You are on page 1of 23

Subscriber access provided by FLORIDA ATLANTIC UNIV

Article
Monitoring grease production by reaction calorimetry and thermoanalytical
methods as an alternative to dropping point determination
Luc Vincent, Sébastien Barale, Aurélie Sandeau, Reina
Véronica Castillo, Nicole Genet, and Nicolas Sbirrazzuoli
Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.7b01920 • Publication Date (Web): 09 Sep 2017
Downloaded from http://pubs.acs.org on September 13, 2017

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 free 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 accessible to all
readers and 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.

Energy & Fuels 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 22 Energy & Fuels

1
2
3 Monitoring grease production by reaction calorimetry and thermoanalytical methods as
4
5 an alternative to dropping point determination
6
7
8
9
10
11
Luc Vincent1, Sébastien Barale1, Aurélie Sandeau², Reina Véronica Castillo2, Nicole Genet2,
12
13 Nicolas Sbirrazzuoli1,*
14
15
16 1
Université Côte d’Azur, Institut de Chimie de Nice, UMR CNRS 7272, 06100 Nice, France.
17
18
19 2
TOTAL, Centre de recherche de Solaize (CReS), Chemin du Canal, 69360 Solaize.
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 * Nicolas.Sbirrazzuoli@unice.fr
55
56
57
58
59
60 1
ACS Paragon Plus Environment
Energy & Fuels Page 2 of 22

1
2
3 Abstract
4
5
6 The present research work concerns the thermodynamic study of lithium grease production by
7
8 reaction calorimetry. In addition, several thermoanalytical techniques such as DSC and
9
10 rheology were employed to highlight grease thermal events. The results showed that the
11
12
13 cooling rate mainly controls the crystallization of soap fibers and the formulation influences
14
15 the reaction heat of saponification. The liquid fraction method (LFM) has been evaluated for
16
17 both simple and complex lithium greases as a substitution technique to dropping point. For
18
19 Lithium grease, the results showed a very good correlation between LFM data and standard
20
21
22
dropping point measurements. On the other hand, the LFM method was not applicable to
23
24 complex lithium soap for dropping point determination by DSC, while it was shown that
25
26 shear strain sweep led to a good correlation with dropping point values.
27
28
29
30
31
32
33
34
35
36
37
Keywords: Lithium grease production; saponification; liquid fraction method; reaction
38
39
40 calorimetry; thermal analysis; rheology.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 2
ACS Paragon Plus Environment
Page 3 of 22 Energy & Fuels

1
2
3 1. Introduction
4
5
6 The manufacturing of greases is known since the beginning of the XXth century, nonetheless
7
8 it remains an empirical domain1. A good understanding of the grease thermal behavior during
9
10 the cooking and cooling steps provides an accurate control of the process and an optimization
11
12
13 of the production time.
14
15
All the parameters of the grease fabrication process have an influence on the final properties
16
17
18 of the grease. For example, the soap fibers size is controlled by the cooling rate and the fiber
19
20 network give the mechanical properties of the grease. Process optimization requires
21
22 investigations on the physical and the chemical events as functions of the temperature.
23
24 Several techniques have been used to characterize physical properties of oil products such as
25
26
27 Differential Scanning Calorimetry (DSC)2 or Pressure Differential Scanning Calorimetry
28
29 (PDSC)3. Calorimetry is a technique which measures the heat flow variation versus time (at a
30
31 given constant temperature) or versus temperature. In particular, reaction calorimetry (RC)
32
33 can be an appropriate tool to study the influence of several parameters on grease production
34
35
36
because the mixing system available is well adapted to grease production processes4,5.
37
38 Compared to the DSC, the main advantage of reaction calorimetry is to dispose of mixing
39
40 system. Thus, with these thermal techniques it is possible to follow the reaction during each
41
42 step of the process. Users can obtain information on the reaction duration and enthalpy of
43
44 phenomena like dehydration, saponification and melting of simple lithium soap. Finally, the
45
46
47 adequate process can be selected when the grease that presents the optimal properties is
48
49 obtained. This technique requires few matters and is transposable to any processes for simple
50
51 greases.
52
53
54 The aim of this study is to show the benefits provided by the use of reaction calorimetry in
55
56 association with DSC and rheology for the industrial elaboration of lithium soap and complex
57
58
59
60 3
ACS Paragon Plus Environment
Energy & Fuels Page 4 of 22

1
2
3 lithium soap and to better understand reactions taking place during the industrial process. In
4
5 addition, new methods for substitution of dropping point technique will be proposed for both
6
7 simple and complex lithium greases.
8
9
10
11 2. Experimental
12
13
14 2.1. Reaction calorimetry (RC)
15
16
17 Reaction calorimetry measurements were performed with a C80 reaction calorimeter
18
19 commercialized by Setaram. Measurements were made using the membrane mixing cell and
20
21 PTFE membranes. The mixing was performed by a system developed in our laboratory and
22
23 specifically designed for grease production.
24
25
26 2.2. Differential scanning calorimetry (DSC)
27
28
29 DSC measurements were carried out on a Mettler-Toledo DSC 1 equipped with a FRS5
30
31
sensor and STAR© software for data analysis. Temperature and enthalpy calibrations were
32
33
34 performed by using indium and zinc standards. Samples crucibles were filled with 8 mg of
35
36 lithium soap and placed in 40 µL aluminum crucibles.
37
38
39
40
41 2.3. Dropping point determination (DP)
42
43
44 The dropping point of a lubricating grease is the temperature at which it passes from a semi-
45
46 solid to a liquid state under specific test conditions. It is a measure of the cohesiveness of the
47
48 oil and thickener of grease. Dropping point measurements have been performed at a heating
49
50 rate of 2 K min-1 using a DP70 Dropping Point System commercialized by Mettler-Toledo.
51
52
53
54
55
56
57
58
59
60 4
ACS Paragon Plus Environment
Page 5 of 22 Energy & Fuels

1
2
3 2.4. Rheology
4
5
6 Rheological analyses were realized with a rotational rheometer (Kinexus Pro+ Malvern) at
7
8 25°C using plate-plate geometries (25 mm diameter, 0.5 mm gap). Strain sweep tests, at a
9
10 frequency of 1 Hz, were previously performed on each sample to determine the linear
11
12
13 viscoelastic region. All samples were stored at room temperature. Before each rheological
14
15 test, grease samples were homogenized to erase the material aging.
16
17
18
19
20 2.5. Grease preparation
21 Simple Lithium greases samples (LiS) were prepared in situ into the calorimeter through a
22
23
24 saponification reaction between 12-hydroxystearic acid (12-HSA) and monohydrated lithium
25
26 hydroxide (LiOH, H2O) within a naphthenic base oil medium. The formulation of the grease
27
28 was defined in order to reach a specific NLGI grade of 2. The NLGI grade expresses a
29
30 measure of the relative hardness of grease used for lubrication, as specified by the standard
31
32
33
classification of lubricating grease established by the National Lubricating Grease Institute
34
35 (NLGI). The reactants were introduced inside the calorimeter equipment with the following
36
37 repartition between both parts of the calorimetric cell: the lithium hydroxide was placed in the
38
39 upper part with 50 % of naphthenic base oil and 12-hydroxystearic acid was incorporated in
40
41 the bottom part. Concerning complex lithium grease samples (LiX), the azelaic acid
42
43
44 complexing agent was introduced in the same part as 12-HSA.
45
46 Greases production processes are well-known since the beginning of XXth century1. A grease
47
48 production requires typically two heating steps and one cooling step. An isothermal step can
49
50 be added between both heating rates to optimize saponification reaction as shown on Fig. 1.
51
52
53
Several parameters such as heating rate and isothermal duration can be adjusted to optimize
54
55 grease fabrication. The thermal and mechanical performances of the grease are mainly related
56
57 to the process. In this work, the process used is described by the following steps. The first
58
59
60 5
ACS Paragon Plus Environment
Energy & Fuels Page 6 of 22

1
2
3 heating step was performed at a heating rate of 1 K min-1 from 60 °C to 115 °C. During this
4
5 first step, the melting phenomenon of 12-HSA is occurred. An isothermal step was then
6
7 performed at 115 °C during 45 minutes. The saponification reaction was initiated by
8
9
10 perforating the PTFE membrane to generate the mixing between 12-hydroxystearic acid and
11
12 lithium hydroxide into naphthenic oil. The saponification occurs during the isothermal step at
13
14 the temperature of 115 °C. The final step was composed of a heating rate at 1°C.min-1 from
15
16 115 °C to 215 °C. The cooling step was performed at a cooling rate of -1 °C.min-1 from 215
17
18
°C to room temperature, i.e. 20°C in this study.
19
20
21
22
23
24 3. Results / discussion
25
26
27 3.1. Obtention of simple lithium soap
28
29
30 An example of curves obtained during the LiS cooking step is reported in Fig.1. The first
31
32
33 thermal event is an endothermic phenomenon characteristic of 12-HSA melting. The melting
34
35 temperature measurement of 12-HSA (peak temperature at 81,1 ± 0.5°C (A), Tonset at 75,2 ±
36
37 0.5°C and Tendset at 85,0 ± 0.5°C, see Fig.1) was in agreement with literatures values (82 ±
38
39 1°C 6 ).
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 6
ACS Paragon Plus Environment
Page 7 of 22 Energy & Fuels

1
2
3
4
5 150 210
6 Saponification 200
7 Exo
120 190
8
9 175°C 180
10 170
90
11 160
Heat Flow / mW

12

Temperature / °C
45 min. 150
13 60 140
14 Contacting 130
15 reactants
30 120
16 115°C
17 110
18 0 Gap B 100
19 C • 90
20 • 80
21 -30 81°C 70
22
60
23
-60
A • 50
24
25 0 30 60 90 120 150 180 210
26 Time / min
27
28
29 Fig. 1. Heat Flow variation obtained by reaction calorimetry during the cooking step of
30
31
32 simple lithium soap (red: temperature into the measuring cell).
33
34
35
36
37
After membrane perforation at 55 minutes (see. Fig.1), all products were mixed and an
38
39
40 exothermic phenomenon was immediately recorded. C80 calorimeter has a high time constant
41
42 (around 360s). Thus, to correct the measured thermal effect, a numerical treatment was
43
44 applied using special software developed in our laboratory7. The exothermal event was related
45
46 to the reaction between lithium hydroxide and 12-HSA as an acid-base reaction. The reaction
47
48
49 starts immediately and generates water. Dehydration heat flow was masked by the high
50
51 enthalpy of saponification. The duration of reaction was evaluated at 45 minutes (section in
52
53 blue on the Fig.1). The gap observed between 55 min (C) and 100 min on the baseline was
54
55 explained by the heat capacity (Cp) variation. The baseline level after the reaction was higher
56
57
due to the heat capacity decreases during the reaction. The mixture was then heated to
58
59
60 7
ACS Paragon Plus Environment
Energy & Fuels Page 8 of 22

1
2
3 complete soap formation and dehydration. Melting point of lithium hydroxystearate was
4
5 reported to be around 180°C. The value obtained on the thermogram of Fig. 1 is around
6
7 175°C (B), which is in good agreement with literature8. However the broad shape of the
8
9
10 thermal event does not correspond to the thermal fingerprint of the melting of a pure
11
12 compound. The experimental conditions were modified in order to better highlight the melting
13
14 phenomenon: a cooling step was added after the isothermal step from 115 °C to 90°C
15
16 followed by a heating to higher temperature (until 240°C) (see. Fig.2).
17
18
19
20
21
20
22
23
24 10
25
26 0
27
28
Heat flow / mW

29 -10
30
31 -20
32
33
34 -30
35
36 -40
37 Exo
38
-50
39
40
41 -60
42 120 140 160 180 200 220
43 Temperature / °C
44
45
46
47
48
49 Fig. 2. Lithium soap melting during the grease cooking step (red LiS no cooling step, black
50
51 LiS with cooling step)
52
53
54
55
56
57
58
59
60 8
ACS Paragon Plus Environment
Page 9 of 22 Energy & Fuels

1
2
3 These modifications provided a sharper melting peak at higher temperature. In this case, the
4
5 melting point of soap was determined more accurately and a value of 187,0°C ± 0.5°C was
6
7 measured using the onset temperature of the melting peak (Tonset).
8
9
10 Lithium Greases are composed of a soap fiber network whose morphology strongly
11
12
impacts the final performances in particular thermal and mechanical stabilities9,10,11. The
13
14
15 structure of the network can be controlled using appropriate heating and cooling rates during
16
17 the processes. The combination of short and long fibers provides interesting properties1. The
18
19 length of the fibers is controlled by the cooling rate. The faster is the cooling rate, the shorter
20
21 are the fibers12. Indeed, the system has more time to crystallize and longer fibers are built
22
23
24 when temperature decreases slowly. Enthalpy of crystallization is also modified with the
25
26 cooling rate (Table 1). If the sample has more time to crystallize, the final sample crystallinity
27
28 is higher, leading to a higher crystallization enthalpy and the shape of curves is substantially
29
30 different as seen in Fig. 3. These effects are perfectly highlighted by calorimetric
31
32
measurements (Table 1 and Fig. 3). The lower cooling rate provides a peak separation with at
33
34
35 least two thermal events as observed at -0.3°C.min-1 with a shoulder at 157 °C and a peak at
36
37 168 °C .
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 9
ACS Paragon Plus Environment
Energy & Fuels Page 10 of 22

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 Fig. 3. Influence of the cooling rate on the crystallization of lithium soap for LiS grease.
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 10
ACS Paragon Plus Environment
Page 11 of 22 Energy & Fuels

1
2
3 Table 1
4
5 Temperature and enthalpy of crystallization of the grease LiS as a function of cooling rate
6
7 Ref Cooling rate Peak Onset Endset Enthalpy
8 Grease (°C.min-1) Temperature Temperature Temperature (J.g-1)
9 LiS Tpeak Tonset (°C) Tendset (°C)
10 (°C)
11 1 0.3 170.4 ± 0.5 174.3 ± 0.5 162.0 ± 0.5 11 ± 1
12 2 0.8 159.5 ± 0.5 164.1 ± 0.5 146.7 ± 0.5 7±1
13
3 1.0 158.2 ± 0.5 162.7 ± 0.5 151.4 ± 0.5 7±1
14
15
16 3.2.Obtention of complex lithium soap
17
18
19 The thermogram obtained with reaction calorimetry measurement during the cooking of
20
21 complex lithium soap (LiX) is shown in Figure 4. The overall shape of the thermogram of the
22
23
24 complex lithium soap is very similar to that obtained with lithium soap. The first thermal
25
26 event occurring between 60°C and 85°C is attributed to the melting of 12 HSA. The second
27
28 thermal event is the result of the convolution of several thermal effects. Indeed, two main
29
30 reactions occur in parallel. Each acid react with lithium hydroxide to give soap and water. 12
31
32
HSA and one part of lithium hydroxide are reacting while azelaic acid is melting. The second
33
34
35 part of lithium hydroxide reacts with liquid azelaic acid. These two reactions release water.
36
37 Dehydration and saponification reactions are exothermic whereas melting of azelaic acid is an
38
39 endothermic event. The overall thermal event mainly reflects the exothermic effects because
40
41 their intensities are higher than the endothermic effect. The enthalpy of saponification is
42
43
44 higher. The melting of lithium stearate is observed at around 175°C and the melting of lithium
45
46 azelate is not visible below 200°C. Heat flow variation measured during complex lithium soap
47
48 formation isn't so different from that of simple lithium soap though mechanism of reaction is
49
50 different. At this stage it isn't possible to come to a conclusion about the complexation. The
51
52
characterization of grease by dropping point measurement is necessary to confirm obtaining
53
54
55 complex lithium soap.
56
57
58
59
60 11
ACS Paragon Plus Environment
Energy & Fuels Page 12 of 22

1
2
3
4
5 180 200
6
7 Exo 180
8
9
10 120 160
11
12 140
Heat Flow / mW

Temperature / °C
13
14
60 120
15
16
17 100
18
19 0 80
20
21 60
22
23
24 -60 40
25 0 30 60 90 120 150 180
26 Time / min
27
28
29 Fig. 4. Thermogram obtained with reaction calorimetry measurement during the cooking of
30
31
32 complex lithium soap (LiX).
33
34
35
36 3.3.Grease characterization
37
38
39 To ensure that grease produced in the calorimeter corresponds to grease targeted and to the
40
41 industrial requirements, three tests are performed in industry. These tests are named: dropping
42
43 point, NLGI grade evaluation and basicity test. For simple lithium soap samples (LiS) and
44
45 complex lithium soap samples, the desired NLGI grade is the grade 2, the dropping point of
46
47
48 these samples must be close to 180°C ± 7°C and 195°C ± 7°C respectively and the basis
49
50 character have to be revealed by the test. If one or more tests do not correspond to the targeted
51
52 values, this means that the synthesis of the greases was not correctly done. All results
53
54 presented in this work were obtained on adequate greases.
55
56
57
58
59
60 12
ACS Paragon Plus Environment
Page 13 of 22 Energy & Fuels

1
2
3 3.3.1. Dropping point
4
5
6 The dropping point is measured with a standardized test given by procedures in ASTM
7
8 standards D556 and D2265 13. The dropping point is defined as the temperature at which the
9
10 grease goes from a semi-solid to a liquid state under specific test conditions. Because most of
11
12
13 the physico-chemical transitions lead to heat exchanges, they can be measured by DSC with
14
15 high accuracy. Thus, in the following the transition temperature measured by DSC was
16
17 compared to dropping point determinated by ASTM D556. Mettler-Toledo society14 has used
18
19 DSC to evaluate the dropping point of edible fat with success. The method applied is called
20
21
22
liquid fraction method (LFM). In this study, this method was transposed to the evaluation of
23
24 the dropping point of lithium greases. According to this method, the temperature of dropping
25
26 point of edible fat corresponds to the temperature at 95% of conversion. All measurements
27
28 have been conducted at a heating rate of 10 K min-1 between 20 and 320°C.
29
30
31 Heat flow variation shows an endothermic event corresponding to lithium stearate melting
32
33 (Fig. 5). The shape of the DSC curve is well defined in agreement with expected results14.
34
35
36
Temperatures at 95% conversion of ten samples are compared to each dropping point values
37
38 and summarized in Table 2.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 13
ACS Paragon Plus Environment
Energy & Fuels Page 14 of 22

1
2
3
4
5 0
6
7 Exo 100
8
9 -1
10 80
11

Extent of conversion
Heat flow / mW

12
-2 60
13
14
15
16 -3 40
17
18
19 20
20 -4
21
22 0
23 -5
24 160 170 180 190 200 210 220
25
26 T / °C
27
28
29 Fig. 5. Application of liquid fraction method to lithium soap. Measured heat flow (blue, left
30
31 axis) and extent of conversion (red, right axis).
32
33
34
35
36 Table 2
37
38 95% melting temperature compared to ASTM dropping point for lithium grease (LiS)
39
40
41
Sample Id Temperature at 95% conversion by Dropping point ASTMD556 (°C)
lithium DSC (°C) precision given by the manufacturer ± 7°C
42
43 stearate precision given by the manufacturer ± 1°C
44 1 195 198
45 2 204 202
46
47
3 208 206
48 4 189 187
49 5 189 189
50 6 194 191
51 7 190 188
52 8 194 194
53
9 187 188
54
55 10 188 187
56
57
58
59
60 14
ACS Paragon Plus Environment
Page 15 of 22 Energy & Fuels

1
2
3 The values measured by DSC are in very good agreement with those obtained for dropping
4
5 point using ASTM D556 method. In the case of dropping measurements, bleeding can occurs
6
7 and the test has to be reproduced. With a DSC measurement the bleeding problem is removed
8
9
10 and measurement duration is reduced. Because of the high sensitivity of the apparatus, DSC
11
12 measurement allows dropping point determination with more precision. In addition, the
13
14 measurement is easier and faster which is also an important advantage in regard to industrial
15
16 applications. Thus the method of the liquid fraction seems easily transposable in the case of
17
18
lithium soap.
19
20
21
22
23 Complex lithium soap has higher dropping point than lithium soap. Generally, complex
24
25 lithium soap without additives has a dropping point near to 250°C. Heat flow measured by
26
27 DSC shows several thermal events not present for lithium soap (Fig. 6). Thus the melting
28
29
30 point of lithium stearate and the melting point of lithium azelate and lithium diazelate can be
31
32 observed on thermogram. We first observe an endothermic peak at a temperature onset of 181
33
34 °C ± 1°C, that correspond to the melting of lithium stearate, followed by at least three melting
35
36 points at 248°C, 269°C and 280°C respectively.
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 15
ACS Paragon Plus Environment
Energy & Fuels Page 16 of 22

1
2
3
4
5
6
0.00 100
7
8
9
10 -0.05 Exo 80
11

Extent of conversion
12
Heat flow / mW

60
13 -0.10
14
15
40
16
-0.15
17
18
20
19
20 -0.20
21
0
22
23 -0.25
24 50 100 150 200 250 300
25 T / °C
26
27
28
29
30 Fig. 6. Liquid Fraction Method (LFM) applied to complex lithium grease (LiX).
31
32
33 Temperature values taken at 95% conversion do not correspond to dropping point values (data
34
35
36 not presented here). A better agreement between the two methods is obtained when an extent
37
38 of conversion of 98% is used instead of 95% (Table 3), but it still exist a great difference
39
40 between the values obtained from the two methods and no any correlation is observed.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 16
ACS Paragon Plus Environment
Page 17 of 22 Energy & Fuels

1
2
3 Table 3
4
5 98% melting temperature obtained from DSC measurements compared to ASTM dropping
6 point for complex lithium grease (LiX)
7
8 Sample Id Temperature at 98% conversion by Dropping point ASTMD556 (°C)
9 lithium DSC (°C) precision given by the manufacturer ± 7°C
10 stearate precision given by the manufacturer ± 1°C
11 1 280 262
12
2 280 257
13
14 3 272 284
15 4 278 290
16 5 288 287
17 6 258 270
18 7 280 295
19 8 279 282
20
21
9 270 292
22 10 252 269
23
24
25 This method cannot be used for complex lithium greases. This can be explained because
26
27 dropping point of complex lithium soap is linked to the melting of a complex fiber network.
28
29
30
This network is the result of several soaps formation. The structure of the fiber network
31
32 confers mechanical properties to the grease and in particular its thermo-mechanical resistance.
33
34 The mechanical properties can be characterized by rheological measurements15,16,17,18,19,20.
35
36 The ASTM method is mainly related to the mechanical response of the sample. This explains
37
38
why the values obtained are not directly correlated with the values obtained with DSC.
39
40
41 Indeed, the calorimetric response reflects the temperature at which 95% of the melting
42
43 enthalpy of the grease has been absorbed by the calorimeter. This temperature does not
44
45 correspond necessarily to the temperature were the sample start to flow. In this context, we
46
47 decided to investigate a potential link between rheological properties and dropping point. This
48
49
50
measurement was performed on complex lithium soap (Fig. 7). Based on dropping point
51
52 definition, the destruction of the fiber network induces the loss of grease mechanical
53
54 resistance. In rheological analysis, the destructuration of a physical network structure can be
55
56 highlighted during a shear strain sweep by a crossing between storage modulus (G’) and loss
57
58
59
60 17
ACS Paragon Plus Environment
Energy & Fuels Page 18 of 22

1
2
3 modulus (G’’). For gels systems, the intersection of G' and G” is used to detect gel-sol
4
5 transition. As mentioned in earlier studies, the grease loses all its mechanical properties in the
6
7 same way that it loses its mechanical stability when the dropping point is reached21. To check
8
9
10 the validity of the proposed method, an exploratory study has been driven on several complex
11
12 greases with various dropping points.
13
14
15
16 100000
G' storage modulus
17 G" loss modulus
18
19
20
21
10000
G' & G" / Pa.s

22
23
24
25
26
27 1000
28
29
30
31
32 100
33 0.1 1 10 100 1000
34
Shear Strain / %
35
36
37
38
39 Fig. 7. Shear strain sweep of complex lithium soap LiX at 25°C using plate-plate geometries
40
41 (25 mm diameter, 0.5 mm gap) at a frequency of 1 Hz.
42
43
44 For each complex lithium grease, the dropping point has been measured by a DP700
45
46 Dropping point system and the crossing point (S) between loss modulus and storage modulus
47
48 have by determined by rheology. The percentage of shear strain, observed at the crossing
49
50 point (S), increases with the dropping point temperature. A representation of the relation
51
52
53 between the dropping point and the crossing point S is shown in Fig. 8. Moreover, the shear
54
55 strain obtained at the flowing point seems strongly correlated to the dropping point
56
57
58
59
60 18
ACS Paragon Plus Environment
Page 19 of 22 Energy & Fuels

1
2
3 temperature as shown in Fig. 8. In the case of complex lithium grease obtained by the same
4
5 process the dropping point could be estimated using rheological measurements.
6
7
8 Table 4
9
10 Flowing point of a complex lithium soap
11
12 Reference of the grease LiX with is Shear strain values of flowing point (%)
13 dropping temperature (°C)
14 LiX298 – 298 °C 48
15 LIX296 – 296°C 46
16
LIX292 – 292 °C 39
17
18 LiX290 – 290 °C 34
19 LiX282 – 282°C 21
20
21
22
23
24
25 50
26
Shear strain values of flowing point / %

27
28 45
29
30
40
31
32
33 35
34
35
36 30
37
38 25
39
40
41 20
42
43 280 285 290 295 300
44
45 Dropping temperature / °C
46
47
48 Fig. 8. Shear strain values of flowing point vs. dropping temperature for complex lithium soap
49
50 LiX.
51
52
53
54
55
56
57
58
59
60 19
ACS Paragon Plus Environment
Energy & Fuels Page 20 of 22

1
2
3 4. Conclusion
4
5
6 The present work reveals that reaction calorimetry can be considered as an interesting
7
8 technique to understand thermal events during grease fabrication and predict their final
9
10 properties before industrial scale production. Accuracy of temperature events measurements
11
12
13 can be exploited to extract reliable data and transpose them to industrial process. Melting
14
15 point can be determined with high accuracy. The dropping point temperature which is the key
16
17 property of lithium soap can be determined with high accuracy using the liquid fraction
18
19 method (LFM) by DSC. This method allows obtaining faster results and can be used to
20
21
22
overcome oil leakage observed in ASTM method. Unfortunately, LFM method is not
23
24 applicable in the case of complex lithium greases due to their particular network. To
25
26 overcome this, we investigate a promising method based on rheological characterization of
27
28 the network destructuration obtained by shear strain sweep and its correlation with dropping
29
30
point temperature.
31
32
33
34
35
36 5. Acknowledgements
37
38 The authors gratefully acknowledge TOTAL Company for the fruitful collaboration and the
39
40
41 financial support.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 20
ACS Paragon Plus Environment
Page 21 of 22 Energy & Fuels

1
2
3 References
4
5
6
7 (1) Boner C.J., Manufacture and application of lubricating greases. National Lubricating
8 Grease Institute, library of congress catalog card number. Original edition 1954 ; reprint
9 1983, 54-11031.
10 (2) Santos J.C.O., Garcia dos Santos I.M., Souza A.G., Sobrinho E.V., Fernandes Jr V.J.,
11 Silva A.J.N. Thermoanalytical and rheological characterization of automotive mineral
12 lubricants after thermal degradation. Fuel 2004, 83, 2393-2399.
13 (3) Zhou J., Xiong Y., Liu X. Evaluation of the oxidation stability of biodiesel stabilized with
14 antioxidants using the Rancimat and PDSC methods, Fuel 2017, 188, 61-68.
15
(4) Drabik J., Trzos M., Ilowska J., Bereska B. Complex greases produced in calorimetric
16
17 reactor. Part I. Optimization of process parameters with respect of the grease quality, Przem.
18 Chem 2013, 92, 10, 1846-1849.
19 (5) Ilowska J., Szwach I., Bereska B., Gniady J., Szmatola M., Semeniuk I., Drabik J.,
20 Kozdrach R. Complex greases produced in calorimetric reactor. Part 2. Effect of process
21 parameters on properties of the greases, Przem. Chem 2013, 92, 10, 1850-1852.
22 (6) Lide D.R., Frederikse H.P.R., CRC handbook of chemist and physics. Boca Raton, FL:
23 CRC Press. 1993; 3465.
24 (7) Barale S., Vincent L., Sauder G., Sbirrazzuoli N. Deconvolution of calorimeter response
25
from electrical signals for extracting kinetic data, Thermochimica Acta 2015, 615, 30–37.
26
27 (8) Delgado M.A., Sanchez M. C., Valencia C., Franco J. M., Gallegos C., Relationship
28 among microstructure, rheology and processing of a lithium lubricating grease, Chemical
29 Engineering Research and Design, 2005, 83(A9): 1085–1092.
30 (9) Podlennykh L., V., Ishchuk L.P. Influence of cooling rate and soap concentration on
31 isothermal crystallization temperature of lithium greases, Plenum Publishing Corporation
32 1987; 50, 9.
33 (10) Vold M.J., Colloidal structure in lithium stearate grease, J. Phys. Chem. 1956, 60(4) 439-
34 442.
35
(11) Bondi A.A., Proceedings of the Third World Petroleum Congress The Hague, E.J. Brill;
36
37 First edition, 1951.
38 (12) Martín-Alfonso J.E., Valencia C., Sánchez M.C., Franco J.M., Gallegos C. Influence of
39 some processing variables on the rheological properties of lithium lubricating greases
40 modified with recycled polymers, Int. J. Materials and Product Technology 2012, 43(1-4),
41 184-200.
42 (13) ASTM D556, Test Method for Dropping Point of Lubricating Grease and ASTM D2265,
43 Standard Test Method for Dropping Point of Lubricating Grease Over Wide Temperature
44 Range, ASTM International, West Conshohocken, PA, 2010, www.astm.org (consulted on
45
10/28/2016).
46
47 (14) Mettler-Toledo, Thermal analysis Usercom n°11, 2000, www.mt.com (consulted on
48 10/28/2016).
49 (15) Franco J.M., Delgado M.A., Valencia C., Sánchez M.C., Gallegos C. Mixing rheometry
50 for studying the manufacture of lubricating greases, Chem. Eng. Sci. 2005, 60, 2409 – 2418.
51 (16) Delgado M.A., Valencia C., Sanchez M.C., Franco J.M., Gallegos C. Influence of Soap
52 Concentration and Oil Viscosity on the Rheology and Microstructure of Lubricating Greases,
53 Ind. Eng. Chem. Res. 2006, 45, 1902-1910.
54 (17) Martın-Alfonso J.E., Moreno G., Valencia C., Sanchez M.C., Franco J.M., Gallegos C.
55
Influence of soap/polymer concentration ratio on the rheological properties of lithium
56
57 lubricating greases modified with virgin LDPE, J. Ind. Eng. Chem. 2009, 15, 687–693.
58
59
60 21
ACS Paragon Plus Environment
Energy & Fuels Page 22 of 22

1
2
3
4
5 (18) Delgado M.A., Valencia C., Sanchez M.C., Franco J.M., Gallegos C. Thermorheological
6 behaviour of a lithium lubricating grease, Tribol. Lett. 2006, 23(1), 47-54.
7 (19) Delgado M.A., Franco J.M., Valencia C., Kuhn E., Gallegos C., Transient shear flow of
8 model lithium lubricating greases, Mech Time-Depend Mater 2009, 13, 63–80.
9 (20) Yonggang M., Jie Z.A. Rheological model for lithium lubricating grease, Tribology
10 International 1998, 31(10), 67-70.
11 (21) Agernas O., Tengberg T., Development of two methods to evaluate lubricating greases
12
13
using a rheometer, Chambers, Göteborg, Sweden 2011.
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 22
ACS Paragon Plus Environment

You might also like