boundary Lubrication Regime with Application to Internal Combustion
871
Engines. Tribology Letters 2003; 15: 443-52.
872 873
[62] Jao T-C, Li S, Yatsunami K, Chen SJ, Csontos AA, Howe JM. Soot characterisation and diesel engine wear. Lubrication Science 2004; 16:111–26.
50 ACS Paragon Plus Environment
Page 51 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
874 875
Energy & Fuels
[63] Narita K. Effects of diesel soot on engine oil performance. Journal of Japanese Society of Tribologists 1997; 42:425–9.
876
[64] Sharma H, Pahalagedara L, Joshi A, Suib S, Mhadeshwar A. Experimental Study of Carbon
877
Black and Diesel Engine Soot Oxidation Kinetics Using Thermogravimetric Analysis.
878
Energy & Fuels 2012; 26:5613–25.
879
[65] Sappok G, Wong W. Detailed Chemical and Physical Characterization of Ash Species in
880
Diesel Exhaust Entering Aftertreatment Systems. SAE Technical Paper 2007-01-0318;
881
SAE International:Warrendale, PA, 2007.
882 883
[66] Kalogirou M, Samaras Z. Soot oxidation kinetics from TG experiments. Journal of Thermal Analysis and Calorimetry - J THERM ANAL CALORIM 2010; 99:1005-10.
884
[67] Harrie J, Debora F, Renate U, Michiel M. Influence of Diesel Fuel Characteristics on Soot
885
Oxidation Properties. Industrial & Engineering Chemistry Research 2012; 51(22): 7559-64.
886
[68] Xinling L, Zhen X, Chun G, Zhen H. Oxidative Reactivity of Particles Emitted from a
887
Diesel Engine Operating at Light Load with EGR. Aerosol Science and Technology 2015;
888
49(1):1-10.
889
[69] Yehliu K, Vander Wal RL, Boehman AL. Development of an HRTEM Image Analysis
890
Method to Quantify Carbon Nanostructure. Combustion and Flame 2011; 158:1837-51.
891
[70] Liati A, Eggenschwiler P, Schreiber D, Zelenay V, Ammann M. Variations in diesel soot
892
reactivity along the exhaust after-treatment system, based on the morphology and
893
nanostructure of primary soot particles. In Combustion and Flame 2012; 160(3):671-81.
894
[71] Al-Qurashi K, Boehman AL. Impact of exhaust gas recirculation (EGR) on the oxidative
895
reactivity of diesel engine soot. Combustion and Flame 2008; 155:675-95.
51 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 52 of 54
896
[72] Liu S, Obuchi A, Uchisawa J, Nanba T, Kushiyama S. An exploratory study of diesel soot
897
oxidation with NO2 and O2 on supported metal oxide catalysts. Applied Catalysis B:
898
Environmental 2002; 37(4):309-19.
899
[73] Kamp CJ, Zhang S, Bagi S, Wong V, Monahan G, Sappok A, Wang Y. Ash Permeability
900
Determination in the Diesel Particulate Filter from Ultra-High Resolution 3D X-Ray
901
Imaging and Image-Based Direct Numerical Simulations. SAE International Journal of
902
Fuels and Lubricants 2017; 10(2):602-18. DOI: https://doi.org/10.4271/2017-01-0927
903
[74] Kamp CJ, Sappok A, Wong V. Soot and ash deposition characteristics at the catalyst-
904
substrate interface and intra-layer interactions in aged diesel particulate filters illustrated
905
using focused ion beam (fib) milling. SAE International Journal of Fuels and Lubricants
906
2012; 5:696-710.
907
[75] Custer N, Kamp CJ, Sappok A, Pakko J, Lambert C, Boerensen C, Wong V. Lubricant-
908
derived ash impact on gasoline particulate filter performance. SAE International Journal of
909
Engines 2016; 9:1604-14.
910
[76] Kamp, C. J., Bagi, S., & Wang, Y. (2019). Phenomenological Investigations of Mid-
911
Channel Ash Deposit Formation and Characteristics in Diesel Particulate Filters (No. 2019-
912
01-0973). SAE Technical Paper. DOI: https://doi.org/10.4271/2019-01-0973
913
[77] Sappok A, Rodriguez R, Wong V. Characteristics and effects of lubricant additive
914
chemistry on ash properties impacting diesel particulate filter service life. SAE International
915
Journal of Fuels and Lubricants 2010; 3:705-22.
916
[78] Mourhatch R, Aswath P. Tribological Behavior and Nature of Tribofilms Generated from
917
Fluorinated ZDDP in Comparison to ZDDP Under Extreme Pressure Conditions-Part 1:
918
Structure and Chemistry of Tribofilms. Tribology International 2011; 44:187-200.
52 ACS Paragon Plus Environment
Page 53 of 54 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Energy & Fuels
919
[79] Mourhatch R, Aswath P. Nanoscale Properties of Tribofilms Formed with Zinc Dialkyl
920
Dithiophosphate (ZDDP) Under Extreme Pressure Condition. Journal of Nanoscience and
921
Nanotechnology 2009; 9:2682-91.
922
[80] Zhang Z, Yamaguchi ES, Kasrai M, Bancroft GM. Tribofilms generated from ZDDP and
923
DDP on steel surfaces. Part 1. Growth, wear and morphology. Tribology Letters 2005;
924
19:211–20.
925
[81] Hu C, Li W, Lin Q, Zheng X, Pan H, Huanga Q. Impact of ferrocene on the nanostructure
926
and functional groups of soot in a propane/oxygen diffusion flame. RSC Advance 2017;
927
7(9):5427-36.
928 929
[82] Esangbedo C, Boehman AL, Perez JM. Characteristics of diesel engine soot that lead to excessive oil thickening. Tribology International 2012; 47:94-203.
930
[83] Knauer M, Carrara M, Rothe D, Niessner R, Ivleva NP. Changes in Structure and
931
Reactivity of Soot during Oxidation and Gasification by Oxygen, Studied by Micro-Raman
932
Spectroscopy and Temperature Programmed Oxidation. Aerosol Science and Technology
933
2009; 43(1):1-8.
934 935
[84] Vander Wal RL, Yezerets A, Currier NW, Kim DH, Wang CM. HRTEM Study of diesel soot collected from diesel particulate filters. Carbon 2007; 45:70–77.
936
[85] La Rocca A, Di Liberto G, Shayler PJ, Fay MW. The nanostructure of soot-in-oil particles
937
and agglomerates from an automotive diesel engine. Tribology International 2013; 61:80-
938
7.
939
[86] Vander Wal RL, Tomasek AJ, Street K, Hull DR, Thompson WK. Carbon Nanostructure
940
Examined by Lattice Fringe Analysis of High-Resolution Transmission Electron
941
Microscopy. Applied Spectroscopy 2004; 58:230-37.
53 ACS Paragon Plus Environment
Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 54 of 54
942
[87] Vander Wal RL, Mueller C. Initial Investigation of Effects of Fuel Oxygenation on
943
Nanostructure of Soot from a Direct-Injection Diesel Engine. Energy & Fuels 2006; 20(6):
944
2364-69.
945
[88] Yang H, Li X, Wang Y, Mu M, Li X, Kou G. Experimental investigation into the oxidation
946
reactivity and nanostructure of particulate matter from diesel engine fuelled with
947
diesel/polyoxymethylene dimethyl ethers blends. Nature publishing group 2016; 1-10.
948
[89] Yehliu K, Armas O, Vander Wal RL, Boehman A. Impact of engine operating modes and
949
combustion phasing on the reactivity of diesel soot. Combustion and Flame 2013;
950
160(3):682-91.
951 952
[90] Stachowiak GW, Batchelor AW. Engineering tribology, 2nd edition. Corrosive and Oxidative Wear, Volume 13, Butterworth Heinmann, Boston; 2001.
953
[91] Aktary M, McDermott MT, McAlpine GA. Morphology and nanomechanical properties of
954
ZDDP antiwear films as a function of tribological contact time. Tribology Letters 2002;
955
12:155–62.
956 957 958 959
[92] Barnes A, Bartle KD, Thibon V. A review of zinc dialkyl dithiophosphates (ZDDPS): characterisation and role in the lubricating oil. Tribology International 2001; 34:389–95. [93] Bagi S, Sharma V, Aswath P. Role of dispersant on soot-induced wear in Cummins ISB engine test. Carbon 2018; 136:395-408.
54 ACS Paragon Plus Environment