Model Water-in-Oil Emulsions for Gas Hydrate Studies in Oil

Jul 18, 2013 - Stable water-in-oil emulsions with water volume fraction ranging from 10 to 70 vol % have been developed with mineral oil 70T, Span 80,...
1 downloads 0 Views 13MB Size
Subscriber access provided by UNIV OF ALABAMA HUNTSVILLE

Article

Model Water-in-Oil Emulsions for Gas Hydrate Studies in Oil Continuous Systems Jose G. Delgado-Linares, Ahmad Afif Abdul Majid, E. Dendy Sloan, Carolyn A. Koh, and Amadeu K. Sum Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/ef4004768 • Publication Date (Web): 18 Jul 2013 Downloaded from http://pubs.acs.org on July 19, 2013

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 34

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

Model Water-in-Oil Emulsions for Gas Hydrate

2

Studies in Oil Continuous Systems

3

Jose G. Delgado-Linares1,2, Ahmad A. A. Majid1, E. Dendy Sloan1, Carolyn A. Koh1, and Amadeu

4

K. Sum1*

5 6 7 8 9 10 11

1

Center for Hydrate Research, Chemical & Biological Engineering Department, Colorado School of Mines, 1600 Illinois St., Golden, CO 80401 – USA 2 Laboratorio FIRP, Escuela de Ingeniería Química, Universidad de Los Andes, Mérida 5101, Venezuela

KEYWORDS: water-in-oil emulsion, hydrates, stability, model emulsion ABSTRACT

12

Stable water-in-oil emulsions with water volume fraction ranging from 10 vol.% to 70 vol.%

13

have been developed using Mineral Oil 70T, Span 80, sodium di-2-ethylhexylsulfosuccinate

14

(AOT), and water. The mean size of the water droplets ranges from 2 to 3 μm. Tests conducted

15

show that all emulsions are stable against coalescence for at least one week at 2°C and room

16

temperature. Furthermore, it was observed that the viscosity of the emulsion increases with

17

increasing water volume fraction, with shear thinning behavior observed above certain water

18

volume fraction emulsions (30 vol.% at room temperature and 20 vol.% at 1°C). Viscosity tests

19

performed at different times after emulsion preparation confirm that the emulsions are stable for

20

one week. Differential scanning calorimetry performed on the emulsions shows that for low

21

water volume fraction emulsions (< 50 vol.%), the emulsions are stable upon ice and hydrate

ACS Paragon Plus Environment

1

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 34

22

formation. Micromechanical force (MMF) measurements show that the presence of the

23

surfactant mixture has little to no effect on the cohesion force between cyclopentane hydrate

24

particles, although a change in the morphology of the particle was observed when the surfactant

25

mixture was added into the system. High-pressure autoclave experiments conducted on the

26

model emulsion resulted in a loose hydrate slurry when the surfactant mixture was present in the

27

system. Tests performed in this study show that the proposed model emulsion is stable, having

28

similar characteristic to that observed in crude oil emulsions, and may be suitable for other

29

hydrate studies.

30 31

1.INTRODUCTION

32

Gas hydrates are crystalline compounds in which small gas molecules are enclathrated by

33

hydrogen-bonded water molecules,1 typically forming at high pressure and low temperature. The

34

condition of high pressure and low temperature in crude oil production pipelines provides

35

favorable conditions for gas hydrate formation. The accumulation and aggregation of hydrates in

36

pipelines can result in pipeline blockage.1,2

37

Significant advances in the area of thermodynamics, such as phase equilibrium of different gas

38

hydrate systems has been achieved.3 In fact, much of the current knowledge in thermodynamics

39

is being used as a predictive tool in phase equilibria programs.4 In contrast, the kinetics of

40

hydrates formation (from nucleation to growth) is less well understood and constitutes the

41

subject of numerous current investigations.4,5 Over the past decade, a conceptual model that

42

depicts hydrate plug formation in water, oil and gas systems has been developed and gained

43

acceptance within the flow assurance community6 (Figure 1).

ACS Paragon Plus Environment

2

Page 3 of 34

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

Due to the turbulence in flow in pipelines, water typically is dispersed into the oil phase,

45

creating a water-in-oil (W/O) emulsion. At the interface of water droplets, hydrate shells will

46

grow. Interactions among hydrate-encapsulated water droplets cause agglomeration of hydrate

47

particles that can become large aggregates and eventually plug pipelines. For the hydrate growth

48

at water/oil interfaces, Taylor et al.7 proposed a mechanism that comprises three stages: 1)

49

formation of a thin and porous hydrate film that propagates across the interface, 2) film

50

thickening over time and filling of the pores, and 3) bulk conversion of the hydrate film with

51

filling of the pores. In emulsified systems, irrespective of the amount of water, formation of

52

hydrates can lead to droplet agglomeration which will then lead to plugging of pipelines,

53

whereas their dissociation promotes the destabilization and/or inversion of the emulsion.8

54

Emulsions

55

Figure 1. Conceptual model of hydrate formation, agglomeration and plugging in a water-in-oil

56

emulsified system under flowing conditions (from Turner,8 in collaboration with J.

57

Abrahamson).6

58

Several studies have demonstrated that emulsion stability is an important criterion in

59

preventing hydrate plug formation;9,10 the higher the emulsion stability, the lower the risk of

60

hydrate agglomeration. In addition, it has been identified that natural surfactants (such as,

61

asphaltenes and organic acids) present in petroleum crude, play a major role in emulsion

62

stability, and as a consequence in hydrate agglomeration and plugging.9-11

ACS Paragon Plus Environment

3

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 34

63

Salager and Forgiarini12 have suggested applying the concepts of the physicochemical

64

formulation of surfactant-oil-water systems to control the transportability of hydrates in

65

emulsified systems. Stabilizing the emulsions using appropriate surfactants will create a stable

66

dispersed system and resistance against agglomeration even in the presence of hydrates. Once the

67

fluid arrives at separation facilities, the formulation may be shifted by various methods, such as

68

by changing the temperature. Emulsion breakage is then attained after hydrate dissociation.

69

Hydrate studies, specifically in water-in-oil emulsions, require the use of a stable and well-

70

defined system in order to obtain reproducible results and to separate the impact of hydrates from

71

those of crude oil chemistry. Many studies using water-in-crude oil emulsions have contributed

72

to important understanding of hydrates behavior. However, the complexity of crude oils presents

73

a number of issues, which involves not only differences from one type of crude oil to another,

74

but also differences from batch to batch of a same crude oil.10,13

75

In this context, the work presented in this paper focuses on the development and

76

characterization of a model water-in-oil emulsion suitable for hydrate studies. As the model

77

system will be used to for hydrate studies in oil continuous systems, stability is the main

78

requirement set for the model system. The emulsion should remain stable throughout the

79

experimental studies, which for our interest is about one week. Another requirement for the

80

model emulsion is to have a system with well-defined surface-active component(s) in order to

81

separate the contributions to hydrate formation and behavior in emulsified system. The use of

82

asphaltenes to stabilize the emulsion is not suitable in the present work since the specific

83

structure and interfacial behavior of asphaltenes are unknown. The model system developed here

84

is a water-in-mineral oil emulsion stabilized by a surfactant mixture (Span 80 and AOT). Results

ACS Paragon Plus Environment

4

Page 5 of 34

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

85

are presented on the development and characterization of the model emulsion, including several

86

tests of the model emulsion with hydrates, at static and dynamic conditions.

87 88

2. EXPERIMENTAL METHOD

89

2.1.Emulsion Preparation

90

Water-in-oil emulsions with water volume fraction (or water cut, as commonly denoted in the

91

oil industry) ranging from 10 to 70 vol.% are prepared using Crystal Plus Mineral Oil 70T (STE

92

Oil Company, Inc.) with reported viscosity of 10.2 cP at 40 °C and density of 0.857 g/cc at 20

93

°C. A compositional analysis of Mineral Oil 70T is shown in Table 1. In all cases, the emulsifier

94

agents are composed of an anionic-cationic surfactant mixture. Two surfactant concentrations are

95

studied and reported here: 1 and 5 wt.% surfactant mixture with respect to the total emulsion

96

system. De-ionized water is used as the aqueous phase in all emulsions.

97

The composition of surfactant mixture is 90 wt.% Sorbitan monooleate (Span 80, MW 428.61

98

g/mol, Sigma Aldrich) and 10 wt.% Sodium di-2-ethylhexylsulfosuccinate (AOT, MW 444.56

99

g/mol, Fisher Scientific). The chemical structure of the surfactant molecules are shown in Figure

100

2. This surfactant mixture is suitable to produce oil continuous emulsions, since both surfactants

101

are oil-soluble (lipophilic) molecules. This is supported by Bancroft’s rule as it is stated that in

102

any emulsions, the external (continuous) phase is the one, which solubilizes the surfactant.14,15

103

Furthermore, the major component of the surfactant mixture, Span 80, is a surfactant with a

104

Hydrophilic Lipophilic Balance (HLB) value of 4.3. The HLB is a measure of the surfactant

105

hydrophilicity, which is defined as 1/5 of the weight percent of ethylene oxide in the

106

molecule.16,17 Generally, surfactants with HLB values below 10 will produce water-in-oil

107

emulsifiers.18

ACS Paragon Plus Environment

5

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

108

Page 6 of 34

Table 1. Mineral Oil 70T Composition (data provided by Chevron ETC).

Component

Mass %

Component

Mass %

C16

0.09

C24

4.23

C17

1.23

C25

3.76

C18

5.22

C26

3.29

C19

11.75

C27

2.66

C20

16.04

C28

2.27

C21

17.04

C29

1.56

C22

12.20

C30+

12.34

C23

6.34

Average MW

311

Density (g/cm3)

0.856

109

110 111

Figure 2. Molecular structure for surfactants: a) Sorbitan monooleate (Span 80); b) Sodium di-2-

112

ethylhexylsulfosuccinate (AOT).

ACS Paragon Plus Environment

6

Page 7 of 34

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

113

The emulsions are prepared by dissolving the surfactant mixture in mineral oil using a

114

magnetic stirrer and low heating (~50ºC). For water cut ≤ 50%, the mixture is stirred using a

115

high-speed homogenizer at 8000 rpm for three minutes, while water is slowly added during the

116

first minute. For 60% and 70% water cut emulsions, the mixture is stirred at 8000 rpm for six

117

minutes while water is added during the first four minutes (a longer stirring time was required to

118

avoid phase inversion). After preparation, the emulsions are poured into 100-ml glass flask. In all

119

cases, with exception of the autoclave experiments, 30 ml of emulsion samples are prepared. One

120

liter of emulsions samples is prepared for the autoclave experiments.

121 122

2.2. Interfacial Tension Measurements

123

Water-mineral oil interfacial tension (IFT) is measured at ambient temperature and pressure

124

through a pendant drop tensiometer with a KSV Instruments, Ltd. equipped with a CAM 200

125

apparatus. IFT tests are performed on systems with different surfactant concentrations in the oil

126

phase. The data of each experiment are collected every 10 s for 290 s. This total time is sufficient

127

to reach an equilibrium value of the interfacial tension, which is calculated by the tensiometer

128

software. The density of both fluids (water and oil) that are required as input in the calculations

129

of the interfacial tension are experimentally measured using an Anton Paar DMS60 densitometer

130

connected to an Anton Paar DMA602HT measuring cell.

131 132

2.3. Emulsion Characterization

133

2.3.1. Stability Test

134

Phase separation is monitored daily by visual observations (classical bottle test) for a period of

135

one week. This test is performed at room temperature (~22°C) and 2°C.

ACS Paragon Plus Environment

7

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 34

136

2.3.2. Droplet Sizes Measurements

137

Droplet size measurements are performed using an optical microscope (Olympus IX71)

138

coupled to a digital camera (Olympus XM10). Several images of the emulsions are recorded over

139

the course of one week and analyzed using ImageJ. In measuring the droplet size, an average of

140

about 250 water droplets is measured for each water cut. The reported mean droplet size is then

141

the numerical average of the measured water droplets.

142 143

2.3.3. Viscosity measurements

144

With the purpose of observing the relationship between viscosity and water cut, viscosity

145

measurements are performed at ambient pressure on the emulsions using a TA Instrument

146

Hybrid Rheometer DHR-3. In this study, cone and plate geometry is used with a gap size of 52

147

μm. The cone has a diameter of 40 mm and an angle of 1.995°. Viscosity measurements are

148

conducted at room temperature (~22°C) and 1°C. To determine the stability of the emulsions

149

over time, viscosity measurements are conducted at three different times: 24 hours, 72 hours, and

150

one week after preparation.

151

A sample of ~580 μl of the emulsion is used in all measurements. The sample is allowed to

152

rest at the experimental temperature for two minutes prior to the measurement. Measurements

153

are performed at shear rates ranging from 0.1 to 1000 s-1. In each test, ten viscosity points per

154

decade on a log scale were taken and each point at steady-state, which we define as three

155

consecutive points within 3% variation.

156 157

2.4. Emulsion Behavior under Hydrate Formation

158

2.4.1. Differential Scanning Calorimetry (DSC)

ACS Paragon Plus Environment

8

Page 9 of 34

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

159

To measure the thermal properties of the emulsions, DSC tests at ambient pressure and high

160

pressure are conducted using a micro-differential scanning calorimeter (μ-DSC VIIa, Setaram

161

Inc.). For ambient pressure DSC tests, ~15 mg of an emulsion sample is placed in a vessel. The

162

sample is cooled from 30°C to -45°C at a rate of 0.3°C/min. The sample is then heated back to

163

30°C at the same rate. This cycle is repeated three times to observe the effect of ice formation on

164

emulsion stability. In the high pressure tests, a similar amount of emulsion sample is added into

165

the vessel. Methane gas (99.99% Ultra High Purity, General Air) is then introduced into the

166

vessel with the sample up to a pressure of 125 bar. The sample is left at 125 bar and 30°C for

167

three hours to allow for saturation of the sample with methane. Next, the sample is cooled from

168

30°C to -25°C at a rate of 0.3°C/min and then heated at a rate of 0.3°C/min to 30°C. Similarly,

169

the cycle is repeated three times to observe the effect of hydrate formation on emulsion stability.

170 171

2.4.2. Micromechanical Force Measurements

172

A micromechanical force (MMF) apparatus is used to measure the cohesion forces between

173

cyclopentane (CyC5, 99% by Acros Organic) hydrates particles in CyC5/mineral oil and

174

CyC5/surfactant/mineral oil systems. A detailed description of MMF apparatus and experimental

175

method are reported elsewhere.19-22 Briefly, a hydrate particle is attached to a spring constant-

176

calibrated glass cantilever and another hydrate particle at the tip of a glass fiber attached to a

177

moving micromanipulator. The cantilevers and hydrate particles are immersed in a cooled CyC5

178

bath at 3.2°C and atmospheric pressure. Forty “pull-off” measurements are conducted, where a

179

“pull-off” corresponds to the contact of the particles, a 10 seconds waiting period, and the pull-

180

off of the particles until they are no longer in contact. The cohesive force (FA) between the

ACS Paragon Plus Environment

9

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 34

181

particles is obtained from Hooke’s Law on the displacement of the cantilever and the cantilever

182

spring constant. The reported cohesive force is the average of forty “pull-off” trials.

183

With the aim of studying the effect of Span 80-AOT surfactant mixtures on the hydrate

184

cohesion forces, two experiments are performed. The first system consisted of 97 wt.% CyC5

185

and 3 wt.% of Mineral Oil 70T. The second system consists of 98 wt.% CyC5 and 2 wt.%

186

Mineral Oil 70T with 5 wt.% surfactant mixture.

187 188

2.4.3. Autoclave Experiments

189

To obtain information about hydrate formation under dynamic conditions of the model

190

emulsion, high-pressure autoclave cell tests are performed. A complete description of autoclave

191

components and operation may be found in the literature.7,8,10,23 The autoclave is a 2-liter stainless

192

steel vessel nine inches in height and four inches inner diameter. Mixing in the cell is done with

193

a four-blade vane impeller, each blade at 90 degrees, 5.5 inches in height and 0.9 inches in width.

194

One liter of emulsion with a 30 vol.% water cut is charged into the autoclave cell. Once sealed,

195

the system is placed in a cooling bath that is initially set at 20°C. Next, the system is stirred at

196

constant 300 rpm. The cell is then pressurized with methane gas (99.99% Ultra High Purity,

197

General Air) to 65.5 barg (950 psig). The system is kept at these conditions for 12 hours to

198

ensure the emulsion is fully saturated with methane. Following this step, the cell is cooled down

199

to 1°C. Throughout the test, the pressure in the cell is maintained at 65.5 barg by feeding the cell

200

with methane gas from a secondary gas reservoir tank. The time when the pressure in the

201

secondary reservoir started to decrease is taken as onset of hydrate formation. The current

202

required by the impeller motor to maintain the stirring at 300 rpm is a qualitative measurement of

203

the relative viscosity of the fluids/slurry in the autoclave cell. The duration of each test is at least

ACS Paragon Plus Environment

10

Page 11 of 34

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

204

24 hours after hydrate formation. The amount of hydrate formed is estimated from the

205

consumption of gas in the reservoir.

206 207

3. RESULTS AND DISCUSSION

208

3.1. Emulsion Stability and Droplet Size

209

Visual observation confirmed that all emulsions (1 and 5 wt.% surfactant) were stable without

210

evidence for droplet coalescence (see Figure 3), that is, no free water observed after one week.

211

However, at low temperature (2°C), emulsions prepared with 10 and 20 vol.% water cut showed

212

settling of water droplets (sedimentation) identified by the presence of a thin oil layer at the top

213

of the samples. This settling of the droplets was observed after 24 hours the samples were placed

214

in a chiller at 2°C. The samples became re-dispersed by daily gentle hand-stirring thus helping

215

the emulsion to remain homogeneous. For our purposes, the settling is not a major concern since

216

these emulsions will be later used in experiments where mixing is typically constantly applied

217

(e.g., rheometer, autoclave). It should be noted that mixtures prepared with less than 1 wt.%

218

surfactant did not result in stable emulsions, resulting in phase separation within 24 hours after

219

preparation. Measurements of water droplets at different times in the studies shows that

220

coalescence was negligible even for samples that showed sedimentation of water droplets (see

221

Supporting Information for droplet sizes measured over days). For these emulsions, increasing

222

the water cut reduces the tendency for sedimentation while the stability of the emulsion is

223

maintained by the presence of the two surfactants.

ACS Paragon Plus Environment

11

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 34

224

225

226

227 228

Figure 3. Images of water-in-oil emulsions for a range of water cut prepared with Mineral Oil

229

70T (continuous phase) and a surfactant mixture of Span 80 + AOT: 1 wt.% surfactant after a)

230

24 hours and b) one week, and 5 wt.% surfactant at c) 24 hours and d) one week.

231

The high stability of the emulsions can be attributed to the strong lipophilic character of the

232

surfactant mixture. Additionally, one may suggest that there are two mechanisms of dispersion

233

stabilization:24,25 steric repulsions produced by the interaction between hydrocarbon tails present

ACS Paragon Plus Environment

12

Page 13 of 34

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

234

in both surfactant molecules and electrical repulsions due to the negative charge present in the

235

anionic group (SO3-) in AOT molecules. The application of surfactant mixtures to stabilize

236

emulsions has been demonstrated to be in most cases more efficient than a single surfactant,

237

owing to synergetic mechanisms that reduce coalescence of droplets.26,27 Furthermore, due to the

238

small size of water droplets, sedimentation of water droplets is reduced.

239

The concentration of the surfactant (1 and 5 wt.%) used in all the emulsions prepared for this

240

study was higher than the critical concentration of aggregation (CCA), that is, the concentration

241

at which inverse micelles may form. Figure 4a shows the interfacial tension values for the

242

mineral oil-water system at different surfactant concentrations. A change in the slope of the IFT

243

curve represents the surfactant aggregation threshold. According to Figure 4a, the CCA of the

244

surfactant mixture in mineral oil is approximately 0.1 wt.%. All emulsions were prepared using

245

either 1 or 5 wt.% of surfactant (Span 80-AOT) with respect to the total mass of the emulsion.

246

The concentration of the surfactants above the CCA in the oil phase guarantees that once the

247

emulsion is created, the new interface is quickly covered by adsorbed surfactant molecules; the

248

barrier against coalescence is instantly present and the coalescence rate decreases.28,29

249

The IFT as a function of time are showing Figure 4b. It can be appreciated that IFT

250

equilibrium values were reached in short times (lesser than 290 s), which is a clear indication of

251

the rapid adsorption of the surfactant at the interface. It is remarkable that, for surfactant

252

concentrations beyond the CCA, equilibrium was attained at even shorter times (< 50 s). This

253

higher adsorption rate can lead to the quick droplet coverage by the surfactant mixture and as

254

consequence, a barrier against coalescence is consolidated during emulsification time (3 min).29,

255

30

ACS Paragon Plus Environment

13

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 34

256 257

Figure 4. Interfacial tension as a) a function of surfactant concentration and b) dependence on

258

concentration and time.

259

The mean droplet size for the emulsions with both 1 and 5 wt.% surfactant is in the range of 2

260

to 3 µm in diameter at room temperature and 2°C, for all water cuts studied in this work. These

261

observations are in good agreement with other reported data31 for water-in-crude oil emulsions.

262

In all cases, the droplet size is relatively similar for all water cut (see Figure S1 in the Supporting

263

Information material). We also observe that the variation in the droplet size for emulsions with 5

264

wt.% surfactant is smaller than those with 1 wt.% surfactant, reinforcing the high stability of the

265

emulsion with higher surfactant concentration (see Supporting Information material for

266

additional results on measures of droplet sizes over the one week testing period). Figure 5 shows

267

examples of microscopy images on the appearance of the emulsions at 10% and 50% water cuts

268

immediately after their preparation for emulsion prepared using 1 and 5 wt.% surfactant.

ACS Paragon Plus Environment

14

Page 15 of 34

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

269 270

Figure 5. Sample microscope images of water-in-oil emulsions for 1 wt.% surfactant with a) 10

271

water vol.% and b) 50 water vol.%, and 5 wt.% surfactant with c) 10 water vol.% and d) 50

272

water vol.%.

273

3.2. Viscosity Results

274

Viscosity profiles for the emulsions prepared with 1 and 5 wt.% surfactant measured at

275

ambient pressure for two temperatures (room temperature and 1°C) are presented in Figure 6. As

276

can be seen, the apparent viscosity of the emulsions increases significantly with the water cut.

277

This increase in apparent viscosity is due to the increase in the packing of the water droplets.27

278

Since the viscosity of the emulsion is proportional to the viscosity of the external phase, the

279

viscosity of the emulsion will deviate further from the viscosity of Mineral Oil 70T as the water

280

cut increases. This observation holds true at both temperatures tested.

ACS Paragon Plus Environment

15

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 34

281

Furthermore, shear-thinning behavior was observed above certain water cuts: above 30 vol.%

282

at room temperature and 20 vol.% at 1°C for both surfactant concentrations. This behavior is

283

typical for complex dispersed systems such as human blood, food emulsions and even water-in-

284

crude oil emulsions,32,33 and can be attributed to, in emulsion cases, the rupture of clusters or

285

droplet aggregates, followed by the droplets alignment under shear.34 Thus the shear-thinning

286

effect is more significant at high water cuts and low temperatures because the system becomes

287

highly packed and the droplet-droplet interactions increase.

288

289

ACS Paragon Plus Environment

16

Page 17 of 34

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

290

Figure 6. Viscosity profiles at various water cuts for emulsion prepared with 1 wt.% of

291

surfactant at a) 22°C and b) 1°C, and with 5 wt.% of surfactant at c) 22°C and d) 1°C.

292

Furthermore, the viscosity profiles at room temperature (~22°C) shows that for the 10 and 20

293

vol.% water cut, the viscosity of the emulsion is relatively similar for both surfactant

294

concentrations. However, at 30 and 40 vol.% water cut, the viscosity of the emulsion is greater

295

for system with 1 wt.% surfactant and at water cut greater that 50 vol.%, the viscosity of the

296

emulsion with 1 wt.% of surfactant is lower than that of system with 5 wt.% of surfactant at the

297

same water cut. The viscosity is related to the size and packing (concentration) of water droplet

298

in the emulsion and it is known that the increase in droplet-droplet interactions increases the

299

viscosity of the emulsion.24 We also see that the viscosity of the emulsions with 1 wt.%

300

surfactant (Figure 6b) is slightly lower than the emulsions with 5 wt.% surfactant (Figure 6d) for

301

all water cuts studied at 1°C.

302

Moreover, in Figures 6b and 6d, it can be noted that for the sample with 70 vol.% water cut,

303

the viscosity of the emulsions drops significantly at a high shear rate. This suggests a possible

304

phase inversion of the emulsion. To confirm this phenomenon, a small sample of the emulsion

305

was stirred at a high shear rate of 1000 s-1 for two minutes. Next, a small drop of the sample was

306

placed into water. As the emulsion completely dispersed in water, it was confirmed that water

307

was the continuous phase.35 Emulsion inversion induced by stirring has been well studied.34 In

308

systems with high internal phase content, physicochemical formulation leads to important

309

changes in emulsion morphology.34 This experimental evidence suggests that hydrates studies

310

using the model emulsion proposed here, will have to consider possible phase inversion for

311

samples with high water cut and experiments with high shear rate.

ACS Paragon Plus Environment

17

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 34

312

The viscosity of the emulsions was measured at different times to determine their stability.

313

Since the viscosity of the emulsion is proportional to the water cut, any instability in the

314

emulsion (phase separation or sedimentation) will be captured by a large deviation in the

315

viscosity measurements. Figure S4a in the Supporting Information shows the apparent viscosity

316

of 30 vol.% water cut emulsion with 1 wt.% surfactant. As can be noticed, there is a noticeable

317

change in the viscosity of the emulsion over the testing time. This is an indication that

318

sedimentation and phase inversion might occur in the sample but was not detected by visual

319

observation. In contrast, the viscosity of a 30 vol.% water cut emulsion prepared with 5 wt.%

320

surfactant shows a relatively small changes in the viscosity over the testing time (Figure S4b in

321

the Supporting Information). A small change in the viscosity of the emulsion indicates that the

322

emulsions prepared with 5 wt.% surfactant remain stable for at least one week, thus making them

323

particularly suitable for our intended hydrate studies. It can also be concluded that the emulsion

324

prepared using 1 wt.% surfactant is less stable as compared to emulsion prepared using 5 wt.%

325

surfactant.

326 327

3.3. DSC Results

328

DSC tests of the emulsions at ambient pressure and high pressure were conducted to determine

329

their stability upon ice and hydrate formations. Figure 7a shows the thermogram for a 10 vol.%

330

water cut emulsion with 1 wt.% surfactant. Cycle 1, there is a small peak at ~ -18°C and another

331

at ~-42°C upon cooling. In cycle 2, two exothermic peaks are seen upon cooling, one at ~ -19°C

332

and another at ~ -42°C. In cycle 3, three exothermic peaks are evident upon cooling: one at ~ -

333

13°C, another at ~ -18°C, and the third at ~ -42°C . The peaks at ~ -42°C in all cycles confirm

334

that the samples are a water-in-oil emulsion.9,36-38 Since there are at least two exothermic peaks

ACS Paragon Plus Environment

18

Page 19 of 34

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

335

upon cooling, it is likely that the emulsion with 1 wt.% surfactant has at least two water droplet

336

sizes. It can also be noticed that the intensity of the peaks at ~ -42°C decrease with each cycle,

337

indicating that the amount of small water droplets decreases with each cycle. This suggests that

338

that agglomeration of the water droplets may be occurring upon ice formation and melting. In the

339

heating portion of the thermogram, only one endothermic peak near 0°C is observed,

340

corresponding to ice melting.

341

342

ACS Paragon Plus Environment

19

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 20 of 34

343

Figure 7. Ambient pressure thermograms of the emulsions prepared with 1 wt.% surfactant for

344

a) 10 vol.% and b) 50 vol.% water cut, and with 5 wt.% surfactant for c) 10 vol.% and d) 50

345

vol.% water cut. For each plot: cycle 1 (red), cycle 2 (blue), cycle 3 (green).

346

Figure 7b shows the thermogram for 50 vol.% water cut emulsion with 1 wt.% surfactant at

347

ambient pressure. It can be immediately noticed that there is only one sharp peak at ~ -40°C

348

upon cooling during cycle 1, confirming the sample is a water-in-oil emulsion.9,36-38 However, for

349

cycles 2 and 3, several exothermic peaks are seen upon cooling. These peaks are low and broad,

350

suggesting that the water droplets are not well dispersed. Also, because the intensity of the peak

351

at ~ -40°C significantly decreases with each cycle, water droplet agglomeration/coalescence is

352

likely occurring upon ice formation/melting. In the heating portion of the thermogram, there is

353

only one endothermic peak correspond to ice melting. These results suggest that ice formation

354

and melting destabilize this emulsion.

355

Figure 7c shows the thermograms for the emulsion with 5 wt.% surfactant and 10 vol.% water

356

cut at ambient pressure. There are two exothermic peaks upon cooling: one at ~ -18°C in cycles 1

357

and 2 and ~ -19°C in cycle 3, and another at ~ -42°C. The latter peak again confirms the sample

358

is a water-in-oil emulsion with small droplets.9,36-38 The near symmetrical shape of the peak at ~ -

359

42°C indicates the water droplets are well dispersed. The other small peaks at high temperature,

360

it may be from ice formation from bulk water or from freezing of aggregates of small water

361

droplets.37 However, from the low intensity of the peak, only a small amount of free water or

362

large water aggregates is present in the sample. In the heating portion of the thermogram, there is

363

only one endothermic peak near 0°C, corresponding to ice melting. The peaks observed in the

364

cooling and heating stages for all three cycles remain relatively unchanged, indicating the

365

emulsion remains stable upon ice formation and melting.

ACS Paragon Plus Environment

20

Page 21 of 34

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

366

Figure 7d shows the thermogram for the emulsion with 5 wt.% surfactant and 50 vol.% water

367

cut at ambient pressure. Several exothermic peaks are observed upon cooling, ones clustered

368

near ~ -20°C and another at ~ -40°C. Similarly, the peak at ~ -40°C confirms a water-in-oil

369

emulsion with small droplets. Since the location of the peak is at a slightly higher temperature, it

370

may indicate that the water droplets are slightly larger in this 50 vol.% water cut sample

371

compared to the one with 10 vol.% (Figure 7c). The other peaks at ~ -20°C suggest the presence

372

of bulk water or isolated large water droplets. For this water cut, it can also be observed that the

373

intensity of the peaks also are relatively unchanged from cycle to cycle, confirming that the

374

emulsion remains stable upon ice formation and melting.

375

Measurements in the DSC were also made to determine the stability of the emulsions upon

376

hydrate formation at 125 bar with methane gas. Figure 8a is the high pressure thermogram for 10

377

vol.% water cut emulsion with 1 wt.% surfactant. In all cycles, there are several exothermic

378

peaks upon cooling, corresponding to both hydrate and ice formations. The heating portion of the

379

thermogram shows ice melting (~0°C) and hydrate dissociation (~15°C). The reported

380

dissociation temperature for methane hydrate at 125 bar is 14.82°C.1 It is seen that the intensity

381

of the ice melting peak slightly increases while the intensity of the hydrate dissociation peak

382

slightly decreases progressively with each cycle, that is, the amount of ice formed increases

383

while the amount of hydrate formed decreases with each cycle. These results indicate that there

384

is likely some agglomeration of the water droplet upon hydrate formation/dissociation, resulting

385

in either larger water droplets or droplet coalescence.

386

Figure 8b shows the thermogram at high pressure for a 50 vol% water cut emulsion with 1

387

wt.% surfactant. The general features of the thermogram are very similar to that in Figure 8a (10

388

vol.% water cut emulsion with 1 wt.% surfactant), however, the ice melting peaks are much

ACS Paragon Plus Environment

21

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 34

389

more intense, indicating there is a much larger amount of ice being formed after the first cycle at

390

higher water cut. For this water cut, agglomeration of water droplet is much more significant

391

upon hydrate formation/dissociation.

392

393 394

Figure 8. High pressure thermogram of the emulsions prepared with 1 wt.% surfactant for a) 10

395

vol.% and b) 50 vol.% water cut, and with 5 wt.% of surfactant for c) 10 vol.% and 50 vol.%

396

water cut: cycle 1(red), cycle 2 (blue), cycle 3 (green).

ACS Paragon Plus Environment

22

Page 23 of 34

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

Figure 8c shows high pressure thermogram results for a 10 vol.% water cut emulsion prepared

398

with 5 wt.% surfactant. For this emulsion, it is observed that only hydrates are

399

formed/dissociated, as evidenced by the endothermic peak at ~15°C. The absence of the ice

400

melting peak and nearly constant intensity of the hydrate dissociation peak show this emulsion is

401

highly stable. It should be noted that hydrate formation is stochastic so the formation peaks upon

402

cooling should not be expected to be reproducible from cycle to cycle.

403

A much different behavior is observed in Figure 8d for a 50 vol.% water cut emulsion with 5

404

wt.% surfactant. Only one exothermic peak is measured upon cooling in each cycle. On heating

405

in the first cycle, only one endothermic peak at ~15°C is seen, corresponding to hydrate

406

dissociation. In cycles 2 and 3, an ice melting peak is observed, accompanied by a less intense

407

hydrate

408

agglomeration/coalescence of water droplets occur upon hydrate formation/dissociation.9

409

Comparing Figure 8d to Figure 7d, it can be concluded that this emulsion is highly stable to ice

410

formation/dissociation but less so to hydrate formation/dissociation.

dissociation

peak.

These

results

suggest

that

after

the

first

cycle,

411

The DSC results are insightful in providing an understanding on the stability of the emulsions,

412

especially with respect to ice and hydrate formation/dissociation. As the water cut increases, the

413

emulsion becomes more packed with the water droplets. The increase in droplet interactions can

414

cause droplet coalescence and thus lower the emulsion stability. We see from these results that

415

the emulsions at low water cut and with 5 wt.% surfactant are the most stable and may be the

416

best ones for our later studies with hydrates. Additional results from high pressure DSC

417

measurements of the emulsions at 20, 30, and 40 vol.% water cut can be found in the Supporting

418

Information.

419

ACS Paragon Plus Environment

23

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 34

420

3.4. MMF Results

421

All of the emulsion systems were prepared using surfactant concentration greater than the

422

CCA. It is thus necessary to measure how the surfactants impact the interaction of hydrate

423

particles. For a system with 3 wt.% of Mineral Oil 70T in cyclopentane, the average interparticle

424

cohesion force was measured to be 4.1±1.9 mN/m. This result is consistent with that measured

425

by Aman et al.39 of 6.1±1.3 mN/m based on Mineral Oil 200T. The presence of 5 wt.% surfactant

426

mixture in the Mineral Oil 70T results in an interparticle cohesion force of 2.1±0.8 mN/m.

427

Comparing the two systems with Mineral Oil 70T, the difference in the cohesion force is not

428

statistically significant, even though there is a change in the morphology of the hydrate particles.

429

As can be seen in Figure 2, both Span 80 and AOT have hydrophilic and hydrophobic groups.

430

The hydrophilic part of the surfactant likely attaches to the surface of the hydrate particles,

431

causing the hydrophobic part to extend toward the oil phase.39

432

The morphological changes observed are likely induced by Span 80.40 Several previous studies

433

have reported similar morphological changes when Span-type surfactants are used.40,41 Figure 9

434

shows the observed changes on the hydrate particles with the introduction of the surfactants used

435

in preparing the emulsion. The increase in the surface roughness on the particle can lead to a

436

decrease in capillary water bridges between hydrate particles and a slight decrease in the

437

cohesion force. The presence of the surfactant mixture (Span 80 and AOT) does not have a

438

significant impact in the cohesion force between cyclopentane hydrate particles. It has been

439

reported that other surface active chemicals such as naphthenic acid can reduce the cohesion

440

force by as much as 80%.21 Therefore, the surfactant mixture used in the emulsions is suitable for

441

hydrate studies.

442

ACS Paragon Plus Environment

24

Page 25 of 34

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

443 444

Figure 9. Morphological changes observed when the surfactant mixture was added into the

445

MMF for the cohesive force measurements: a) system without surfactant, and b) system with

446

addition of 5 wt.% surfactant mixture.

447 448

3.4. Autoclave Test Results

449

Contrasting results are observed for the tests performed in the autoclave cell experiments with

450

hydrate formation for model emulsion and West African Crude (WAC) emulsion compared with

451

the system without surfactant. For the sample without surfactants (mineral oil + water), large

452

hydrate chunks were seen at the end of experiment upon opening the autoclave cell (Figure 10a).

453

For the model emulsion and WAC system, small particles of loose hydrate slurry were obtained

454

(Figure 10b and 10c). In the case of mineral oil + water (no surfactant), the hydrates formed

455

quickly agglomerate, forming a solid mass of hydrates. In the model emulsion, the water droplets

456

in the emulsion are quickly converted to hydrate, and because of the emulsion stability, the

457

hydrate particles remain dispersed and create a slurry. Similar results are seen for WAC

458

emulsion, which contains natural surface-active component, creating a loose hydrate slurry.

459

Another useful quantity from the autoclave cell test is the motor current (measure of relative

460

viscosity) required to maintain the stirrer at 300 rpm. Figure 11a shows the motor current over

ACS Paragon Plus Environment

25

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 34

461

the course of the test after hydrate formation for all three systems tested. As clearly seen in the

462

Figure 11a, the model emulsion shows lower motor current requirement than the mineral oil +

463

water system and WAC emulsion system, suggesting a lower relative viscosity. This result is

464

again related to the dispersion of the hydrate particles formed in the systems with surfactant,

465

whereas large hydrate chunks in the mineral oil + water system resulted in relative high viscosity

466

conditions. In comparing the model emulsion with WAC emulsion, it can be observed that the

467

motor current for WAC system is slightly higher than that of model emulsion. This is as

468

expected as the viscosity of the WAC at 25°C is 60 cP42 while the viscosity of Mineral Oil 70T at

469

the same temperature is 20 cP. However, the motor current throughout the time of the

470

experiment for model emulsion has the same general trend to that of WAC emulsion, suggesting

471

the model emulsion has a similar behavior as the WAC emulsion.

472

473 474

Figure 10. Type of hydrate particles observed for a) mineral oil + water system b) model

475

emulsion system and c) WAC emulsion at the end of high pressure autoclave experiment.

ACS Paragon Plus Environment

26

Page 27 of 34

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

476 477

Figure 11. a) Relative power input for the stirrer and b) hydrate fraction in the autoclave as a

478

function of time. Curves are for mineral oil + water (red), model emulsion (blue) and WAC

479

emulsion (orange).

480

It is also interesting to compare the amount of hydrate formed for the systems tested in the

481

autoclave cell. As shown in Figure 11b, a similar amount of hydrates formed in all systems by

482

the end of the experiment. However, based on the slope of the curves, hydrates formed at a faster

483

rate in the model emulsion followed by WAC emulsion and finally the mineral oil + water

484

mixture. This result can be interpreted based on size of the water droplets in the emulsion. The

485

model emulsion is highly stable with water droplets ranging 2-3 μm, while WAC emulsion has

486

water droplet ranging 3-5 μm. As for the mineral oil + water mixture, it does not form a stable

487

emulsion and any entrainments of the phases are quickly separated. The small water droplet size

488

in the emulsions creates a high interfacial surface area to form hydrates, which it is not the case

489

in the mineral oil + water mixture, where the interfacial area is only generated from the mixing

490

imposed by the stirrer. Since water droplets are smaller (larger surface area) in the model

491

emulsion, hydrates formation is the fastest among the three systems.

ACS Paragon Plus Environment

27

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 34

492 493

4. CONCLUSIONS

494

Stable water-in-oil emulsions with water cut ranges from 10 to 70 vol.% have been prepared

495

using Mineral Oil 70T, Span 80, sodium di-2-ethylhexylsulfosuccinate (AOT) and water.

496

Classical bottle tests conducted show that all emulsions, both with 1 and 5 wt.% surfactant

497

mixture) are stable against coalescence for at least one week (the period of study). This high

498

stability is attributed to the strong lipophilic interaction among the surfactants molecules. From

499

the microscopy images of the emulsions, the average water droplet size ranges from 2 to 3 μm.

500

Rheological measurements performed on these emulsions show a strong dependency of the

501

viscosity on the water cut. The viscosity of the emulsion increases with increasing water cut. It

502

was also observed that these emulsions show shear thinning behavior above certain water cuts

503

(30 vol.% at 25°C and 20 vol.% at 1°C). This behavior is similar to water-in-oil emulsions

504

prepared with a crude oil.9 The DSC measurements show that the emulsions prepared with 5

505

wt.% surfactant with less than 50 vol.% water are the most stable upon ice and hydrate

506

formation/dissociation. For samples with high water cut (> 50 vol.%), the emulsions are slightly

507

less stable after ice and hydrate formation/dissociation due to the increased packing density of

508

water droplets and their interaction. MMF measurements show that the presence of the surfactant

509

mixture has little impact in the cohesion force of hydrate particles, although a change in hydrate

510

surface morphology was observed with the surfactant added. Autoclave cell experiments showed

511

two different types of hydrate formation: large hydrate chunks for the system without surfactant

512

and a loose hydrate slurry for the system with surfactant. It is also observed that the relative

513

viscosity of the hydrate slurry is lower and the formation rate is higher for the emulsion

514

compared to the mineral oil + water system. Autoclave cell experiments also shows that the

ACS Paragon Plus Environment

28

Page 29 of 34

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

515

model emulsion system behaves similarly to a crude oil emulsion. The tests performed in this

516

study demonstrate that the proposed model emulsion is stable, having similar characteristic to

517

that observed in crude oil emulsions, and may be suitable for other hydrate studies requiring

518

long-term (~one week) stability.

519 520

SUPPORTING INFORMATION

521

Average water droplet size for emulsion over the range of water cuts, average droplet size for

522

emulsions over the one-week testing period, viscosity of emulsions over time, and high pressure

523

DSC thermograms for emulsions with 20, 30, and 40 vol.% water cut. This material is available

524

free of charge via the Internet at http://pubs.acs.org.

525 526 527 528

AUTHOR INFORMATION *Corresponding Author: Telephone: +1(303) 273-3873. Fax: +1(303) 273-3730. E-mail: [email protected]

529 530

ACKNOWLEDGMENT

531

The authors acknowledge the funding and support from the Hydrate Consortium (current and

532

past members): BP, Champion Technologies, Chevron, ConocoPhillips, ENI, ExxonMobil,

533

Halliburton, Multi-Chem, Nalco, Petrobras, Schlumberger, Shell, SPT Group, Statoil, Total. We

534

also thank Dr. Jefferson Creek (Chevron ETC) for kindly providing the analysis of the mineral

535

oil used in this study.

536

ACS Paragon Plus Environment

29

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

537

REFERENCES

538

(1)

539 540

Page 30 of 34

Sloan, E.D.; Koh, C.A. Clathrate Hydrates of Natural Gases. 3rd ed. CRC Press: Boca Raton, FL, 2008.

(2)

541

Natural Gas Hydrates in Flow Assurance; Sloan D., Koh, C., Sum, A.K., Eds.; Elsevier: Amsterdam, The Netherlands, 2010.

542

(3)

Ballard A. L.; Sloan E. D. Fluid Phase Equilib. 2004, 218, 15–31.

543

(4)

Sum A. K.; Koh C.;Sloan E. D. Ind. Eng. Chem. Res 2009, 48, 7457-7465.

544

(5)

Sum A. K.; Koh C.; Sloan E. D. Energy Fuels 2012, 26, 4046-4052.

545

(6)

Turner, D. Clathrate Hydrate Formation in Water-in-Oil Dispersions. Ph.D. Dissertation,

546

Colorado School of Mines, Golden, CO, 2005.

547

(7)

Taylor C.; Miller K.; Koh C.; Sloan D. Chem. Eng. Sci. 2007, 62, 6524-6533.

548

(8)

Greaves D.; Boxall J.; Mulligan J.; Sloan D.; Koh C. Chem. Eng. Sci. 2008, 63, 4570-

549

4579.

550

(9)

Lachance J.; Sloan E. D.; Koh C. Chem. Eng. Sci. 2008, 63, 3942-3947.

551

(10) Sjöblom J.; Øvrevoll B.; Jentoft G.; Lesaint C.; Palermo T.; Sinquin A.; Gateau P., Barre´

552

L.; Subramanian S.; Boxall J.; Davies S.; Dieker L.; Greaves D.; Lachance J.; Rensing P.;

553

Miller K.; Sloan D.; Koh C. J. Dispersion Sci. Technol. 2010, 31, 1110-1119.

554

(11) Fadnes F. H. Fluid Phase Equilib. 1996, 117, 186-192.

555

(12) Salager J.; Forgiarini A. Energy Fuels. 2012, 26, 4027-4033.

ACS Paragon Plus Environment

30

Page 31 of 34

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

556 557

Energy & Fuels

(13) Webb E.; Rensing P.; Koh C.; Sloan E. D.; Sum A. K.; Liberatore M. Energy Fuels. 2012, 26, 3504-3509.

558

(14) Bancroft W.D. J. Phys. Chem. 1913, 17, 501-519.

559

(15) Bancroft W.D. Phys. Chem. 1915, 19, 275-309.

560

(16) Griffin W. C. J Soc. Cosmet. Chem. 1949, 1, 311-326.

561

(17) Griffin, W. C. J Soc. Cosmet. Chem. 1954, 5, 249-256.

562

(18) The HLB System a Time-Saving Guide to Emulsifier Selection. ICI Americas Inc., Edited

563 564 565 566 567

by CHEMMUNIQUE. (19) Aman Z. M. Interfacial Phenomena of Cyclopentane Hydrate. Ph.D. Dissertation, Colorado School of Mines, Golden, CO, 2012. (20) Aman Z. M.; Brown E. P.; Sloan E. D.; Sum A. K.; Koh C. A. Phys. Chem. Chem. Phys. 2011, 13, 19796−19806.

568

(21) Aman Z. M.; Sloan E. D.; Sum A. K.; Koh C. A. Energy Fuels 2012, 26, 5102-5108.

569

(22) Dieker, L. E., Cyclopentane Hydrate Interparticle Adhesion Force Measurements. M.S.

570

Thesis, Colorado School of Mines, Golden, CO, 2009.

571

(23) Turner D.; Miller K.; Sloan E.D. Chem. Eng. Sci. 2009, 64, 3996-4004.

572

(24) Salager J.-L. Pharmaceutical Emulsions and Suspensions. Emulsion Properties and

573

Related Know-How to Attain Them, Françoise Nielloud, Gilberte Marti-Mestres (Eds.),

574

Marcel Dekker, Inc., New York, 2000; pp 73-125

ACS Paragon Plus Environment

31

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

575 576 577 578 579 580

Page 32 of 34

(25) Walstra P. Encyclopedia of Emulsion Technology. Emulsion Stability, Paul Becher (Ed.), Marcel Dekker, New York, 1996; pp 1-62.

(26) Boyd J.; Parkinson C.; Sherman P. Factors Affecting Emulsion Stability, and the HLB Concept, J. Colloid Interface Sci. 1972, 41, 359-370. (27) Tadros T. Applied Surfactants. Principles and Applications. WILEY-VCH Verlag GmbH & Co. Weinheim, 2005.

581

(28) Salager J.-L., Perez-Sanchez M., Ramirez-Gouveia M., Anderez J., Briceno-Rivas M. I.,

582

Stirring-Formulation Coupling in Emulsification. IXth European Congress on Mixing,

583

Paris, 122-130, 1997.

584 585

(29) Walstra P. Encyclopedia of Emulsion Technology. Formation of Emulsions, Vol. 1 Basic Theory, Paul Becher (Ed.), Marcel Dekker, New York, 1983; pp 57-127.

586

(30) Ford R. E.; Furmidge C. G. L. Studies at Phase Interfaces II. The Stabilization of Water-

587

in-Oil Emulsions Using Oil-Soluble Emulsifiers. J. Colloid Interface Sci. 1966, 22, 331-

588

341.

589 590

(31) Nöik C.; Trapy J., Mouret A. Design of a Crude Oil Dehydration Unit. SPE 77492, SPE Annual Technical Conference and Exhibition, San Antonio, TX, 2002.

591

(32) Briceño M. I. Pharmaceutical Emulsions and Suspensions. Rheology of Suspensions and

592

Emulsions, Françoise Nielloud, Gilberte Marti-Mestres (Eds.), Marcel Dekker, Inc., New

593

York, 2000; pp 557-607.

ACS Paragon Plus Environment

32

Page 33 of 34

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

594 595 596 597

Energy & Fuels

(33) Quintero C. Comportement Rhéologique et Propriétés Interfaciales des Systèmes Émulsionnés d'Intérêt Industriel. Ph.D. Dissertation, Université Paris-Sud 11, Paris, 2008. (34) Rondón-González M.; Madariaga L. F.; Sadtler V.; Choplin L.; Salager J.-L. Ind. Eng. Chem. Res. 2009, 48, 2913-2919.

598

(35) Ese M.H.; Kilpatrick P.K. J. Dispersion Sci. Technol. 2004, 25, 253-261.

599

(36) Dalmazzone C.; Noïk C.; Clausse D. Oil & Gas Sci. Technol – Rev. IFP. 2009, 64, 543-

600 601 602 603 604 605 606

555. (37) Santini E.; Liggieri L.; Sacca L.; Clausse D.; Ravera F. Colloids Surf., A. 2007, 309, 270279. (38) Lachance J.W. Investigation of Gas Hydrates Using Differential Scanning Calorimetry with Water-In-Oil Emulsion. M.S Thesis, Colorado School of Mines, Golden, CO, 2008. (39) Aman Z. M.; Dieker L.E., Aspenes G.; Sum A. K.; Sloan E. D.; Koh C. A. Energy Fuels 2010, 24, 5441-5445.

607

(40) Karanjkar P; Lee, J; Morris J. Crystal Growth and Design 2012, 12, 3817-3824.

608

(41) Taylor C.J.; Dieker, L.E.; Miller K.T.; Koh, C.A; Sloan E.D. J. Colloid Interface Sci.

609

2007, 306, 255-261

610

(42) Rensing P.J. Studies on the Rheology of Ice and Clathrate Hydrate Slurries Formed In Situ

611

from Water-in-Oil Emulsions. Ph.D. Thesis, Colorado School of Mines, Golden, CO,

612

2010.

613

ACS Paragon Plus Environment

33

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

107x34mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 34 of 34