Investigation on Physical Properties and Morphologies of

May 9, 2016 - weighted sample (about 10 mg) was injected into a measuring cell, and the amount of water ... scattering cross section I(q) in absolute ...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV LAVAL

Article

Investigation on Physical Properties and Morphologies of Microemulsions formed with SDBS, Isobutanol, Brine, and Decane, Using Several Experimental Techniques Ayako Fukumoto, Christine Dalmazzone, Didier Frot, Loïc Barré, and Christine Noik Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.6b00595 • Publication Date (Web): 09 May 2016 Downloaded from http://pubs.acs.org on May 10, 2016

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 32

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

Investigation on Physical Properties and Morphologies of Microemulsions formed with SDBS, Isobutanol, Brine, and Decane, Using Several Experimental Techniques Ayako Fukumoto, Christine Dalmazzone,* Didier Frot, Loïc Barré, and Christine Noïk. IFP Energies nouvelles, 1-4 avenue de Bois Préau, 92852 Rueil-Malmaison, France

Abstract

In the context of chemical enhanced oil recovery (EOR), there is no well-established way to characterize and understand the physical properties and structures of microemulsions composed of crude oil and industrial surfactants due to the complexity. Making a comparison to a wellstudied model system is a simple and effective way to investigate the complex microemulsions. The purpose of this study is to provide and complete experimental characteristic data of modelmicroemulsions as a basis of analysis of the complex system. Types of microemulsions studied in the present work were oil in water (O/W), bicontinuous, and water in oil (W/O). The system was composed of water, sodium dodecyl benzene sulphonate (SDBS), isobutanol, sodium

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 32

chloride (NaCl), and decane. We performed several experiments at room temperature such as spinning drop method, differential scanning calorimetry (DSC), Karl-Fischer titration, Hyamine titration, dynamic light scattering (DLS), small-angle X-ray scattering (SAXS), and electron cryomicroscopy (cryo-SEM). The measurements provided a deeper insight into a correlation between physical properties and morphologies of water/oil in the microemulsions.

Introduction

Water and oil can be dispersed with each other by adding surfactants in a system. Microemulsions are thermodynamically stable, composed of water, oil, surfactants, and sometimes co-surfactant such as alcohol.1 One can observe the structural transition of microemulsions from water in oil (W/O) to oil in water (O/W) through bicontinuous by changing physical parameters of a system like temperature, salinity, surfactant concentration, and water to oil ratio (WOR). Bicontinuous phase can be pictured as a structure that has aqueous and oil domains separated with surfactant layer with periodicity. It is similar to that of lamellar phase, but each phase is connected to itself and the domains are chaotically intertwined.2 Commonly, microemulsion’s system is classified into Winsor I to III types.3 A Winsor I consists of O/W microemulsions as a lower phase and excess oil as an upper phase. A Winsor II consists of excess water phase as a lower phase and W/O microemulsions as an upper phase. A Winsor III has three phases; a bicontinuous microemulsions in between excess water (upper) and oil (lower) phases. Types of microemulsions are determined by the behavior of surfactants that inhabit at water/oil interface.4,5 When the surfactant has an equal affinity for oil and water, the interfacial tensions (IFT) between oil and water can be ultralow, and this leads to bicontinuous structure.6–8

ACS Paragon Plus Environment

2

Page 3 of 32

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

Microemulsions have been studied for applications such as chemical enhanced oil recovery (EOR) and pharmaceutical.9–12 Physical properties and characterizations of the microemulsions are usually investigated by different experimental measurements such as, conductivity, IFT, remained surfactant concentration, water content, and so on. In the context of chemical EOR, there is no well-established way to characterize and understand the physical properties and the structures of microemulsions composed of crude oil and industrial surfactants due to their complexity. Making a comparison to a well-studied model system is a simple and effective way to investigate the complex microemulsions. Numerous works have been done with different model systems in the past. Saito and Shinoda4 were one of the first to report the phase transition of the microemulsions (W/O to O/W) by studying a phase stability of W/O microemulsion as a function of temperature. Lindman et al.13 demonstrated that both oil and water are continuous in a middle phase with a nuclear magnetic resonance (NMR) experiment. Aveyard et al.14 investigated on phase diagrams and physical properties of the microemulsions. They measured IFT and concentration of the surfactant in aqueous and oil phases as a function of temperature and salinity. Viscosity, conductivity, and hydrodynamic radius were studied as well on W/O microemulsions. The structure of bicontinuous phase was studied with a small-angle X-ray scattering (SAXS)15, and a model16 was proposed to understand the characteristic length of the structure. Images of the structures were presented by Jahn and Strey17 using a freeze fracture electron microscopy. Garti et al.18 studied distributions of water in W/O microemulsions with a differential scanning calorimetry (DSC). Dynamical behavior of the bicontinuous phase was studied with a quasielastic light scattering.19 Among the many model systems studied, few studies share the same component. Therefore it is needed to provide characteristics of experimental results of a certain model-microemulsion for

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 32

further investigations on the complex microemulsions. With this aim, we have developed a complete methodology based on the application of several probes. This methodology will be relevant to more complex systems such as the one including real oils and will become of prime importance to the reemerging field of chemical EOR, especially surfactant flooding. The purpose of this study is to show experimental characteristics of modelmicroemulsions as a basis of analysis of the complex system. In this study, we performed several experiments at room temperature such as spinning drop method, DSC, Karl-Fischer titration, Hyamine titration, dynamic light scattering (DLS), SAXS, and electron cryomicroscopy (cryo-SEM). The system studied in this work is composed of water, sodium dodecyl benzene sulphonate (SDBS), isobutanol, sodium chloride (NaCl), and decane. The anionic surfactant SDBS was chosen because it is a well-defined molecule used in chemical, biochemical, and industrial works. This paper is organized as follows. First, experimental techniques are described in details. Results from the measurements are discussed in two parts. One is from a physicochemical and the other is from a morphological point of view. The characteristics of the bicontinuous microemulsion at a salinity of 52 g/L were focused in conclusions.

Materials and Methods

Materials. SDBS (technical grade) was purchased from Sigma Aldrich. Isobutanol (purity > 99%) and decane (99% purity) were purchased from Alfa Aesar. NaCl (99.9% purity)

ACS Paragon Plus Environment

4

Page 5 of 32

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

was purchased from VWR chemicals. All compounds were used without further purification. Aqueous solutions were prepared with Milli-Q® water made with a resistivity of 18.2 MΩ/cm. Preparation of samples. Microemulsion was prepared at room temperature in a glass tube by following procedures. First, solutions of SDBS (140 g/L) and NaCl (200 g/L) were prepared separately in flasks. The mass of SDBS and NaCl was measured with an electronic mass balance (AE200, Mettler toledo) with a precision of ± 0.1 mg. Second, 1 ml of the SDBS solution was inserted in a glass tube with a pipette. After, water and NaCl solutions were inserted in the glass tube with pipettes. Here, the volume of water and the solution was calculated from the set salinity (from 30 to 80 g/L) and set volume of aqueous phase (6.79 ml). 0.42 ml of isobutanol was added in the aqueous phase and the solution was carefully mixed. Finally, 6.79 ml of decane was inserted in the glass tube. Samples were mixed gently by turning the tubes upside down for approximately 20 times, and left to be equilibrated for 1 month at room temperature. To summarize the composition of the samples, the concentration of SDBS in aqueous phase was 20.6 g/L, WOR was set to 1, and isobutanol was 3 % (ratio in total volume). Spinning drop method. IFT of microemulsion and aqueous/oil phase was measured at 20 oC using the spinning drop video tensiometer (SVT 20N, Dataphysics). Equilibrated solutions were sampled for this method. First, density of the sample was measured with a densimeter (DMA 4500 M, Anton Paar) at 20 oC. After, about 1 ml of the aqueous phase was inserted in a glass capillary, and a drop of the phase, contacting with the aqueous phase, was injected. The capillary was rotated at high speed (5000 rpm). DSC. Crystallization and dissociation temperatures/enthalpies of microemulsions were measured using a DSC apparatus (DSC 1, Mettler toledo). The calibration of the setup was

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

Page 6 of 32

carried out with high purity indium. Water was used to determine the uncertainty of the phase change temperature and the enthalpy, which were found to be 0.40 oC and 20 %, respectively. Between 10 to 20 mg of microemulsions were inserted in a cell. Temperature of the furnace was decreased from 20 to -80 oC at a cooling rate of 5 oC·min−1, and kept constant for 5 min. After, temperature was increased to 20 oC at a heating rate of 5 oC·min−1. Temperature of crystallization and enthalpy of fusion were measured to investigate the water morphology and salinity of the microemulsions. This scanning was also conducted with SDBS + NaCl + isobutanol + water solutions as a reference. Karl-Fischer titration. Mass fraction of water in oil phase or microemulsion phase was measured by the Karl-Fischer titration. The titration was conducted with 787-KF Titrino from Metrohm with a reagent of HYDRANAL®-Composite 5 purchased from Fluka. A weighted sample (about 10 mg) was injected into a measuring cell, and the amount of water in the sample was calculated according to the volume of reagent consumed. All the samples were analyzed at least three times. Hyamine titration. Concentration of SDBS in aqueous phase was estimated by the Hyamine titration (Hyamine® 1622-solution, Merck Millipore) using a titrator (862-compact titrosampler, Metrohm). 1 ml of the sample was inserted in a measuring cell. The titrant (Hyamine) was added in the sample with an increment of 0.03 ml. An electrode was used for the potentiometric titration to determine the amount of the anionic surfactants in the sample. DLS. Dynamics of microemulsions were studied using a DLS setup (Vasco-flex, Cordouan technologies) equipped with a diode laser (λ = 656 nm) and a single photon detector. Measurements were conducted at a scattering angle (θ) of 90o. A detailed description of the setup

ACS Paragon Plus Environment

6

Page 7 of 32

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

can be found elsewhere.20 It is noteworthy that this measurement did not require any sampling. The equipment was specially modified in order to mount a tube on the setup so that the laser beam can directly go through the phase of interest in order to observe its dynamics. In this measurements, intensity fluctuations of the scattered light were collected as a function of time. From the measured fluctuation, the autocorrelation function of the intensity G2(q,t) can be derived using the Siegert relation, and can be expressed as

G2 (q, t ) = α + βg1 (q, t ) , 2

(1)

where α and β are a baseline and the coherence factor, respectively. q (= 4πnsin(θ/2)/λ0) and t represent scattering vector and time, respectively. n is the refractive index of the medium and λ0 is the radiation wavelength in vacuum. The normalized field autocorrelation function g1(q,t) is related to the dynamic behavior of the sample. For monodispersed particles, it is appropriate to describe g1(q,t) as a single exponential decay: g 1 (q , t ) = exp (− Γt ) ,

(2)

where Г is the decay rate. The diffusion coefficient of the particles can be given as D = Г/q2. In the case of hard spheres well dispersed in a solution with no interactions and only diffuse as Brownian motion, the hydrodynamic radius RH of the sphere is given by the Stokes-Einstein equation as

D = kT / 6πν 0 RH ,

(3)

where k, T, and ν0 represent the Boltzmann constant, temperature, and viscosity of solvent, respectively.

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 32

SAXS. An in-house experimental setup was used to investigate the structural characteristics of microemulsions. Sample was injected into a capillary with a diameter of 1.5 mm. Based on measurements of an external standard (Ag behenate), the position of each pixel, relative to the position of the direct beam, was converted into the magnitude of the scattering vector q (= 4πnsin(θ/2)/λ0). The range of the scattering angle (θ) enables a range of q from 7 × 10−3 to 3 × 10−1 Å−1 to be covered. After normalization with respect to thickness, transmission, and measuring time, the solvent signal was subtracted from the sample signal, and the raw intensities were converted to the scattering cross section I(q) in absolute scale (cm-1). Further details of the description of the setup can be found elsewhere.20–22 Cryo-SEM. This technique was used to visualize the structure of bicontinuous microemulsions. In this study, a small piece of the middle phase solutions at a salinity of 52 g/L was shock frozen in nitrogen slush at −210°C, fractured, and transferred in a vacuum to the scanning electron microscope (Supra 40, Zeiss) fitted with a cold stage unit. Rheometer. Viscosity of O/W and W/O microemulsion was measured with AR2000 rheometer (TA Instruments) using a double cylinder geometry. After carefully loading the sample in the geometry, 1 min of equilibration time was imposed. Shear rate measurements were performed between 0.1 and 700 s-1. This measurement was conducted in order to estimate the hydrodynamic radius of the microemulsions from the DLS measurements as explained later in the results and discussion.

Results and Discussion

ACS Paragon Plus Environment

8

Page 9 of 32

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

Physicochemical analysis. Phase behavior. Phase transitions of the samples were investigated with a salinity scan in the range from 24 to 96 g/L with a minimum interval of 2 g/L. The samples were kept at room temperature for 1 month in order to have equilibrated phases. It was found that the phases reached equilibrium 4 days after the preparation, because heights of the phases remained constant since. Middle phases were observed at salinities from 36 to 74 g/L. The phase transitions in terms of the microemulsions can be described as follows: Winsor I (O/W) ranged from 24 to 34 g/L, Winsor III (bicontinuous) ranged from 36 to 74 g/L, and Winsor II (W/O) ranged from 76 to 96 g/L. The photographic and schematic images of some of the formulated samples are shown in Figure 1. The aqueous phase was white-translucent at a salinity of 30 g/L because of dispersed oil droplets. At salinities of 40, 52, and 64 g/L, three phases were observed. The middle phase at 40 g/L was slightly yellow-colored. The color is due to SDBS molecules because the similar color was also seen with a concentrated SDBS solution. As for samples at 52, 64, and 80 g/L, all phases were transparent. Note that the dark region at the top of the lower phase is seen in the photographic image because of an optical effect at the interface. IFT. Figure 2 shows the IFTs between equilibrated microemulsion and oil/water phases measured with the spinning drop method. At salinities of 30, 52, and 80 g/L, the measurements were conducted three times, and mean values are plotted. The error of the measurements were found to be less than 7 %, which is smaller than the symbol (filled square) in the figure. As can be seen from the figure, the order of magnitude of IFT was -2 when the system is Winsor III. For comparison, the IFT measured by Moiré23 with the same technique is plotted in the figure. The IFT reported from the literature is substantially smaller than ours at a salinity of 52 g/L. When a system is Winsor III, the interfaces can be easily disordered because of the ultralow IFT,

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 32

especially on the occasion of samplings. It can cause a transfer of SDBS molecules between phases, which can result in the deviation of IFTs between the measurements. At salinities of 30 and 80 g/L, the results agree well with the literature. This is reasonable because when the system is Winsor I/ II, sampling is easier. One can say that optimum salinity of the system is in the vicinity of 50 g/L, meaning that the amphiphilicity is high in this region. Water content of oil phase and surfactant concentration in aqueous phase. Compositions of the aqueous and oil phases were investigated by the Karl-Fischer and the Hyamine titration. The results are summarized in Figure 3. It shows that about 0.4 to 1 wt% of water was in oil phase when systems are Winsor I and III. The water content increased rapidly at a salinity of 80 g/L, which is in the range of Winsor II (W/O microemulsions). As for the SDBS concentration in aqueous phase, while about 95 % of the surfactant remained in the aqueous phase when the system is Winsor I, less than 1 % remained when it is Winsor II/ III. The results refer that these compositions of the oil and aqueous phases are good indications for characterizing the state of a system; low water content of oil phase and high concentration of surfactants in aqueous phase indicate a system is Winsor I, when the properties are contrary, a system is Winsor II. Both parameters are low when Winsor III, because surfactants are concentrated in the middle phase, thus few oil can transfer to the aqueous phase. Salinity and water content of microemulsion. From the heat profiles of the melting process obtained from DSC measurements, one can estimate the concentration of NaCl in the microemulsions by comparing the onset temperature of the progressive melting of ice with that of the aqueous solutions. Figure 4 shows heat flow profiles of the aqueous solutions and decane. During the melting process, one peak was detected when no NaCl was present. Two negative peaks were observed with the solutions with NaCl; first one detected around -20 oC corresponded

ACS Paragon Plus Environment

10

Page 11 of 32

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

to eutectic melting of NaCl and ice, and the second broad peak was detected soon after the first one, which was due to the progressive melting of ice. The melting temperature of decane was about -30 oC, which is in a good agreement with the literature.24 From Figure 4, a relation between temperature of the progressive melting of ice and salinity was found as T= 4.99×10-4x2 – 1.11×10-1x with the average absolute deviations (AAD) of 0.13 oC, where T is the melting temperature in oC and x is the concentration of NaCl in g/L. It is noted that this relation is only valid within the studied salinities. Here, the temperature of progressive melting of ice was extrapolated from an intersection point of a baseline and a line which goes through the peak point of the progressive melting. For an inclination of the line, the same value of the tangent at an inflection point of the eutectic melting was used. The heat flow profiles of the microemulsions at different salinities are shown in Figure 5. At a salinity of 80 g/L, a negative peak was observed around -30 oC, which corresponds to a dissociation of decane. It is suspected that the heat absorbed by melting of dispersed water droplets was too small to be detected (water content was 3 wt%, see Figure 3). As for salinities at 30, 40, 52, and 64 g/L, three peaks were detected. First peak observed around -30 oC was induced by decane melting. Second and third peaks were due to eutectic and progressive melting of ice, respectively. The temperature of the progressive melting of ice in microemulsion was determined with the same procedure as the aqueous solutions, and the salinity was estimated using the equation shown above. The comparison between the initial concentration of NaCl and the value obtained from the DSC measurements are shown in Figure 6 (a). Error bars of y axis were estimated from the uncertainty of the DSC setup measured with water. Note that the concentration of NaCl was not estimated with the microemulsion at a salinity of 80g/L since a significant peak of the ice melting was not detected.

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 32

The estimation from the DSC measurements slightly underestimates the salinity. The deviations were 7, 10, 14, and 9 g/L at 30, 40, 52, and 64 g/L, respectively. Water content of microemulsion can also be estimated by comparing the total enthalpies (surface area of the peaks) of the eutectic and progressive melting of ice in microemulsions with that in aqueous solutions. A relation between the melting enthalpy and salinity (within the studied range of salinities), when water content was 100 %, was found as ∆H= 5.00 × 10-1 x – 3.25× 10-2 with the ADD of 2.4 %. Here, ∆H represents the melting enthalpy in J/g. Likewise, oil content of microemulsion can be estimated by comparing the enthalpy of decane melting. Melting enthalpy of decane was found as 240 J/g, which is in the error range of the literature25. This value was used as reference of oil content of 100 %. The comparison of water content measured with the Karl-Fischer method and the DSC measurements are shown in Figure 6 (b). Squares were calculated using the melting enthalpies of the aqueous solutions as reference. While circles were calculated with the decane melting, assuming wwater= 1 – wdecane. Here w represents mass fraction of each component. The estimation of DSC using the aqueous melting roughly equals the value measured with the Karl-Fischer titration. The estimation from the decane melting, however, deviated about 20 wt% from the Karl-Fischer titration. The possible reason can be an assumption we made in the estimation, which ignored the mass contribution of surfactant, NaCl, and alcohol. Due to this assumption, water content was overestimated. Despite the time cost, it is recommended to use melting profiles of aqueous solutions when estimating water contents from DSC measurements. From Figure 6 (b), one can say that a salinity of 52 g/L is close to the optimum salinity because the volume of water in microemulsion was estimated to be 48 wt% with the DSC result.

ACS Paragon Plus Environment

12

Page 13 of 32

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

Morphological analysis. Images of bicontinuous microemulsions. Cryo-SEM allowed us to have the images of the structure of the bicontinuous microemulsion. Obtained images of the sample at a salinity of 52 g/L are shown in Figure 7. From Figure 7 (a), it was observed that the bicontinuous phase is a homogeneous structure at this scale. Figure 7 (b) shows an image at a higher resolution. From the color contrast, which is due to the difference of chemical composition and topology, it is confirmed that water and oil domain are intertwined at this scale. The reader is directed to the literature17,26, which shows similar images of the bicontinuous phase taken with an electron microscopy. Crystallization of microemulsions. Dalmazzone et al.27 demonstrated the characterization of opaque-emulsions using a DSC. Temperature of crystallization is related to morphology of water in microemulsions since a probability of nucleation of ice is dependent of the volume of water. It has been shown in the literature that the smaller the volume is, the lower the freezing temperature is. As shown in Figure 5, during the cooling process, sharp positive peaks were observed around -20 oC at salinities of 30 and 40 g/L, which are in the same range as measured in the aqueous solutions (Figure 4). It indicates the existence of water in a volume scale of mm3. This is understandable for the sample at a salinity of 30 g/L, since the microemulsion was O/W, meaning oil droplets dispersed in a continuous water phase. As for 40 g/L, one could conclude that water channels in the microemulsion were so large that the peak did not shifted despite the formation of the bicontinuous structure. However, it is not the case for this sample. The heat profile similar to 30 g/L was obtained because water and oil domains in the bicontinuous microemulsion were separated during the cooling process. The phase separation was visually confirmed by a simple experiment; about 5 ml of the bicontinuous microemulsion was frozen in

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 32

a household freezer, and defrosted in a room temperature. The phase separation was observed at a salinity of 40 g/L, but not at salinities of 52 and 64 g/L. We believe that the microemulsion at 40 g/L destabilized because it is near the verge of Winsor III region and the structure can change more easily than that at 52 and 64 g/L. For the samples at 30 and 40 g/L, second peaks were observed around -35 oC, which correspond to decane crystallizations. Following the second peaks, third peaks were detected between -45 and -40 oC. These small peaks were attributed to NaCl crystallizations. It must be noted that the bicontinuous microemulsions at a salinity of 52 g/L exhibited only one peak of crystallization. This supports the fact that for the bicontinuous phase at around the optimum salinity, water and oil phases are neighboring each other laying the surfactant molecules in between the layers. Thus the crystallizations of water and oil were induced together. At a salinity of 64 g/L, the peak of decane crystallization was following a small sharp peak observed at -32 oC. This sharp peak suggests the presence of water in a volume scale of µm3 in the bicontinuous phase. It can be interpreted as a bicontinuous phase with a poor connection of the water channel. As for W/O microemulsions at a salinity of 80 g/L, one peak was obtained at -35 oC, which is mainly due to the decane crystallization. It is suspected that the crystallization of water droplets were induced by the surrounding decane, hence the one peak corresponds to both decane and water freezing. Dynamical behavior. The dynamics of O/W and W/O microemulsions can be evaluated using the hydrodynamic radius from the Stokes-Einstein equation as mentioned before since they are composed of dispersed particles. However for bicontinuous microemulsions, it is not suitable for the evaluation. To compare the dynamics of all phases, the diffusion coefficient D was used for the sake of simplicity. The cumulant method was applied to have a representative decay rate Γ of the auto correlation data. The diffusion coefficient D, which describes the averaged

ACS Paragon Plus Environment

14

Page 15 of 32

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

fluctuation of the scattered intensity, was obtained ( D = Γ / q ). It is noted that D obtained from 2

a middle phase does not represent the diffusivity of components in the phase, but it only corresponds to the characteristic time of the dynamics, which is defined as τc= 1/ Γ . The higher D is, the smaller the characteristics time is. Figure 8 shows the measured autocorrelation functions of microemulsions at 30, 52, and 80 g/L. The solid lines represent the exponential decays using Γ . It is noteworthy that all the samples were quite transparent and there was no multi-scattering of the light. Also, there was no dust in the samples, which was confirmed by measuring the total scattered intensity versus long time. In Figure 9, D of aqueous, middle, and oil phases are plotted at different salinities. The error bars in the figure represent the reproducibility of the measurements, which was found to be around 10 %. D of the middle phase increased, but it decreased beyond this concentration. The order of magnitude of D of middle phases measured in this study are consistent with the literature,19,28 which reports the diffusion coefficient of microemulsions formed with anionic surfactants, brine, and simple alkane oil. In bicontinuous microemulsions, the fluctuation is dependent on the size of water and oil domains because it is mainly due to a topology change in the phase; breaking and mending of water/oil channels. As can be seen in the figure, D of middle phase maximized near the optimum salinity because the domain size is smaller and hence more frequent topology change was observed. As for the aqueous phase, the fluctuation decreased along with the salinity increment when system was Winsor I. This can be due to increments of size of dispersed oil droplets or interactions between the particles. When system was Winsor II or III, no fluctuation of the intensity was observed in the aqueous phase (plotted as D= 0 in the figure). Although the D of oil phase in Winsor I and III was one order of magnitude smaller than that in Winsor II, the fluctuation of the phase was confirmed at all salinities studied in this work.

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 32

This is consistent with our observation with the Karl-Fischer titration (Figure 3), which confirmed the presence of water in the oil phase in all Winsor systems. Characteristic lengths of structures. Structures of the microemulsions were investigated, and the characteristic lengths were extracted from the SAXS measurements. Figure 10 shows the SAXS spectra of microemulsions. Note that the signals of the solvents (water and decane) were subtracted with respect to volume fractions. The volume fractions of water (φw) and decane (φo) were estimated by the water fraction in mass (ww) obtained from the Karl-Fischer titration and the density (ρME) as φw= ww×ρME/ρw, where ρw represents the density of water. φo is calculated as 1- φw. From the figure, the dispersion of the measured intensity is clearly visible at large q values for 80 g/L. For other spectra, especially for bicontinuous microemulsions, uncertainties are smaller than the symbol sizes. The scattering curve of microemulsion at salinities of 40, 52, and 64 g/L exhibited a single broad peak, which is a typical pattern from bicontinuous microemulsions.16,29–31 As for 30 and 80 g/L, no characteristic peak was observed. In the small q regime, the scattered intensity is plateau for the O/W emulsion at 30 g/L. On the other hand, the intensity decreased significantly for the W/O microemulsion at 80 g/L. One can conclude that the microemulsion at 80 g/L had significant interactions of dispersed water droplets such as aggregations. In the large q regime, the scattering intensities of all the tested samples decreased with the same scale as q-3.2, meaning that the structure on a small length scale exhibits no difference depending on the salinity. Theoretically speaking, one should observe the decay of the curve scaling as q-4 (Porod’s law) in the large q regime, which is a characteristic of a sharp interface (i.e., a step function of scattering length density). The difference between the law and our measurements is attributed to a SDBS molecule at the interface. This makes the scattering length density more complex than the simple

ACS Paragon Plus Environment

16

Page 17 of 32

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

approximation using the solvents such as φwφo(ρw-ρo)2. It is suggested to use a profile of the scattering length density at the interface that is closer to the actual one. The scattering length densities at the interface can be calculated by combining the bound coherent scattering length and the molar volume32 of the atom/group. The values related to the interface are: 7.12 × 1010 and 9.4 × 1010 cm-2 for decane and water, respectively. As for SDBS, a sulfonate ion is 23 × 1010, a phenylene is 9.5 × 1010, a straight alkyl chain is 8.4 × 1010, and a methyl is 4.5 × 1010 cm-2. The scattered intensity I(q) of the O/W microemulsion at 30 g/L was fitted with the Guinier approximation, which is given by using radius of gyration Rg as

(

)

I (q ) = I (0) exp − q 2 Rg / 3 . 2

(4)

Assuming that droplets are spherical particles homogeneously dispersed in the solution, the statistical radius of the particles R0 can be estimated by Rg2= 3/5 R02.33 The scattered intensity I(q) of the bicontinuous microemulsions was fitted with the model proposed by Teubner and Strey16 (T-S model). In the T-S model, a real space correlation function ϒ(r) is described as ϒ(r) = sin(Kr)/Kr·exp(-r/ξ),

(5)

where K is given as K= 2π/d. The scattered intensity I(q) is described with d and ξ as

I (q ) =

8πc 2 η 2 / ξ a 2 + c1 q 2 + c 2 q 4

,

(6)

where is the mean square fluctuation of the scattering length density (ρ) given by φwφo(ρwρo)2. Physical parameters d and ξ can be written using a2, c1, and c2 in Eq. (6) as

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

 1  a 1 / 2 c  ξ =   2  + 1  4c2   2  c2  

−1 / 2

 1  a 1 / 2 c  , d = 2π   2  − 1  4c2   2  c2  

Page 18 of 32

−1 / 2

.

(7)

By substituting Eq. (7) into Eq. (6), I(q) is described with three fitting parameters d, ξ, and . Here, the parameters d and ξ represent periodicity of the oil and water domain and the decay length of the periodic order, respectively. Solid lines in Figure 10 were drawn by T-S model and the Guinier approximation using the parameters listed in Table 1 for the microemulsions at 30, 40, 52 and 64 g/L. As for the microemulsion at 80 g/L, no line is drawn because the Guinier fitting does not agree well with the measurement since there is no plateau in the spectra. One can observe the good agreement in the SAXS measurements and the fitted lines in the q range before 0.07 Å-1. For the bicontinuous microemulsions, it is found that the domain size (d) takes minimum and the correlation length (ξ) takes maximum at 52 g/L. Uncertainties of d and ξ were estimated to be within 0.2 nm. A radius of dispersed O/W microemulsions (R0) was found to be 6 nm. Although no structural length was determined for W/O microemulsion at 80 g/L, one can presume from the curve of the spectra that the size of droplet is bigger but in the same order of magnitude as compared to that of O/W microemulsion at 30 g/L. Chen et al.34 reported the physical meaning of ξ by using the ratio of the two parameters ξ/d; the ratio is related to a polydispersity of the domain size, the smaller the ratio is, the bigger the polydispersity is. We found 0.31, 0.39, and 0.32 for the bicontinuous microemulsions at salinities of 40, 52, and 64 g/L, respectively. The ratio maximized at the salinity of 52 g/L, indicating the most ordered structure among the tested samples. ξ/d is reported in literature for different systems; 0.41-0.65 for D2O + decane + sodium bis (2-ethylhexyl) sulfosuccinate (AOT)

ACS Paragon Plus Environment

18

Page 19 of 32

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

system studied by Teubner and Strey16, 0.35-0.41 for D2O + toluene + butanol + NaCl + SDS (sodium dodecyl sulfate) studied also by Teubner and Strey16, 0.29-0.40 for water + NaCl + decane + AOT system reported by Chen et al.29 Note that ξ/d obtained in this study is in the same range as the literature, however, when comparing the structural order with another system using the parameter, one should be careful with the magnitude of hydrophilicity/hydrophobility of the system, which depends on the composition of the system. For example, ξ/d at an optimum salinity should be a good comparison for the polydispersity when comparing different systems. The observation of d and ξ/d supports the results of the DLS measurements, which showed the more frequent topology change around the optimum salinity. The correlation between D, d, and ξ/d are summarized in Figure 11. The hydrodynamic radius of O/W and W/O microemulsions can be estimated from the DLS measurements with the Stokes-Einstein equation. Here, the viscosity of the solvent ν0 was estimated from the Einstein’s equation as ν= ν0(1 + 2.5φ), where ν represents the viscosity of microemulsions measured with the rheometer and φ is the volume fraction of a dispersed phase. Only one characteristic decay was found in both microemulsions. The calculated RH is listed in Table 1. For the microemulsion at 30 g/L, RH was found to be bigger than R0. The difference can be caused from polydispersity of the dispersed phase or a contribution of counter ions to the hydrodynamic radius. RH of the microemulsion at 80 g/L was observed to be bigger than that of the microemulsion at 30 g/L. This is consistent with the assumption we made from the curve of SAXS spectra.

Conclusions

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 32

In this work, O/W, bicontinuous, and W/O microemulsions composed of water, NaCl, SDBS, isobutanol, and decane were studied at room temperature using several experimental techniques such as spinning drop method, DSC, Karl-Fischer titration, Hyamine titration, DLS, SAXS, and cryo-SEM. The measurements provided us a deeper insight into a correlation between physical properties and morphologies of water/oil in the microemulsions. The optimum salinity was found to be around 50 g/L from the IFT measurement. The bicontinuous microemulsion at a salinity of 52 g/L showed following characteristics. From the DSC measurement, it was found that the concentration of NaCl slightly decreased compared to the initial salinity of aqueous solution. Also, water content of the bicontinuous microemulsion was estimated to be 48 wt% from the DSC measurement, which agrees well with the value measured with the Karl-Fischer titration. The value supports that a salinity of 52 g/L is close to the optimum salinity, at which the volume of water and oil solubilized in microemulsion is equal. The DSC measurement also revealed that water and oil crystalized at the same time in the bicontinuous microemulsion. This leads to the morphological observation that water and oil domains were neighboring each other and well intertwined, which can also be seen in the images from cryo-SEM. Physical properties of excess water and oil phases were studied with the KarlFischer titration and the Hyamine titration. It was confirmed that very few surfactants (0.4 %) remained in the aqueous phase and a small amount of water (0.5 wt%) existed in oil phase. Correspondingly, the time fluctuation of the excess oil phase measured with the DLS corroborated the presence of water droplets in the phase. Dynamics of the middle phase were investigated with the DLS in terms of the diffusion coefficient, which maximized near the optimum salinity. SAXS measurements revealed that the domain size and its polydispersity took minimum at a salinity of 52 g/L. By combining the observation of DLS and SAXS, one can

ACS Paragon Plus Environment

20

Page 21 of 32

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

conclude that the fluctuation of the middle phase is mainly due to topology changes (breaking and mending of water/oil channels). We conclude that characteristics of the model-microemulsions were well studied and revealed in this study. The methodology that has been developed is currently being applied to complex oil systems which are more representative of real crude oils, that means solutions of asphaltenes. This will be the subject of a future paper. In a second step, we intend to investigate systems representative of microemulsions that should be used in chemical EOR operations (surfactant flooding), including real crude oils and complex mixtures of surfactants. The results shown here will be a great help when interpreting that of more complex systems, composed of industrial surfactant or crude oil, which are of our next interests.

AUTHOR INFORMATION Corresponding Author *

E-mail: [email protected]

Notes The authors declare no competing financial interest. Acknowledgments The authors thank M. Moiré (IFPEN) for cryo-SEM measurements and A. Knorst-Fouran (IFPEN) for helpful discussions. References

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 32

(1) Shinoda, K.; Friberg, S. Microemulsions: Colloidal aspects. Advances in Colloid and Interface Science 1975, 4, 281–300. (2) Kaler, E. W.; Bennett, K. E.; Davis, H. T.; Scriven, L. E. Toward understanding microemulsion microstructure: A small‐angle x‐ray scattering study. J. Chem. Phys. 1983, 79, 5673–5684. (3) Winsor, P. A. Solvent properties of amphiphilic compounds; Butterworths: London, 1954. (4) Saito, H.; Shinoda, K. The stability of W/O type emulsions as a function of temperature and of the hydrophilic chain length of the emulsifier. Journal of Colloid and Interface Science 1970, 32, 647–651. (5) Carnali, J. O.; Ceglie, A.; Lindman, B.; Shinoda, K. Structure of Microemulsions in the Brine/Aerosol OT/Isooctane System at the Hydrophile-Lipophile Balance Structure of Microemulsions in the Brine/Aerosol OT/Isooctane System at the Hydrophile-Lipophile Balance Temperature Studied by the Self-Diffusion Technique. Langmuir 1986, 2, 417–423. (6) Shinoda, K.; Sagitani, H. Emulsifier selection in water/oil type emulsions by the hydrophile—lipophile balance—temperature system. Journal of Colloid and Interface Science 1978, 64, 68–71. (7) Boyd, J.; Parkinson, C.; Sherman, P. Factors affecting emulsion stability, and the HLB concept. Journal of Colloid and Interface Science 1972, 41, 359–370. (8) Strey, R. Microemulsion microstructure and interfacial curvature. Colloid & Polymer Science 1994, 272, 1005–1019. (9) Herzig, E. M.; White, K. A.; Schofield, A. B.; Poon, W C K; Clegg, P. S. Bicontinuous emulsions stabilized solely by colloidal particles. Nature materials 2007, 6, 966–971. (10) Liu, H.; Wang, Y.; Lang, Y.; Yao, H.; Dong, Y.; Li, S. Bicontinuous cyclosporin a loaded water-AOT/Tween 85-isopropylmyristate microemulsion: structural characterization and dermal pharmacokinetics in vivo. Journal of pharmaceutical sciences 2009, 98, 1167–1176. (11) Mittal, K. L.; Kumar, P. Handbook of microemulsion science and technology; Marcel Dekker: New York, 1999. (12) Sripriya, R.; Muthu Raja, K.; Santhosh, G.; Chandrasekaran, M.; Noel, M. The effect of structure of oil phase, surfactant and co-surfactant on the physicochemical and electrochemical properties of bicontinuous microemulsion. Journal of Colloid and Interface Science 2007, 314, 712–717. (13) Lindman, B.; Kamenka, N.; Kathopoulis, T. M.; Brun, B.; Nilsson, P. G. Translational diffusion and solution structure of microemulsions. J. Phys. Chem. 1980, 84, 2485–2490. (14) Aveyard, R.; Binks, B. P.; Clark, S.; Mead, J. Interfacial tension minima in oil-watersurfactant systems. Behaviour of alkane-aqueous NaCl systems containing aerosol OT. J. Chem. Soc., Faraday Trans. 1 1986, 82, 125–142. (15) Auvray, L.; Cotton, J.-P.; Ober, R.; Taupin, C. Concentrated Winsor microemulsions: A small angle X-ray scattering study. J. Phys. France 1984, 45, 913–928. (16) Teubner, M.; Strey, R. Origin of the scattering peak in microemulsions. J. Chem. Phys. 1987, 87, 3195. (17) Jahn, W.; Strey, R. Microstructure of microemulsions by freeze fracture electron microscopy. J. Phys. Chem. 1988, 92, 2294–2301. (18) Garti, N.; Aserin, A.; Tiunova, I.; Fanun, M. A DSC study of water behavior in water-in-oil microemulsions stabilized by sucrose esters and butanol. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2000, 170, 1–18.

ACS Paragon Plus Environment

22

Page 23 of 32

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

(19) Peter, U.; Roux, D.; Sood, A. K. Observation of a topological relaxation mode in microemulsions. Physical review letters 2001, 86, 3340–3343. (20) Eyssautier, J.; Frot, D.; Barré, L. Structure and dynamic properties of colloidal asphaltene aggregates. Langmuir 2012, 28, 11997–12004. (21) Barré, L.; Simon, S.; Palermo, T. Solution properties of asphaltenes. Langmuir 2008, 24, 3709–3717. (22) Yarranton, H. W.; Ortiz, D. P.; Barrera, D. M.; Baydak, E. N.; Barré, L.; Frot, D.; Eyssautier, J.; Zeng, H.; Xu, Z.; Dechaine, G.; Becerra, M.; Shaw, J. M.; McKenna, A. M.; Mapolelo, M. M.; Bohne, C.; Yang, Z.; Oake, J. On the Size Distribution of Self-Associated Asphaltenes. Energy Fuels 2013, 27, 5083–5106. (23) Moiré, M. Etude des propriétés interfaciales eau/huile/tensioactifs par microfluidique (Study of the interfacial properties of water/oil/surfactants by microfluidics). Ph.D. Thesis, University Pierre et Marie Curie, UPMC, Paris, France, 2015. (24) Kukarni, A. A.; Zarah, B. Y.; Luks, K. D.; Kohn, J. P. Phase-equilibriums behavior of system carbon dioxide-n-decane at low temperatures. J. Chem. Eng. Data 1974, 19, 92–94. (25) Finke, H. L.; Gross, M. E.; Waddington, G.; Huffman, H. M. Low-temperature Thermal Data for the Nine Normal Paraffin Hydrocarbons from Octane to Hexadecane. J. Am. Chem. Soc. 1954, 76, 333–341. (26) Issman, L.; Talmon, Y. Cryo-SEM specimen preparation under controlled temperature and concentration conditions. Journal of microscopy 2012, 246, 60–69. (27) Dalmazzone, C.; Noïk, C.; Clausse, D. Application of DSC for Emulsified System Characterization. Oil & Gas Science and Technology - Rev. IFP 2009, 64, 543–555. (28) Shukla, A.; Graener, H.; Neubert, Reinhard H H. Observation of two diffusive relaxation modes in microemulsions by dynamic light scattering. Langmuir : the ACS journal of surfaces and colloids 2004, 20, 8526–8530. (29) Chen, S. H.; Chang, S. L.; Strey, R. On the interpretation of scattering peaks from bicontinuous microemulsions. In Trends in Colloid and Interface Science IV; Zulauf, M., Lindner, P., Terech, P., Eds.; Progress in Colloid & Polymer Science; Steinkopff, 1990; pp 3035. (30) Silas, J. A.; Kaler, E. W. Effect of multiple scattering on SANS spectra from bicontinuous microemulsions. Journal of Colloid and Interface Science 2003, 257, 291–298. (31) Chen, S.-H.; Chang, S.-L.; Strey, R. Structural evolution within the one-phase region of a three-component microemulsion system: Water–n-decane–sodium-bis-ethylhexylsulfosuccinate (AOT). J. Chem. Phys. 1990, 93, 1907. (32) Fedors, R. F. A method for estimating both the solubility parameters and molar volumes of liquids. Polym Eng Sci 1974, 14, 147–154. (33) van Blaaderen, A.; Vrij, A. Synthesis and characterization of colloidal dispersions of fluorescent, monodisperse silica spheres. Langmuir 1992, 8, 2921–2931. (34) Chen, S. H.; Chang, S. L.; Strey, R. Simulation of bicontinuous microemulsions: comparison of simulated real-space microstructures with scattering experiments. J Appl Crystallogr 1991, 24, 721–731.

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 32

Table 1 Structural parameters of microemulsions. φw is volume fraction of water in microemulsions. ν0 is viscosity of solvent in microemulsions. d, ξ, and η are fitting parameters of T-S model. I(0) and R0 were determined by the Guinier approximation. RH is hydrodynamic radius given by Eq. (3) NaCl, φw ν0, d, ξ, η×1010, I(0), R0, a RH,b -2 -1 g/L mPa·s nm nm cm cm nm nm 30 87.4 1.29 12.4 5.9 9.8 40 57.4 - 23.7 7.4 1.19 52 40.1 - 20.0 7.7 1.36 64 26.8 - 20.6 6.6 1.40 80 2.2 1.14 14.6 a b Calculated using SAXS results. Calculated using DLS results.

Figure 1 Photographic and schematic images of formulated microemulsions with increase of salinity. Samples were kept at room temperature for 1 month. System was composed of water, NaCl, SDBS, isobutanol, and decane. WOR is unity. WI, WII, and WIII represent Winsor I, II, and III, respectively. A and O denote aqueous and oil phases, respectively.

ACS Paragon Plus Environment

24

Page 25 of 32

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

Figure 2 IFT between equilibrated microemulsion and water/oil phases as a function of salinity. Data from literature23 is shown as a comparison.

Figure 3 Water content of oil phase measured by Karl-Fischer titration and concentration of SDBS in aqueous phase measured by Hyamine titration.

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 32

Figure 4 Heat flow profiles of aqueous solutions composed of SDBS + water + isobutanol with different salinities, and decane.

ACS Paragon Plus Environment

26

Page 27 of 32

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

Figure 5 Heat flow profiles of microemulsions with different salinities. Types of microemulsions are O/W at 30 g/L, bicontinuous at 40, 52, and 64 g/L, and W/O at 80 g/L.

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 32

Figure 6 Comparison of concentration of NaCl in microemulsion calculated from DSC measurements and that at initial state (a). Comparison of water content of microemulsion calculated from DSC measurements and measured with the Karl-Fischer titration (b). Numbers shown next to symbols represent the salinities. Square is calculated using results of aqueous solution melting. Circle is calculated using results of decane melting. The lines correspond to f(x)= x.

ACS Paragon Plus Environment

28

Page 29 of 32

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

Figure 7 Images of bicontinuous microemulsions (middle phase) at a salinity of 52 g/L obtained with cryo-SEM. (a) and (b) were taken at different resolutions (see scales shown in each image).

Figure 8 Normalized autocorrelation against time measured at different salinities. Types of microemulsions are O/W at 30 g/L, bicontinuous at 52 g/L, and W/O at 80 g/L. Symbols are from DLS measurements and solid lines correspond to the exponential decay.

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

Page 30 of 32

Figure 9 Diffusion coefficient (D) of relaxation process of aqueous, middle, and oil phases as a function of salinity.

Figure 10 SAXS spectra of microemulsions. Types of microemulsions are O/W at 30 g/L, bicontinuous at 40, 52, and 64 g/L, and W/O at 80 g/L. Solid lines are the best fits by T-S model or Guinier equation. A dashed line is the power law of q−4.

ACS Paragon Plus Environment

30

Page 31 of 32

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

Figure 11 Characteristic lengths of bicontinuous microemulsions versus diffusion coefficient. d and ξ/d represent domain size and polydispersity of bicontinuous microemulsions obtained from SAXS measurements. D is the diffusion coefficient derived from the average decay rate. Numbers shown on the upper base of the figure represent the salinities.

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

Page 32 of 32

For Table of Contents use only Investigation on Physical Properties and Morphologies of Microemulsions formed with SDBS, Isobutanol, Brine, and Decane, Using Several Experimental Techniques

Ayako Fukumoto1, Christine Dalmazzone1,*, Didier Frot1, Loïc Barré1, and Christine Noïk1

1

IFP Energies nouvelles, 1-4 avenue de Bois Préau, 92852 Rueil-Malmaison, France

*Corresponding author: Email : [email protected]

Toc figure

ACS Paragon Plus Environment

32