Вы находитесь на странице: 1из 97

Preface

This presentation is submitted as the summer dissertation project as the major prerequisite for the Master of Science degree in Chemical Engineering at the University of Nottingham, United Kingdom. Both laboratory experiments and thesis construction were conducted under the full supervision of Dr Buddhi

Hewakandamby in the Department of Chemical and Environmental Engineering, University of Nottingham between June 2010 and September 2010. I hereby declare that this thesis is a work of my own and has not been previously submitted by me at another institution for any degree. I also cede copyright of the thesis in favour of the University of Nottingham, United Kingdom. This piece of work has not been shared or used by any other persons other than me and my project supervisor, Dr Buddhi Hewakandamby. Where I have used and quoted the work of other people, they have been acknowledged and detailed explicitly throughout the entire presentation. ____________________ Chan Yung Khiong 17th September, 2010

Abstract
The drive for the study of droplet coalescence stretches across numerous liquidliquid emulsion engineering applications including food, lubricants, pharmaceutical and oil industry. Such research is also very beneficial for the environmental sectors. However, liquid-liquid emulsion is complex in nature and its full comprehension is yet to be achieved. To help break this barrier a step further, emulsion was studied in this experiment on a smaller scale before escalating to more complicated conditions. The experiment was composed of the investigation of binary

coalescence dynamics of a pair of kerosene droplets forming at low inlet flowrate through capillaries in reverse osmosis (RO) water. Throughout the whole

experiment, the evolution of the kerosene droplet coalescence process was recorded using a high speed video camera and the mean binary coalescence times were calculated for different conditions. Various concentrations of glycerol were added into the system to alter the interfacial tension and hence surface free energy between the kerosene droplets and RO water. Increasing the glycerol concentration decreased the interfacial tension and hence, surface free energy. Two methods were employed during the experiment: (i) Non-flow induced and (ii) Flow induced. For non-flow induced method, the mean binary coalescence time increased with decreasing surface free energy which was in good agreement with most of the literature review while the opposite result was found for the induced flow method. Without the presence of glycerol, the mean coalescence time was higher for induced flow method than non-induced flow method which drew a conclusion that an application of force pressing the droplets together yielded higher stability. However, the addition of glycerol reverses this effect and as a consequence, with increasing glycerol concentration, the induced flow method reduces the stability of the droplets and thus coalescence time was lower in comparison with those obtained from the non-induced flow method. The nature of glycerol on kerosene

droplets when force is not injected is to be investigated further to gain more insight in coalescence aspect. Sceptical results were found on the non-induced flow method which could be due to errors during experimentation. Further repetitions on the non-flow induced method are to be performed.

Acknowledgement
I would like to take this opportunity to first thank Dr Buddhi Hewakandamby for his utmost kind assistance and care in facilitating the progress of the experiment and my thesis. Along the way, he told various motivational short stories of his own life and shared his wisdom which filled the atmosphere with more than just pure academics. He was attentive to the various problems encountered for the past 3 months and making matters unworkable seem hopeful by nourishing me with logical ideas. I would also like to thank my parents who have been giving me the essential support and motivation to reach the few final steps of this project with a positive academic attitude. Without their support, such completion might not have been successful. A word of thanks also goes out to Katerina Loizou who shared the same experiment throughout to cut time the entire summer instead of alternating turns. Together we worked as a team and she made the long and tedious laboratory hours a bit more enjoyable through exchanging personal life experiences and cultures. I would like to thank the laboratory technicians, Phil, Mick, Terry, Fred, Marion and Vikki for getting the materials I need without any difficulties and hesitations. Appreciation also goes out to Dr. David Hann and Andy Matthews who made it possible for me to use of the high speed camera which is the one of the primary elements of the experiment. I would also like to express my gratitude to Aime for lending me the wonderful creation of his thermocouple. To Natalia and Anna, thanks for lending the pump. Also thanks to Nazrul for sharing the bits of information. Chan Yung Khiong 17th September 2010

Contents Preface..........................................................................................................1 Abstract........................................................................................................2 Acknowledgement..........................................................................................4 List of tables and figures.................................................................................8 1 Introduction...........................................................................................11 1.1 Quality energy source demand...........................................................11 1.2 Environmental issues.........................................................................12 1.3 Lifestyle...........................................................................................13 1.4 Relationship between this study and the motivations.............................14 2 Literature review....................................................................................16 2.1 Mechanisms for two phase separation..................................................16 2.2 Interfacial dynamics of droplet coalescence..........................................17 2.2.1 Interfacial rigidity......................................................................19 2.2.2 Interfacial mobility....................................................................20 2.3 Role of surfactants and other surface active agents...............................21 2.3.1 Emulsifiers...............................................................................22 2.3.2 Demulsifiers.............................................................................23 2.3.3 Other surface active agents........................................................25 2.4 Effect of interfacial tension on coalescence...........................................26 2.4.1 The Marangoni Effect.................................................................27 2.5 Role of interfacial repulsive and attractive forces...................................30 2.6 Effects of hydrodynamics on the coalescence........................................32 2.7 Drop size on coalescence...................................................................36 2.8 Thermodynamics..............................................................................37 3 Aims and objectives................................................................................39 3.1 Aims...............................................................................................39 3.2 Objectives........................................................................................39 4 Methodology..........................................................................................41 4.1 Materials..........................................................................................41

4.1.1 Rig..........................................................................................41 4.1.2 Rig core...................................................................................43 4.1.3 Attachments.............................................................................44 4.1.4 Delivery...................................................................................45 4.1.5 Image capturing and video recording...........................................47 4.1.6 Lighting system........................................................................47 4.1.7 Temperature measurement........................................................48 4.1.8 Chemicals................................................................................51 4.1.9 Others materials.......................................................................51 4.2 Experimental procedures...................................................................51 4.2.1 Cleaning..................................................................................51 4.2.2 Set up of experiment.................................................................53 4.2.3 Method of data acquisition..........................................................54 4.2.4 Interfacial tension measurements Pendant drop method.............56 4.2.4.1 5 Equations.....................................................................................58

Results and Discussion............................................................................60 5.1 Time evolution of the coalescence process...........................................60 5.2 Effect of induced flow and glycerol concentration..................................60

6 7

Conclusion.............................................................................................88 Future work...........................................................................................89 7.1 Experiment inefficiency and improvements...........................................89 7.1.1 Rig body..................................................................................89 7.1.2 Rig core...................................................................................93 7.1.3 Pumps.....................................................................................96 7.1.4 Thermocouple...........................................................................98 7.1.5 Lens........................................................................................98 7.2 Environment.....................................................................................99 7.2.1 Vibrations................................................................................99 7.2.2 Temperature............................................................................99

7.3 Data collection..................................................................................99 7.4 Image processing............................................................................100 7.5 Batch saturated..............................................................................101 7.6 Overnight.......................................................................................101 8 Bibliography.........................................................................................102

Error: Reference source not found

List of tables and figures


Table 1. Calculation of kerosene drop dimensions.............................................59 Figure 1. Water in oil emulsion under macroscopic inspection (Mullin, 2006)........12 Figure 2. Oil spillage in the sea (Solutions, 2002).............................................13 Figure 3. Emulsion used in manufacturing drugs (Pharma, 2009)........................14 Figure 4. Sequential mechanisms governing phase separation. Adapted from Paunov (2008).............................................................................................17 Figure 5. Ostwald Ripening involving droplets molecules (Bfigura, 2007)............17 Figure 6. Effect of adding surfactants on interfacial tension................................24 Figure 7. Picture showning Marangoni stress in a wine glass..............................28 Figure 8. Successive stages of the Marangoni flow in a wine glass......................30 Figure 9. Total free energy required for coalescence (Paunov, 2008)...................31 Figure 10. Components of the resultant net energy...........................................32 Figure 11. Rig body with cover, protrusion and walls.........................................41 Figure 12. Laboratory experiment set up.........................................................42 Figure 13. Dissembled rig core with cover........................................................43 Figure 14. Rig core with needles. Figure 17. Rig core in rig body. Figure 19. Pump 1. Figure 15.Top view of rig core...........43 Figure 18. Capillary attachments..........44 Figure 20. Pump 2.........................45 Figure 16. Rig core immobilized by Perspex Glass.............................................44

Figure 21. Pump 1 calibration graph................................................................46 Figure 22. Pump 2 calibration graph................................................................46 Figure 23. High speed camera Phantom v12.1..................................................47 Figure 24. High speed camera bird eyes view...................................................47 Figure 25. Dedocool lighting system................................................................48 Figure 26. Thermocouple...............................................................................49 Figure 27. LabVIEW temperature measurement................................................49 Figure 28. Thermocouple simulation................................................................50 Figure 29. Temperature calibration equation....................................................50 Figure 30. Schematic diagram of rig................................................................54 Figure 31. Image of the kerosene droplet required for the calculation of the interfacial tension.........................................................................................57 Figure 32. Dimensions of the kerosene drop needed to be determined for interfacial tension determination...................................................................................59

Figure 33. Droplets evolution at 0s.................................................................65 Figure 34. Droplet evolution at 0.256s............................................................65 Figure 35. Droplet evolution at 1.101s............................................................65 Figure 36. Droplet evolution at 1.726s............................................................65 Figure 37. Droplet evolution at 1.731s............................................................65 Figure 38. Droplet evolution at 1.732s............................................................65 Figure 39. Droplet evolution at 1.733s............................................................65 Figure 40. Droplet evolution at 1.734s............................................................65 Figure 41. Droplet evolution at 1.735s............................................................65 Figure 42. Droplet evolution at 1.736s............................................................65 Figure 43. Droplet evolution at 1.740s............................................................66 Figure 44. Droplet evolution at 1.741s............................................................66 Figure 45. Droplet evolution at 1.742s............................................................66 Figure 46. Droplet evolution at 1.743s............................................................66 Figure 47. Droplet evolution at 1.753s............................................................66 Figure 48. Droplet evolution at 1.754s............................................................66 Figure 49. Droplet evolution at 1.757s............................................................66 Figure 50. Droplet evolution at 1.758s............................................................66 Figure 51. Droplet evolution at 1.797s............................................................66 Figure 52. Droplet evolution at 1.820s............................................................66 Figure 53. Interfacial tension per area against % concentration v/v glycerol........67 Figure 54. Induced flow cumulative percentage against coalescence time graph for 0% glycerol.................................................................................................68 Figure 55. Induced flow individual percentage against coalescence time graph for 0% glycerol.................................................................................................69 Figure 56. Induced flow cumulative percentage against coalescence time graph for 0.5% glycerol...............................................................................................70 Figure 57. Induced flow individual percentage against coalescence time graph for 0.5% glycerol...............................................................................................71 Figure 58. Induced flow cumulative percentage against coalescence time graph for 1.0% glycerol...............................................................................................72 Figure 59. Induced flow individual percentage against coalescence time graph for 1.0% glycerol...............................................................................................73 Figure 60. Induced flow cumulative percentage against coalescence time graph for 5.0% glycerol...............................................................................................74 Figure 61. Induced flow individual percentage against coalescence time graph for 5.0% glycerol...............................................................................................75

Figure 62. Induced flow cumulative percentage against coalescence time graph for 10.0% glycerol.............................................................................................76 Figure 63. Induced flow individual percentage against coalescence time graph for 10.0% glycerol.............................................................................................77 Figure 64. Induced flow mean coalescence time against glycerol concentration.. . .78 Figure 65. Non-induced flow cumulative percentage against coalescence time graph for 0% glycerol.............................................................................................79 Figure 66. Non-induced flow individual percentage against coalescence time graph for 0% glycerol.............................................................................................80 Figure 67. Non-induced flow cumulative percentage against coalescence time graph for 0.5% glycerol..........................................................................................81 Figure 68. Non-induced flow individual percentage against coalescence time graph for 0.5% glycerol..........................................................................................82 Figure 69. Non-induced flow cumulative percentage against coalescence time graph for 1.0% glycerol..........................................................................................83 Figure 70. Non-induced flow individual percentage against coalescence time graph for 1.0% glycerol..........................................................................................84 Figure 71. Non-induced flow cumulative percentage against coalescence time graph for 5.0% glycerol..........................................................................................85 Figure 72. Non-induced flow individual percentage against coalescence time graph for 5.0% glycerol..........................................................................................86 Figure 73. Non-induced flow mean coalescence time against glycerol concentration. ..................................................................................................................87

Figure

73.

Non-induced time

flow

mean glycerol

coalescence concentration.

against

1 Introduction
Emulsions are crudely dispersions of fluids in another immiscible fluid phase. In the context of this project, only liquid-liquid emulsions are considered. Such typical occurrences are complex in nature and they appear in various aspects of engineering and science applications. For this reason, emulsion remains worthy to be understood through research to yield benefits in the world of engineering and science. In the next few sub-chapters, several major emulsions related industries are elaborated.

1.1 Quality energy source demand


To date, exploitation of crude oil to meet the insatiable global energy demand in the petroleum industry has advanced to the point where the era of large fields with both high quantity and satisfactory quality crude oil is at the brink of exhaustion. The most crucial issue when focusing on existing large oilfields is to increase the recovery rate. This can however be connected with typical flow assurance problems which is often related to the multiphase transportation issues of the crude oil. At many oil fields, the co-production of water from wells is in most cases substantial which can be up to an extent of 50% to 70% during extraction (). It is also known that in the petroleum industry, the issue of more than 95% of the crude oil emulsions which are in the form of stable water-in-crude oil emulsion exists as shown in Figure 1 () (). Through the extensive experiences in the petroleum industry, water-in-crude oil emulsions has nevertheless been seen to be encountered at many stages during drilling, producing, transporting and processing of crude oils. Such emulsions are extremely stable and viscous materials that increase pumping and transportation expenses, cause corrosion of pipes, pumps, production equipment and distillation columns and poison downstream refinery catalysts () which adds to the already

costly extraction process due to rise in demand () and oil resource exhaustion (). Recent studies also have shown that a significant portion of the operating cost associated with the daily oil production is spent on the oil-water separation () (). Moreover, the performance and properties of crude oils are to a great degree determined by the presence of water () (). Minimizing the water levels in the crude oils by oil dehydration method can reduce pipeline corrosion dramatically, improve crude oil quality and maximize pipeline usage for transportation (). Crude oil free from water is mandatory for pipeline flow and refinery operations. The emulsion ageing tends to increase its stability; the breaking must be carried out as soon as possible in the production facility close to the well ().

Figure 1. Water in oil emulsion under macroscopic inspection ().

1.2 Environmental issues


In both offshore and onshore oil processing, one of the most challenging environmental problems today is the effective removal of stable oil deposits from water sources such as reservoirs, oceans and seas which is essentially necessary for industrial and domestic consumption. These are known as oil-in-water emulsions stabilised by naturally occurring surfactants. Offshore oil spillage as shown in Figure 2, runoff oil released during oil well extraction and oily wastewater from other industries such the recycling industry and water desalination are undesirable events contribute to this concern. The oil content which has low biodegradability must be

separated as much and quick as possible from the aqueous phase to prevent mass ecological problems.

Figure 2. Oil spillage in the sea ().

1.3 Lifestyle
Daily products in the olden days have shifted from simple to sophisticated genres with variety if improvements in the new millennium. As such, the demand for such astonishing materials has increased tremendously. One of the sources for such complexity often can be subjected to the incorporation of numerous components into a product so that the benefits of each constituent may be garnered simultaneously for product enhancement. This is often desirable due to the resultant change in tactile properties, for instance in mayonnaise and other foods which are oil-in-water emulsions. Alternatively, there may be a more sophisticated purpose such as when an oilsoluble material must be delivered to an aqueous environment. In this case, the material may first be dissolved in an organic liquid which in turn is dispersed in an aqueous medium to create the emulsion. Such schemes find application in drug delivery as shown in Figure 3. To create dispersion they must inherently be insoluble in one another and so a third component usually must be added to prevent them from phase separating. Other products include, lubricants, cosmetics

and paints (). Such technology is often the correct entry for emulsion applications. A difficulty that arises in such applications however is the incompatibility of the materials making up the emulsion.

Figure 3. Emulsion used in manufacturing drugs ().

1.4 Relationship between this study and the motivations


In a nutshell, there is an endless list of activities that make use of the emulsion technology to function in the designated way. While effective separation of unwanted oil dispersions from water and water from crude oil has long been a challenging technical task in the petroleum, oil spills and water industries, such emulsions or dispersions are useful in the food and pharmaceutical industries. At present, there exist several oil-water separation methods which are the current state of the art technologies. However, most of them have rather limited applications in separating oil form the produced fluids due to their large capital and or operating costs, low separation efficiency and or slow separation process long residence time. Therefore, some significant improvements in oil water separation techniques are still required in practice. This applies to the food and pharmaceutical and other industries as well to improve the quality of the products. During these industrials operations, the emulsion drops are subjected to flow fields which may cause them to break or coalesce. The simultaneous occurrence of the

two phenomena controls the drop size distribution, which in turn has significant effect on the processing conditions and characteristics of the product. The addition of components such as surfactants can cause the inversion of different phases. Moreover, the lifetime of emulsions may vary from seconds to minutes to hours to days to weeks to months and to years depending the nature of the surfactants, nature of both fluids and their volume ratio. Despite the large amount of work devoted to this issue, predicting the emulsions lifetime still remains a challenge. It is necessary to go back to the science of the process to understand the emulsion stability mechanisms from both macroscopic and microscopic points of view. Understanding the underlying principle mechanisms of one type of emulsion with regards to its stabilization strength and formation can improve the design of the phase separator equipments and demulsifying/ emulsifying agents or techniques so that effective separation of the dispersed phase or emulsification can be made successful. Generally, an improved understanding of coalescence behaviour is a prerequisite for better engineering and thus would potentially allow one to improve formulations, quality, stability, process yields and stability during product shelf-life, through engineering of interfaces, bulk rheology, and process. In particular, it is the interfacial dynamics between the droplets that is of great interest in this project. The interfacial dynamics involve the mechanisms that governed the probability of coalescence between the droplets and are the very first several things to look at, in the beginning of an emulsion research.

2 Literature review
In the field of liquid-liquid emulsion, the behaviour of coalescence in many forms such as between a pair of drops, a pair of bulk phases, many drops, drop and bulk phase have been increasingly popular and continuously studied theoretically and experimentally in detail for more than 50 years now. Investigators have conducted experiments in many possible ways with either pure chemicals or in the presence of added components that affect the interfacial properties. The effects of varying interfacial tension, hydrodynamics and other conditions such as temperature and pH levels were investigated to confirm the proposed coalescence mechanisms and to characterize the coalescence behaviour of various liquids systems.

2.1 Mechanisms for two phase separation


From the visible perspective regardless of the presence of surfactants which is discussed later in Section 2.2, an emulsion will eventually phase separate given sufficient time without introducing external forces or energy. Figure 4 shows the generally accepted time evolution of an emulsion breaking occurrence (phase separation). The four mechanisms responsible in destabilizing the emulsion: Flocculation The droplets form aggregates of two or more drops. The interparticular distance between the droplets is strongly diminished due to a net attraction between the droplets.

Creaming Due to the difference in the density between the dispersed phase and continuous phase, this results in the formation of a dispersed phase concentration gradient in the mixture.

Coalescence This significant occurrence involves specifically the amalgamation of two drops.

Ostwald ripening In more microscopic terms as shown in Figure 5, the diffusion of droplets molecules through the medium causes small droplets to decrease in size and disappear while large drops grow. It was found that this process actually takes place prior to coalescence ().

Figure 4. Sequential mechanisms governing phase separation. Adapted from Paunov ().

Figure 5. Ostwald Ripening involving droplets molecules ().

However with respect to the field of this research project, only the coalescence stage is of concern. Therefore, the interaction between the two droplets at very close proximity to each other especially at dynamics at the interface is reviewed in this section.

1.1 Interfacial dynamics of droplet coalescence


According to an early theory developed by Smoluchowski (), two drops merge and coalesce immediately upon collision straightforwardly entailing that there is no resistance to coalescence imparted by the thin liquid film trapped between the drops (). However, in practice, this theory is only applicable only when the continuous phase has very low viscosity and there is no surface active species

present in the medium that can stabilize the drops. Also, the aforementioned theoretical work by Smoluchowski is only applicable to non-deformable particles (). However, many studies now also include deformable droplets where the viscosity of the dispersed phase or droplets is low. It is generally accepted that the coalescence behaviour of approaching emulsion droplets is said to be controlled by the dynamics of the film between their surfaces. Zapryanov et al () reported that for both types of oil/water emulsion, coalescence between two droplets occurs in three specific steps: (1) Approach of the droplets through the continuous phase where the droplets eventually come into contact driven by the applied forces of the flow field or induced flow; (2) Deformation of the droplets by flattening of their contact surfaces to form a thin film between them which starts to drain through a process commonly known as film drainage; (3) Thinning of this film to a critical thickness below which the droplets coalesce due to intermolecular forces like van der Waals forces become dominant and cause the film to rupture followed by the formation of the bridging or merging of the two droplets. Danov et al() and Aarts and Lekkerkerker () supported the sequential mechanisms listed above and investigated the dependence of external forces on the interaction of the droplets such as van der Waals, electrostatic, hydrodynamics and also the energy for the deformation of the droplets which are also studied by many researchers which are discussed later. Auflem () also reported that the main mechanisms that influence coalescence are film drainage and film rupture which are considered acceptable mechanisms in the vast literature. Despite such logical

hypothesized mechanisms, the coalescence probability depends only on the details on the drainage of the film between the droplets. However, with regards to the film thinning occurrence, many have thought that this film should be nevertheless invisible to the naked eye until Pu and Chen () investigated on the surprising phenomenon of jumping coalescence of two largely separated water droplets of unequal size under microgravitation where essentially no external force acts which defies largely against the film thinning theory. In their work, coalescence was successful even when the continuous phase film was seen to be very obviously larger than even the size of the drops. It was concluded that the coalescence driving force within the liquid drops under microgravitation to be responsible for the jumping phenomenon but no further investigation was attempted at such unusual phenomenon under microgravitation. However, quite recently, results obtained from a study by Kim and Longmire () might be able to lead clues to the coalescence under microgravitation. They concluded the vortex rings within colliding drops must be oriented such that they induce streaming flow with a strong component toward the centerplane and the drops must collide with sufficient inertia such that they deform significantly, increasing the velocity magnitude in the streaming flow and also if the large velocities in the streaming flow approach the center plane, then they can induce a faster outflow in the thin film between the drops.

1.1.1 Interfacial rigidity


Often in many emulsion studies, surface rigidity plays as one of the most important dynamics that governs the coalescence probability (). Surface rigidity reflects the tendency of the interface to deform as the two drops approach each other. In the aspect of forces in the vicinity of the interacting drops, it is obvious that the separation distance between the two facing surfaces at which the deformation takes place, increases and decreases with attraction and repulsion respectively i.e. the

higher the attractive force between the droplets, the larger the distance at which deformation starts to occur. Perhaps, the drainage of the thin film may be analyzed by means of a force balance comprising the force exerted on the droplets by the flow field and the resistance to drainage due to the viscous flow in the film. Ivanov et al () observed that often in the beginning of the deformation the drop caps acquire a bell-shaped form called a dimple. Ivanov and Dimitrov () added that this film thins with time and at a certain critical thickness ruptures as observed by Zapryanov et al (1983). Chesters () concluded that the rigidity of the colliding interfaces of the drops governs the thinning rate of the liquid film to its critical thickness. Specifically, the interfacial rigidity determines how much the colliding interfaces will flatten and is influenced by drop size and interfacial tension. If the film can be thermodynamically stable, the thinning stops at the equilibrium thickness where the disjoining pressure in the film equilibrates the capillary pressure in the drops. In the case of thermodynamically unstable films which always rupture: then the film lifetime depends mainly on the rate of thinning and the critical thickness. If the interface is undeformable, which means that the pressure in the liquid film is lower than the Laplace pressure inside the drops, the film can easily be drained.

1.1.2 Interfacial mobility


The interfacial rigidity is often related to the interfacial mobility. The flow in the film is coupled to the flow inside the approaching particles via the mobility of the interfaces, which is especially relevant for pure fluids. Abid and Chesters () observed that the interfacial mobility is governed by the tangential stresses exerted on the film by the drops. Saboni et al () found that this tangential stress depends on the viscosity ratio of the dispersed and continuous phases. There are three types of interfacial behaviour during coalescence identified based on the

dispersed/continuous phase viscosity ratio: (1) immobile interfaces, when the

dispersed phase viscosity is much higher than the continuous phase one, or when surfactants are present that retard the drainage of the liquid film; (2) fully mobile interfaces, when the continuous phase viscosity is very large compared to that of the dispersed phase; (3) the most commonly encountered partially mobile interfaces, when the viscosity ratio is moderate ().

1.2 Role of surfactants and other surface active agents


Previously the information garnered only indicated the interfacial dynamics without specifying whether the dynamics of the coalescence behaviour are the

consequences of the addition of surfactants or otherwise. The role and types of the surfactants should be clearly understood beforehand. According to Bancroft () rule, a liquid containing the surfactant becomes the continuous phase. In other words, when dispersing equal volumes of liquids for instance oil and water, the emulsion obtained is oil-in-water if the surfactant is more soluble in water and vice versa. Surfactants are substances that alter the interfacial tension between the two or more phases of liquids and by altering the interfacial tension; this can influence the coalescence behaviour very much which is discussed in Section 2.4. However, it is important to distinguish that there are two types of surfactants emulsifier and demulsifier. Many articles have confused readers where the authors did not specifically draw the line between emulsifiers and surfactants; they indicate that surfactants are emulsifiers and the other way around. Perhaps it is crucial to review the simple definitions of the types of surfactants available in the chemical industry. It remains very important to identify the surfactant type because different categories have different effects ().

Surfactants commonly are divided into two families (i) emulsifiers and (ii) demulsifiers. An emulsifier when added into a mixture of oil and water stabilizes the emulsion. Hence, emulsions are dispersions of two immiscible liquids, kinetically stabilized by the action of a surfactant or specifically an emulsifier. The function of the demulsifier can only work to break the emulsion which is already stabilized. In other words, demulsifier can only be added to a mixture which is already emulsified or stabilized by natural occurring or induced emulsifiers. Hence demulsifiers are emulsion breakers that separate for instance, crude oil emulsion into distinct oil and water phases (). However, demulsifiers can also be added to emulsion free of emulsifiers to facilitate phase separation.

1.2.1 Emulsifiers
In the absence of emulsifiers, drops aggregate rapidly as a consequence of the van der Waals force. In their presence, the emulsifier particles adsorb to the interface of the drops creating a barrier that decelerate aggregation and Ostwald ripening preventing the higher probability of drops coalescing. Hence, they favour the occurrence of immobile interfaces, delaying the drainage of the intervening film between flocculated drops. Depending on their interfacial properties, these films can drain and rupture after a period of time significantly longer than that of without the presence of emulsifiers or, remain stable for long periods of time () (). Conversely, emulsifiers lower the interfacial tension of these films, favouring the appearance of surface oscillations and holes which happens to reduce the drainage rate (). Hodgson and woods() investigated the effect of additions of SDS on the coalescence of oil toluene drops in water. Most of the surfactants used by the aforementioned researchers previously are regular emulsifiers which can be found in a handbook by Mukerjee and Mysels (). Chen and Pu () investigated on the addition of an emulsifier concentration and found out that the binary coalescence time was increased under microgravitation. They concluded that a film is apparent

when sodium sulphate dodecyl (SDS) was present at a higher level concentration. This film prevents the coalescence of the droplets. There are other works that used regular emulsifiers to investigate on the coalescence behaviour of various liquid systems and yielded similar outcome () () () () () () ().

1.2.2 Demulsifiers
Based on the amount of work done on liquid coalescence using surfactants, compared with emulsifiers, the work done using demulsifiers was not as extensive. Few researchers developed similar scientific hypothesis that supported the replacement of the emulsifiers by demulsifiers at the interfaces and the

consequence of this is the increased facilitation of coalescence. Demulsifiers are surfactants found to develop high surface pressure at the crude oil-water interface and promote phase separation (). It was known that during the demulsification process, the effective demulsifiers replace the emulsifiers adsorbed on the oil water interface (). Deng et al () added that the presence of demulsifiers results in replacement of rigid film of natural crude oil emulsifiers by a film which is conducive to coalescence of water droplets. When two drops approach each other due to external forces, the thickness of the intervening continuous phase film decreases. The shear stress associated with drainage tends to concentrate the emulsifier molecules outside the film and their concentration inside the film is lowered. Thus an interfacial tension gradient is set up with high interfacial tension inside the film and low tension outside the film. Such phenomenon is discussed in more detail in Section 2.4.1. The demulsifier molecules migrate from the surface adjacent to the interface where the demulsifier molecules are adsorbed in the spaces left by the emulsifier molecules in the film so the interfacial tension falls quickly. Once the surface layer is depleted, demulsifier molecules have to diffuse from the bulk which is a slow process and therefore the interfacial tension falls more slowly. Ultimately the interface becomes saturated and

the equilibrium interfacial tension is reached. As indicated by the variation in interfacial tension with time, the rate of adsorption of the demulsifier at the crude oil water interface is much faster than that of emulsifiers in the crude oil. Adsorption of the demulsifier reverses the interfacial tension gradient and enhances film drainage. Ultimately, a stage is reached when the film becomes very thin and due to the proximity of the dispersed phase the van der Waals forces of attraction dominate and the droplets coalesce. More presently, Abdurahman and Yunus () confirmed with higher concentration of the demulsifier used, the droplet size was larger than that of with lower concentration. It was also observed that increase in demulsifier will accelerate the coalescence of droplet faster. But regardless of concentration levels, the presence of a demulsifier is characterized by very short initial coalescence time. Demulsification was also studied by Borges et al (), Rondon et al () () () () ()

Figure 6. Effect of adding surfactants on interfacial tension.

1.2.3 Other surface active agents


Other surface active agents have been used as well. In crude oil emulsions, it has been known that they are stabilized by naturally present substances such as asphaltenes and resins which behave like emulsifiers (). Authors McLean et al ()

were reported that, asphaltenes and resins adsorb at the water-oil interfaces and form interfacial films that confer stability against phase separation or in other words these surfactants provide steric hindrance to droplet-droplet coalescence which are similar to functions of the regular surfactants or specifically emulsifiers. The stability of water in oil emulsions in petroleum systems using the asphaltenes were also investigated by other researchers () () (). Such mechanism was also similarly described by Sundararaj () with polymeric materials () which also act as surfactants and they are called compatilizers. Electrolytes (), salts like sodium chloride NaCl () () and even gelatine () were also used as surfactants to investigate coalescence behaviour or droplets and yielded the same stabilization effects as emulsifiers. The paradigm that indicates that the water-in-oil emulsions are stabilized by asphaltenes often seems to be a gross oversimplification without any specific details regarding the true observable contribution of asphaltenes. Because of this, Czarnecki and Moran () created a model to explain the mechanism of water-in-oil emulsion stabilization in petroleum systems in order to provoke further potential discussions. Their model suggested that only a small sub fraction of asphaltenes and not all is being used as the stabilization. There appears to be another chemical responsible for the stabilization which is a low molecular weight surfactant material. The competition for the oil/water interface between these two substances is based on the difference of their adsorption kinetics where the asphaltenic material adsorbs slowly and irreversibly and forms rigid skins while the other material adsorbs faster. Abdurahman and Yunus () also investigated the dependency of co-surfactants. For instance, by increasing the ratio of resin concentration to asphaltene concentration effectively increases the water separation from crude oil emulsion while an increase in the asphaltene concentration in the crude oil decreases water separation rate. The effect of resin is that it solubilises the asphaltenes into the oil phase minimizing asphaltene interaction with the water droplets and hence more coalescence.

1.3 Effect of interfacial tension on coalescence


Previously, we have seen only the decrease of magnitude of the interfacial tension of fluid systems by addition of surface active or surfactants. In this section, the effect of altering the interfacial tension on the coalescence behaviour is elaborated in more detail. Hartland and Wood () concluded that with decreasing interfacial tension, the film drainage rate will be reduced, which leads to longer coalescence times. They also found that during the coalescence of a liquid drop with a flat interface the drainage rate decreased with a decrease in interfacial tension and applied force or with an increase of the drop volume. Li and Slattery () confirmed that the coalescence time of nitrogen bubbles increased when the surface tension was decreased by changing the concentration of sodium chloride in the aqueous continuous phase. So far, many researchers have only found the effect of decreasing the interfacial tension of binary liquid/liquid systems and gas/liquid systems by addition of emulsifiers on the coalescence frequency () () () () () () () () () () () () (). On the other hand, Abdurahman et al () found that by using many demulsifiers, which did not only decrease the interfacial tension, but increased the interfacial tension and promoted separation of oil and water phases. Surprisingly, Wang et al () found that by decreasing the interfacial tension, the coalescence time decreases which is a contrast to the present literature. Such unusual findings are not very common in the literature.

1.3.1 The Marangoni Effect


Several researchers investigated into more detail with regards to the mechanisms responsible for the low probability of coalescence in the presence of certain emulsifiers. The most commonly accepted idea is that the flow out of the thin film produces a gradient in the surfactant concentration, with the lowest concentration

at the centre of the film and the highest ones near the edge of the film (). This yields the Marangoni stresses that are hypothesized to immobilize the interfaces within the film, and thus slow the drainage process () under micrograviation. This was also previously observed in a recent dissertation by Yoon (). Other previous researchers also yield such phenomenon () (). With surfactant concentrations being the highest at the edge this yield the surface tension at the edge of the film to be the lower while the surface tension at the centre of the film is highest. According to the Marangoni effect which was first observed by James Thomson (), and later in more detail by Marangoni () suggests that the film moves from region of low interfacial tension (emulsifiers concentration is high) to high interfacial tension (emulsifiers concentration is low) causing a film at the centre of the point of contact between the drops. However, Dai , Graham and Leal () showed that during coalescence, the interface in the thin gap actually exhibits a significant degree of mobility which explains that at least to a certain extent, the fact that the assumption of a completely immobilized interface is not in agreement with their experimental conclusions. Hence the

unexpected new result is that the role of Marangoni effect on the coalescence process does not occur via immobilization of the interface within the thin gap region, but rather is due to its effects on the hydrodynamics outside the thin film. In particular, Marangoni stresses immobilize the drop interface outside the thin film, and this increases the total external hydrodynamic force that pushes the drops toward each other. This, in turn increases the degree of flattening and dimpling of the thin film, and it is primarily this change that slows the film drainage process and thus increases the required drainage time prior to coalescence. To help understand the Marangoni mechanisms better, it is best to demonstrate the appearance of the wine tears.

Figure 7. Picture showning Marangoni stress in a wine glass.

Stage 1 Alcohol is first spread across a curved dish. Stage 2 As evaporation occurs everywhere along the free surface, a film can be seen around the outer edges of the dish. Stage 3 The film seems to spread out further. The alcohol concentration in the thin layer is thus reduced relative to that in the bulk owing to the enhanced surface area to volume ratio. As surface tension decreases with alcohol concentration, the surface tension is higher in the thin film than the bulk. The associated Marangoni stress drives up flow throughout the thin film Stage 4 The alcohol thinly spread across the curved dish climbing up to reach the top of the dish where it accumulates in a band of fluid that thickens. Stage 5 Eventually the thick band becomes gravitationally unstable and releases the tears of wine. Marangoni flows are created successfully as tensions draw the liquid towards the centre of the dish.

Stage 1

Stage 2

Stage 3

Stage 4

Stage 5
Figure 8. Successive stages of the Marangoni flow in a wine glass.

However, it seems that the Marangoni effect is only applicable to emulsifiers which happen to cause the increased stability of drops. With respect to demulsifiers, no

such effect was reviewed. Perhaps, with demulsifiers, the reason for the promotion of coalescence frequency can be due to their natural structure which is beyond the scope of this project. Previously, it was mentioned that the most demulsifiers decrease the interfacial tension of liquid systems which is in common with emulsifiers. With this, if the Marangoni stresses hold true, the demulsifiers replace the spaces left by emulsifiers as the thin film is being squeezed out so quick that the concentration of demulsifiers at the centre of the film is very large compared to that at the edge of the film and hence the reverse of interfacial gradient which promotes coalescence. However, this does not apply to demulsifiers which increase the interfacial tension which was observed by Abdurahman et al () because it would defy the coalescence role of demulsifiers.

1.4 Role of interfacial repulsive and attractive forces


The coalescence behaviour of emulsion droplets upon collision depends on the interplay between these two types of forces. If the total interaction is very repulsive, the droplets can rebound; if the intermolecular repulsive forces can hold the film at an equilibrium separation, the droplets may flocculate; and if the total interaction is attractive, the film will become thinner and finally rupture, i.e. the droplets coalesce. Coalescence requires rupture of a thin liquid film between the drops. In most real-life applications there are effects associated with the thinning of the liquid film and interfacial repulsion originated from Derjaguin Landau Verwey Overbeek (DLVO) and non-DLVO forces. Such DLVO forces are the van der Waals and electrical double layer forces which have received much studies () whereas non-DLVO forces are attractive depletion steric and repulsive structural forces () which on the other hand did not receive much attention.

The condition for rupture is the critical film thickness such that the attractive force which is the van der Waals forces must be larger than the electrostatic repulsive force and the Marangoni effects as discussed in Section 2.4. This is illustrated in Figure 9 where the total free energy of the particles to be at very minimal distance must be at most at zero Joules. k is the Boltzmann constant and T is the temperature. D is the distance between the two surfaces of the drops facing each other and U is the free energy of the particles.

Figure 9. Total free energy required for coalescence ().

Figure 10. Components of the resultant net energy.

The red line in Figure 9 shows the net energy which is composed of the double layer repulsive force and the van der Waals attractive force as shown in Figure 10. Hence in order for coalescence to commence, the net energy curve must be below zero value at the vertical axis. Many researchers failed to take into account their significance when characterising the model for coalescence behaviour. Because other relevant forces play an important role in determining the stability of emulsion droplets, neglecting them (van der Waals) will unavoidably lead to deviation from the real situation (). Chen () confirmed that the van der Waals disjoining pressure destabilized the film, whereas the electric double layer stabilized it using a model that describes the film profile evolution between two equal sized drops and predicts the film stability, time scale and film thickness given only the radius of the drops and the required physical properties of the fluids and surfaces.

1.5 Effects of hydrodynamics on the coalescence


While some researchers focussed on coalescence under stagnant flow, several other researchers investigated droplet coalescence under flowing conditions. For instance, Hartland and Wood () found that during the coalescence of a liquid drop with a flat interface the drainage rate decreased with applied force. Many research groups have studied theoretically and numerically the film drainage process under constant approaching velocity or constant driving force () ()() () () (). Their results indicate that the film drainage rate is mainly controlled by the interfacial tension, the viscosity ratio between the dispersed and suspending phases and the external force. These factors have profound effect on the drainage rate, which often depends on the deformability and tangential mobility of the droplet surface as discussed previously. Wang et al () criticized that most of the models assumed simple boundary conditions, such as constant interaction force or constant approach velocity. In reality both will vary during the collision of a drop with a flat interface or with another drop.

Al-Mulla and Gupta () concluded that using a Coutte device at lower shearing rate favoured coalescence. Similarly, Nandi et al () indicated that their surfactant stabilized emulsions were prepared in a stirred tank and low shear rates prevented drop breakup and at high shear rates, the emulsions were stabilized further. However, Nandi et al ()extended their shearing experiments with surfactant-less liquid mixtures and found that without surfactants, higher shearing rates promoted coalesce. Chen and Tao () also found that by increasing the stirring intensity it increases the oil in water emulsion stability. Previously, Leal () added the film drainage is slowed down upon addition of emulsifiers in a four-roll mill where the binary coalescence of a pair of droplets were exposed to external shearing flows. Dai and Leal () validated the film drainage and rupture using polymeric materials in which the experiment was based on the coalescence of a pair of equal size drops undergoing head-on collision in a biaxial linear flow under constant linear velocity. They also investigated that there is an equally important effect which is due to the increasing hydrodynamic force pushing the drops together causing the film to be more strongly deformed into a dimpled configuration which further slows the film drainage process. They also confirmed the validity of the Marangoni effects such that the immobilization of the interface within the thin film as expected and the force pushing the drops together is increased by the Marangoni immobilization of the interface outside the thin film. This caused the thin film to be much more dimpled and deformed than it is in the absence of surfactant at the same capillary number and the more dimpled film shapes slow the rate of film drainage. An observation holds true in this respect in the opposite manner in which, Bremond and Bibette () investigated the high affinity of a pair of emulsion drops to favour coalescence under decompression or rather, separation. Briefly, the experiment was undergoing first expansion and then decompression of a flow of trains of emulsion droplet pairs passing through a coalescence chamber. No coalescence was observed until the decompression of the downstream droplet of the pair. It was also

observed that at this instant, both droplets form a pair of facing nipples in the contact area prior to coalescence. The separation term is used because the action of decompressing the downstream droplet leads to acceleration, causing it to move away faster from the upstream droplet. Gary leal () () () studied the stability of film drainage with respect to flow induced coalescence. They investigated on the growth of disturbance in relation with the film drainage phenomenon. Contrarily in the same perspective, the demulsifying experiment conducted by Abdurahman and Yunus () on different stirring rates concluded that with higher stirring rates and hence, larger turbulence and high shear rates, the average droplet size was larger. This relatively means that shear induces higher drainage rates. However, despite the obvious introduction of shear stress into the medium, no significant observation was made on the flow patterns. The capillary number, which is the ratio between viscous and interfacial forces, characterizes the droplet deformation and plays an important role in determining whether two droplets coalesce upon their collision. It was found that the coalescence behaviour of two droplets in a simple shear flow strongly depends on the capillary number. Coalescence only occurs at a capillary number that is lower than a critical value. Also in a simple shear flow, Loewenberg and Hinch () numerically studied the collision between two deformable droplets. They found that the collision behaviour is dictated by capillary number and predicted that, for flow with capillary number very much less than 1, coalescence tends to occur when the droplets are pressed together, whereas for capillary number ~ 0 the tendency for coalescence reaches its maximum when the droplets begin to separate in the extensional quadrant. The latter behaviour was experimentally verified by Guido and Simeone (). In a different flow fashion where a pair of equal size drops was on a head-on collision, the findings were very similar (). Leal () investigated on the

effect of capillary numbers on the coalescence of the droplets and added that coalescence requires very gentle collisions, i.e., collisions at low capillary numbers. A simplistic view of the coalescence process begins with the observation that the thin layer of fluid that separates the two drops once they collide must become thin enough for van der Waals attraction to destabilize the film to produce film rupture. During the whole collision process, the net force on each drop is actually zero assuming that they are neutrally buoyant. However, when they are in close proximity, it is convenient to think of the net force as being the sum of two equal magnitude but oppositely directed forces; one the hydrodynamic force due to the external flow which tends to push the drops together when they first come into apparent contact, and the other the lubrication force due to the extra pressure in the thin film. From this point of view, the question of whether a pair of drops can coalesce is then the question of whether the film between them thins sufficiently before the drops rotate to the orientation where the external force changes from pushing the drops together to pulling them apart. Leal () investigated such phenomenon in a study of the effect of the capillary number on droplet coalescence in four roll mill device which generates flow to the surrounding fluid. The device is advantageous because there is a stagnation point at the geometric centre where coalescence and drop breakup experiments can be conveniently carried out and monitored. It was found that at a capillary number larger than the critical capillary number, the pair of droplets was first in a horizontal orientation. Then the orientation varied such that there was an increase in inclination of the droplets and the droplets were getting closer to each other. However, at an angle of 45 degrees, the droplets move away from each other due to the influence of the external forces. In contrast, at a capillary number lower than the critical capillary number, the droplets coalesced at an angle of inclination of only 10 degrees. It is also important to note that at horizontal level, after the collision, the droplets rotated afterwards.

Such are glancing collision, where the two drops rotate in the flow from the point of initial contact to the point where they hydrodynamic force along the lone of centres changes sign and the drops begin to separate. The most important finding here was that the influence of the capillary number. Hu et al () studied the collision between two droplets of equal size in a simple shear flow. They showed that, for a fixed initial position and viscosity ratio, the minimum separation between two droplets is solely determined by the capillary number.

1.6 Drop size on coalescence


It is commonly accepted that with large drops the large contact area reduces the film drainage rate resulting in larger coalescence time (also called rest-time) compared to small drops. Results produced by Dreher et al () show that the coalescence time depends almost linearly on drop size. Moreover, Mackay and Mason () found that for easily deformable drops, coalescence time and drop stability increase with drop size and was later validated by Chen el al (). Hartland and Wood () also found that during the coalescence of a liquid drop with a flat interface the drainage rate decreased with an increase of the drop volume.

1.7 Thermodynamics
Surface free energy is another term in the aspects of thermodynamics for interfacial tension. There information regarding the interfacial mechanisms which can still be considered to be under proposition or being hypothesized because there is no concrete or credible proofs that can defend its existence. However, the only basis that we can still rely on is the change of the surface free energy which changes with the curvature of the droplets. The curvature of the droplets is related to the interfacial tension and this is not only available in the literature but also logical and visible to the eye. By incorporating thermodynamics, one can visualize the coalescence behaviour in terms of the period of coalescence by changing the

interfacial

tension.

With

this

and

the

established

numerical

equations

of

thermodynamics through Gibbs free energy, which can link the surface curvature to the concentration of surfactants, then there is solid verification that the presence of surfactants can increase or decrease the coalescence time. Following the conclusions obtained for published articles, the behavior of a system under the influence of a surfactant is to be observed. The surfactant is expected to lower the interfacial tension of the system, keeping in mind that the molecular weight of the surfactant to be used must be low, so that coalescence is enhanced. Thermodynamics can be utilised to speculate the behavior as shown below:

Helmholtz equation: F=U-TS Differential of Helmholtz equation at constant temperatures : dF=dU-TdS Internal energy: dU=TdS+dn+ dA (3) Combining equations (2) and (3) : dF=dn+ dA. (4) where, F= Helmholtzs Free energy, U= internal energy T= temperature, S= entropy, = chemical potential, n= number of species,=interfacial tension and A= surface area. (2) (1)

As a conclusion, lowering the interfacial tension and keeping the system constant, will result to the decrease of dF, thus the energy required to create a system after the spontaneous energy transfer from the environment has taken place, is lowered and coalescences is favored.

2 Aims and objectives


2.1 Aims
The aim of this project is to investigate the coalescence behaviour of oil-in-water emulsion under the influence of surfactants and the presence of stagnation in the vicinity of the region of coalescence. It is crucial to identify that the coalescence process be it successful or unsuccessful, to be free of external forces other than the generation/induction of flow to allow the growth of the droplets. This will involve the study of the interactions between the close proximity facing surfaces of a pair of oil droplets suspending in water undergoing coalescence or otherwise through a high speed camera. Observations of the physical changes of the interface between the pair of oil droplets are to be validated with the available literature to further support or criticize the suggested mechanisms and wherever possible, draw new conclusions. With these contributions, more understanding of the coalescence behaviour can be gained. This is beneficial to both the science and engineering fields in the context of

better design for emulsion breaker devices or techniques and the synthesis of better and more efficient demulsifying chemicals.

2.2 Objectives
The behaviour of these coalescences is variable due to the presence of factors such as external hydrodynamic forces and surfactants. The specific contributions made by each factor with regards to the coalescence behaviour are unpredictable. Hence there are many potential investigations available to investigate the dependence of each factor on each other. The objectives of this project are to: Investigate the effect of different surfactants on the coalescence frequency Investigate on the effect of surfactant concentration on the interfacial tension of the system Investigate on the relationship between interfacial tension and coalescence frequency Investigate the effect of induced and non induced flow on the coalescence frequency

1 Methodology
This section brings about the details in which how the experiment was carried out, what materials were used and their sources and what were measured. The experiments of binary drop coalescence were carried out using the experimental set up as shown in Figure 1. Figure 2 shows the whole experimental set up. For step by step procedures, this is available in the Appendix.

1.1 Materials
1.1.1 Rig
The coalescence cell or the rig shown in Figure 11 is a vertical rectangular acrylic cell with a square cross section of side length 26 cm and height 25 cm. The rig is made out of square and rectangular pieces of Perspex glass glued together using thermal glue and let dried. The bottom centre of the rig has a protrusion made out of Perspex glass which is also glued to the bottom. The protrusion is made such that the rig core can be inserted without any mobilisation. The rig has a cover to prevent the entry of dust particles during experiment.

Figure 11. Rig body with cover, protrusion and walls.

1.1.2 Rig core


The rig core consists of several components combined together including a 4 screws, cover, needles and connectors. The rig core was a cube of sides 7cm with a vertical cylindrical hollow centre which was designed to fit in through the protrusion

in the rig as shown in Figure 13. Two Perspex glasses of dimensions are cut and inserted into the bottom of the rig to prevent any movement of the rig core during experiment in Figure
Figure 12. Laboratory experiment set up

as

shown

14.The rig

core also had four V-shaped grooves to fit in the plastic needles such that the rig core cover is pressed firmly with screws on top of the rig core body to lock in the

plastic needles as shown in Figure 14 and 15. By adjusting the tightness of the screws, this adjusts the levels of the tip of the pair of needles. The needles were bent at 90o in the mechanical workshop as accurate as possible. The outer diameter of the needles was 0.966mm. The tip of the vertical needles should be aligned with each other horizontally and in line with each other so that droplets can be produced equally with sides touching each other.

Figure 13. Dissembled rig core with cover.

Figure 14. Rig core with needles.

Figure 15.Top view of rig core.

Figure 16. Rig core immobilized by Perspex Glass.

1.1.3 Attachments
Drops were formed through two polyethylene capillaries with outer diameter 0.7mm attached to the needle on one end and the other to the syringe connector as shown in Figure 16. Each capillary was connected to the inlet of the needle at one end and the other end to the syringe connectors shown in Figure 17. Hence when the rig core with its attachments was placed within the rig, care was taken not to damage or apply pressure on the vulnerable capillaries. The cover was slid open and a thin film of cling paper is used to cover the opened part of the cover to prevent entry of dust particles as shown in Figure 18.

Figure 17. Rig core in rig body.

Figure 18. Capillary attachments.

1.1.4 Delivery
Two different infusion pumps were used due to lack of resources: (i) Pump 1 and (ii) Pump 2 as shown in Figures 19 and 20 respectively. Due to this reason and the unequal inefficiencies of both pumps, pump calibration was necessary. The pump calibration was carried out such that the true mean volumetric flowrate was calculated for a particular pump setting flowrate. The methodology of the pump calibration can be viewed as the volume of liquid discharged per unit time for a set of pump setting flowrates (0.1ml/hr 1.0ml/hr). The liquid used for pump

calibration was RO water. 5mL syringes were inserted securely into the infusion pumps and connected to the syringe connectors as shown in Figures 18 and 19. The desired volumetric flowrate of the pump is checked through the pump calibration curve obtained from pump calibration. For instance, if 0.5ml/hr was needed to induce flow, then the corresponding pump setting flowrate at the x-axis of the pump calibration graph with error bars as shown in Figures 21 and 22 was used.

Figure 19. Pump 1.

Figure 20. Pump 2.

This configuration allows equal and steady delivery in each syringe of low inlet flowrates required in this work. Throughout the experiment, a mean setting flowrate of 0.5ml/hr was used corresponding to setting flowrates of 0.63ml/hr for Pump 1 and 0.62ml/hr for Pump 2.

Figure 21. Pump 1 calibration graph.

Figure 22. Pump 2 calibration graph.

Syringes were purchased from the Medical School University of Nottingham.

1.1.5 Image capturing and video recording


The coalescence was captured with a high speed video camera (Phantom v12.1) placed in front of the rig, as shown in Figures 23 and 24 which has a CMOS sensor with a maximum resolution of 1280 800 pixels. The average recording rate used throughout the coalescence process was 1000 pps, while even higher recording rates up to maximum 6200 pps at maximum resolution (1280 800 pixels) were necessary to capture the formation of the liquid bridge between the two drops as soon as coalescence started.

Figure 23. High speed camera Phantom v12.1.

Figure 24. High speed camera bird eyes view.

1.1.6 Lighting system


Back lighting for the camera was provided by Dedocool lighting system which was composed of a spotlight and a motor as shown in Figure 25. The spotlight was placed behind of the rig to allow light to pass through the camera where the

camera was focussing. Dedocool lighting system was able to avoid generating temperature gradients within the rig. The spotlight was clamped with its position adjusted to obtain the best image and videos possible in terms of contrast, colour and brightness. Baking paper was clamped when necessary to provide a translucent environment of the lighting to capture clear images and videos when the frame rate used was at maximum. This was necessary because at maximum frame rate, the exposure to light was very low and so the spotlight is switched on at its brightest level. To balance the brightness, the baking paper was used.

Figure 25. Dedocool lighting system.

1.1.7 Temperature measurement


The temperature measurement was done using a thermocouple designed,

programmed and fabricated by Aime as shown in Figure 26. The temperature detector is a metal which is attached at the side of the rig submerged in the mixture. Calibration of the thermocouple was necessary for accurate temperature measurements. Calibration was performed to obtain an equation shown in Figure 29 which relates the voltage and the temperature by measuring different temperatures (0oC 100oC) of mixtures of the ice cubes and boiling water.

Figure 26. Thermocouple.

Figure 27. LabVIEW temperature measurement.

A voltmeter was used to measure the temperature in terms of voltage. The thermocouple is then inserted into the rig attached to the walls of the rig and

submerged in the liquid mixture. The temperature was read off on the computer screen using LabVIEW as shown in Figure 26. The program was composed by Aime as shown in Figure 27.

Figure 28. Thermocouple simulation.

This equation is to be inserted into LabVIEW. LabVIEW was used as the program to measure the temperature during the experiment.

Figure 29. Temperature calibration equation.

1.1.8 Chemicals
Kerosene was sourced from Sigma Aldrich and used as the dispersed phase in all the experiments. The continuous phase was either pure RO water or RO water with different surfactants glycerol or sodium dodecyl sulphate (SDS) concentrations. Glycerol was purchased as 200mL bottles from Boots while SDS powder was obtained from Sigma Aldrich.

1.1.9 Others materials


25L PVC container Glassware beakers (200mL, 500mL and 5L)

Measuring cylinder Pipette and pipette filler Propanol (cleaning solvent) Methanol (cleaning solvent) Latex gloves

1.1 Experimental procedures


This section will give specific treatment of major experimental procedures so that in future if this work is to be repeated, results can be compared to verify the validity of the results obtained in this current experiment.

1.1.1 Cleaning
The investigation of surface science requires very clean and pure system without any contamination. Any contamination will affect results and disallow comparison between repeated results which leads to waste of time. Hence extreme care was taken when preparing prior to setting up of the experiment. Below is a list cleaning procedures that was performed and is required to be followed for future experiments: Latex sterilised gloves were worn throughout the whole experiment. They are rinsed with methanol and RO water prior to making any contact with the experimental materials. Before the start of each experiment the rig was strongly sprayed and rinsed several times with tap water from the laboratory room. Care was taken to

remove any oily stains or dust or solid particles in the rig by spraying strongly the tap water. The rig is then filled with tap water and left overnight to pull out the contamination particles remained as residues in the rig. The rig core body, cover and the Perspex glass pieces were also washed and sprayed with tap water and dipped into the rig of tap water to allow any residual oily stains to float and rise to the surface.

All connectors, needles and screws were dipped in methanol in a beaker under sonification thermal bath at least twice for 10 minutes before rinsing them with RO water and leaving them to dry in the fume cupboard. The methanol dissolved any residual stains attached to these smaller components and the RO water washes the dissolved particles away.

The capillaries were filled and washed with methanol, RO water and air in that order for several times using new clean syringes. The capillaries were then curled into a small beaker where they were washed with methanol and rinsed with RO water.

Stirrer is to be rubbed and rinsed with methanol, RO water and dry in fume cupboard.

All glass beakers were strongly sprayed with tap water to remove oily stains and then with methanol and tap water and rinse with RO water and left dry in the fume cupboard.

Pipette is washed with RO water several times.

1.1.1 Set up of experiment


The setting up of the experiments was started off with the cleaning procedures mentioned in the previous Section 4.2.1. Once the equipments were cleaned, RO water was brought to mix and saturate with 100mL of kerosene in two 5L glass beakers. This was done so that there is no diffusion within the liquids during experiments. The beakers were left for at 12 hours sealed with cling film paper to prevent entry contamination in the fume cupboard. The rig was to be emptied of the overnight tap water and rinsed with RO water. The rig core was set up with the needles, connectors and tubes. Care was taken when handling the needles without touching the tips. The distance separating the needles was taken to be around 2.8mm, measured using a vernier calliper. It was very

important that the needles are secured tightly with the tips aligned horizontally in line with each other as shown schematically in Figure 30. The needles were made also vertically parallel to each other. The excess kerosene on floating on top of the saturated RO water in the 5L beaker was removed out carefully using a pipette and placed in a beaker. The syringes were then filled with the saturated kerosene. The syringes were also rinsed with the relevant aqueous phase before the experiment. Once the rig core and the connectors are inserted into the rig as shown in Figure 17, the syringe connectors are attached to the syringes located at the infusion pumps. Care was taken not to touch the capillaries inside the cell after the cleaning. The saturated RO water was then fed into the rig slowly without splashing on the rig core while saturated kerosene was delivered to both capillaries by the syringes through pumping. Air within the syringes was discharged by continuous pumping of the saturated kerosene.

Figure 30. Schematic diagram of rig.

1.1.2 Method of data acquisition


Pairs of kerosene drops were formed at the tips of the needles through the capillaries and settled on top of the saturated RO water in the cell. At the low flow rates used the drops formed in each capillary were well separated from subsequent ones and did not interfere with each other and with the coalescence process. It was made sure that minute organic phase was generated so that it does not affect the coalescence process for instance 20% of the rig is covered with kerosene.

The coalescence process in the rig was carried out using two methods:i. Non-induced flow

ii. Induced flow For both methods, the binary coalescence time was recorded using the camera software Phantom Showcase. The kerosene inlet flowrate used throughout the whole experiment was 0.5mL/hr (mean pump flowrate refer to Figures 20 and 21) According to Ban et al. (), the coalescence time is defined as the difference between the time the two drops come into contact and the time the two drops begin to coalesce. Begin to coalesce can also refer to the instant at which the bridge between the two drops is formed. It is important to define illustratively the point at which the two drops come into contact and also the point at which the two drops begin to coalesce. This is because later in the experiments, there was difficulty in judging the point where the two drops start come into contact. The difference between the two methods was that for non-induced flow, the flow of kerosene in the syringe was such that it was stopped at the instant just before the two drops come into contact for the first time. The binary coalescence time for this method was measured at this instant until the formation of the bridge was first witnessed clearly on the camera software depicted on the laptop screen. As for the induced flow method, the flow of kerosene was maintained at 0.5ml/hr until the end of the coalescence process or the drops leave the needles due to unsuccessful coalescence. Hence, if coalescence occurred, the binary coalescence time was calculated from the point at where the drops first contact until the formation of the bridge occurs. For the two methods, glycerol was added at 4 different concentrations v/v of 0.5%, 1.0%, 5.0% and 10.0% to compare the system when glycerol concentration was zero.

For each glycerol concentration, 10-15 pairs of drops were formed. Between the generations of each pair, 5 minutes were given to allow the flow patterns to disperse so that during the coalescence process, the environment was maintained as a stagnant one so that there is no potential shear or disturbance that can affect the coalescence of the pair of droplets. 5mL of kerosene in a syringe was more than sufficient to generate data for all concentration. It was also noted that the change in the liquid level height in the rig was insignificant and did not affect the coalescence process. At the end of the experiment, the rig was emptied and cleaned to repeat the same batch of experiments. This procedure allowed consistency in the system for comparison of results. The coalescence between two drops can be affected by factors such as external vibrations and temperature gradients ()(). Due to the limitations of the experiment, it was unable to avoid any such effects; no thermostatic environment was available and also vibrations were inevitable. Despite the thermocouple was available to monitor the temperature change of the system, if a large deviation from the constant temperature of the system occurs, there was nothing that could be done to remove this occurrence. There were also insufficient benches to place equipments that produce vibrations. However, despite such disturbances, they were attempted to be minimized to as low as possible.

1.1.1 Interfacial tension measurements Pendant drop method


One of the ways to measure the surface tension is by using the pendant drop method which was used in this project. The pendant drop method can be used to determine the static surface and interfacial tensions of liquids. is probably the most convenient, versatile and popular method to measure interfacial tension between emulsion. It involves the determination of the profile of a drop of one liquid suspended in another liquid at mechanical equilibrium. It was important to eliminate any vibration that can disrupt the still.

The measurement of the interfacial tension requires certain dimensions of the kerosene droplet in the medium. It was relatively difficult to measure the dimensions of the oil droplet because there was no close access to it where the droplet can be measured off directly using ruler. Hence, measurements of the oil droplet are to be made through the use of the camera and an image processing software called GNU Image Manipulation Program (GIMP) which is freely distributed over the internet (). The image of the drop in the medium was as shown in Figure 31 was taken by Phantom Camera Control software while GIMP calculates the dimensions necessary for the interfacial tension to be measured by converting pixels to distances in SI units. For each glycerol concentration, this was done several times to account for the error in consistency. Hence the average interfacial tension was used. A droplet was generated each time slowly at 0.5ml/hr and the volume of the droplet at which the droplet is detached from the tip of the needle is recorded. The droplet was regenerated again at about 95% of the recorded volume. The flow was then stopped and the droplet was monitored for at least minute before the image of the oil droplet in the aqueous phase was taken.

Figure 31. Image of the kerosene droplet required for the calculation of the interfacial tension.

1.1.1.1 Equations There are various methods of calculating the boundary tensions from the pendant drop profiles, but the method of Andreas, Hauser and Tucker () is the most

commonly used. With this method, the equatorial diameter de and the diameter ds in a selected plane, which is located by measuring vertically from the vortex a distance equal to de are measured as shown in Figure 6. The ratio of the two diameters ds/de is designated as S, the drop shape; the quantity 1/H is a function of S. The interfacial tension is calculated by the equation (1): =gde2H (1)

where g is the acceleration due to gravity and is the difference in density of the two phases. The relationship between S and 1/H is relatively important to characterize all the profiles of the oil droplet. The range of calculated values of S for 1/H has been extended ()()(). Misak () stated that there are 5 equations that can relate 1/H with S for 5 corresponding ranges of S that are sufficiently accurate to be used in the most exacting interfacial tension calculations : For S = 0.401 to S = 0.46 1H=0.32720S2.56651- 0.97553S2+ 0.84059S-0.18069 For S> 0.46 to S = 0.59 1H=0.31968S2.59725- 0.46898S2+ 0.50059S-0.13261 For S> 0.59 to S=0.68 1H=0.31522S2.62435- 0.11714S2+ 0.15756S - 0.05285 (4) (3) (2)

For S> 0.68 to S=0.90 1H=0.31345S2.64267- 0.09155S2+ 0.14701S-0.05877 For S> 0.90 to S=1.00 1H=0.30715S2.84636- 0.69116S3+ 1.08341S2-0.18341S-0.20970 (6) (5)

The dimensions were calculated and shown in Table 1. The density of glycerol was 800kg/m3.

Table 1. Calculation of kerosene drop dimensions. Figure 32. Dimensions of the kerosene drop needed to be determined for interfacial tension determination.

Conc. v/v glycerol % 0 0.5 1.0 5.0 10.0

de mm 4.52 3 4.51 2 4.50 1 4.51 1 4.46 8

ds mm 2.71 1 2.71 5 2.71 8 2.78 8 2.81 0

S=(ds/ de) 0.5993 81 0.6017 29 0.6038 66 0.6180 45 0.6289 17

H 0.8186 22 0.8270 47 0.8347 61 0.8870 51 0.9284 6

Disper
sed

VROwat
er

VGlyce
rol

Continu
ous

kg/m
3

m3 8000 8000 8000 8000 8000

m3 0.0 40.0 80.8 421. 0 888. 9

kg/m3 998.2 999.5 1000.8 1011.3 1024.4

800 800 800 800 800

kg/ m3 198. 2 199. 5 200. 8 211. 3 224. 4

kg/m2 s2 0.048 59 0.048 18 0.047 81 0.047 55 0.047 33

2 Results and Discussion


2.1 Time evolution of the coalescence process
The time evolution of the coalescence process of two kerosene drops is depicted from Figures 33-52 at the highest frame rate of 6200pps at maximum resolution. From Figure 33, it can be seen that there is a visible thin gap between the two drops and the point of contact is defined in Figure 35. Although in other perception, the dark bridge that occurred at Figure 35 seemed relatively like a shadow, but generally, this point is the start of the binary coalescence time. The bridging of the drops starts at Figure 37 and this represents the point where the binary coalescence time stops. The evolution of this coalescence process corresponded to a binary coalescence time of about one second which happened very rarely throughout the period of the experiment. All other figures show the aftermath of the bridging process which leads to coalescence of the two droplets. The bridge gets larger in size until it becomes the same size as the two droplets. The formation of nipples can be seen on the two opposing sides of the two droplets. They deformed such that the nipples disappear and reappear; the frequency of the sudden expansion and contraction of the coalesced droplets gets smaller with time until it settles on top of the needles as shown as Figure 50.

2.2 Effect of induced flow and glycerol concentration


Figures 53-63 and 64-71 illustrate the cumulative coalescence time distributions and also the individual coalescence time distributions during binary kerosene drop coalescence at the same centre-centre capillary separation distance of 2.8mm for the two methods (i) Induced flow and (ii) Non-induced flow respectively. In a general perspective, the coalescence times for all experiments were not as randomly distributed as shown in the figures. The interfacial tension per unit area was found to decrease with increasing concentration of glycerol as shown in Table 1

and Figure 53 which was also the same result as obtained by Wang et al (). From Figure 53, the trend decreased sharply at lower concentration. At higher concentrations of glycerol, the decrease in interfacial tension per unit area was reduced. Glycerol was used in this experiment to confirm the work of Wang et al () because it was surprising to notice that with glycerol, the interfacial tension decreased but the coalescence time decreased which was a contrast to the current literature. However, this work proved that their findings were nevertheless substantial and I further extended their work without inducing flow to monitor the coalescence trend. From the results obtained, it can be seen that for induced flow, the coalescence time decreases with increasing concentration of glycerol which is in agreement with the work done by Wang et al. (). The only difference between this work and theirs was that they had oil as the continuous phase and deionised water as the dispersed phase while in this work, it was the opposite. Despite this phase inversion issue, the results remained comparable with that of Wang et als because both experiments investigated the same interface between water and oil under the influence of glycerol. However, the result was a contrast for non-induced flow. For non-induced flow, increasing the concentration of glycerol led to stabilization of the drops and is in very good agreement with the vast literature. At 10% concentration v/v glycerol, the non-induced method had 0% coalescence and this was the reason for the missing graphs for the non-induced method for 10% concentration v/v glycerol. But however, comparing the coalescence time between between the two methods at 0% glycerol, the non-induced flow actually promotes coalescence which is in good agreement with the literature review where some authors mentioned that with lower force, coalescence frequency was enhanced. But as the concentration of glycerol is increased, such effect was reversed. One logic reason could be the

inaccuracy of the measurement of the binary coalescence time for the non-induced flow. As it was difficult to judge the first point of contact between the two drops, most of the readings in the data collection could be wrong. This can be possible because it may have been overlooked that they made their first contact but in reality they did not and this could be the reason for the increased stability of the drops. Neglecting this error by assuming that the drops did made their first contact; another error that arises could be the contamination of the liquid system by solid particles that may have entered the liquid system unnoticeably during the experiment. Putting aside the human error aspects for such trends, in scientific terms and together with equipment deficiencies, such trend can be possible, given that glycerol behaves like regular surfactants which happened to have been stabilizing emulsion and drops in liquids. When the pumps were switched off, due to the relaxation of the syringes, the flow might still be induced at very low flowrate. This will keep pushing the drops together, sweeping away glycerol particles away from the centre of the gap between the drops. When this happens, the interfacial tension is highest at this centre and Marangoni effects are induced thus increasing the drainage time. However, this is a defying statement for the induced flow method where the drainage time was decreased. So far, glycerol has not been identified as a general emulsifier or demulsifier in the literature. But however, Griffin () did found that the HLB (hydrophilic-lipophilic balance) value of glycerol is 11.09. This indicated that specifically glycerol is an oil in water emulsifier. Perhaps with this statement, the results obtained from this experiment can be held true coupled with several other explanations such as the flows within the drops. Simply, with glycerol as an oil-in-water emulsifier, for non-induced flow, with increasing concentration of glycerol, this would lead to stabilization. For 0% glycerol concentration, the result obtained from induced flow method is in good agreement with other researchers ()()() as discussed in the literature review. Inducing flow

produced force which presses the drops together and causes the continuous film to drain which depends on the capillary pressure and on van der Waals forces given by p=2R+(AH6h3) () where AH is the Hamaker constant. This was also observed by Wang et al. (). On the aspects of flowrates, the experiments performed only used non-induced flow and induced flow at 0.5ml/hr. It was found that with an increase in the drop diameter and hence radius which was caused by induction of flow, this decreases the overall pressure and lead to larger coalescence times. Borreal and Leal ()() also indicated that as the inlet flow rate increases, the effect of drop size on coalescence time diminishes. It can be concluded that an increase in the inlet flow rate for the same capillary separation distance leads to an increase in the coalescence time. Furthermore, Borreal and Leal () also indicated that with increasing inlet flowrate, Q, the flat film area a between the two drops increases as suggested by the equation 1 for coaxial drop coalescence where ao is the initial value of the flat film area which is equal to 0 for drop-drop coalescence. The increase in area will therefore increase the time for the film to drain to its critical value thus increasing the coalescence time a=tQ2R+ ao2 (1) Another factor that affects the coalescence behaviour in this experiment was the solubility of a component into the opposite phase which allows mass transfer to occur. Nielsen et al () found that coalescence taking place in mutually saturated phases always results in longer coalescence times compared to the unsaturated ones. Charles and Mason () also had similar findings on the same aspect with different liquids. Wang et al () found that coalescence times increased obviously in the saturated mixtures which were in agreement with Nielsen et al. However, in the current experiment, the saturation method was different from Wang et als. Their saturation method included glycerol within the mixture of the oil and aqueous

phases whereas the current experiment had the saturation of water with oil only. Wang et al used glycerol in the saturation process and this halted all possible mass transfer. However, in this experiment, when glycerol is transferring to the oil phase due to unsaturation with glycerol in the first place, then a concentration gradient of the glycerol within the drop can appear. The film region between the pair of drops is easily saturated as the area involved is very small compared to the rest of the continuous water phase. As a result there would be a high glycerol concentration on outer edge of the drops that could give rise to Marangoni flows that will help film drainage (). In the current experiment, such explanation could render validity for the decreased coalescence time when the mixture is not saturated with glycerol and this allowed mass transfer. Wang et al () however, found that this was not the case. This suggests that mass transfer alone cannot fully account for the increased coalescence efficiency in the presence of glycerol in water. The addition of glycerol tends to make the interface to deform easier and the non uniform distribution of surfactants on the interface has been shown to cause fusion of drops because of fluctuations generated at the interfaces (). This was also observed by Dickinson () where thermal fluctuations were known to cause coalescence of drops. Finally, it was noted that the hydrophobic part of glycerol within the oil phase is not long enough to generate steric interactions that would prevent drops from approaching very closely and would delay coalescence ().

Figure 33. Droplets evolution at 0s.

Figure 37. Droplet evolution at 1.731s.

Figure 34. Droplet evolution at 0.256s.

Figure 38. Droplet evolution at 1.732s.

Figure 35. Droplet evolution at 1.101s.

Figure 39. Droplet evolution at 1.733s.

Figure 36. Droplet evolution at 1.726s.

Figure 40. Droplet evolution at 1.734s.

Figure 41. Droplet evolution at 1.735s.

Figure 42. Droplet evolution at 1.736s.

Figure 43. Droplet evolution at 1.740s.

Figure 46. Droplet evolution at 1.743s.

Figure 44. Droplet evolution at 1.741s.

Figure 47. Droplet evolution at 1.753s.

Figure 45. Droplet evolution at 1.742s.

Figure 48. Droplet evolution at 1.754s.

Figure 49. Droplet evolution at 1.757s.

Figure 51. Droplet evolution at 1.797s.

Figure 50. Droplet evolution at 1.758s.

Figure 52. Droplet evolution at 1.820s.

Figure 53. Interfacial tension per area against % concentration v/v glycerol.

Induced method

Figure 55. Induced flow individual percentage against coalescence time graph for 0% glycerol.

Figure 57. Induced flow cumulative percentage against coalescence time graph for1.0% glycerol. 54. 0% Figure 56. Induced flow cumulativepercentage against coalescence time graph for 0.5% glycerol. Figure58. Induced flow individual percentage against coalescence time graph for 0.5%glycerol.

Figure 59. Induced flow individual percentage against coalescence time graph for 1.0% glycerol.

Non-Induced flow method

10
Figure 60. Induced flow cumulative percentage against coalescence time graph for 5.0% glycerol. Figure 61. Induced flow individual percentage against coalescence time graph for 5.0% glycerol. Figure 67. Non-induced flow cumulativepercentage against coalescence time graph for 0.5% glycerol. Figure69. Non-induced flow individual percentage against against glycerol concentration. glycerol. Figure 68. Figure 64.flowcumulativepercentage against coalescence timetimegraph 10.0% glycerol. Figure65. Non-induced flow cumulativepercentage against coalescence time graph for 0.5%glycerol. Figure66. Non-induced flow individual percentage againstcoalescence time graphfor 1.0% glycerol. 62. Induced flow individualmean coalescence time coalescence graph for 10.0% 63. Induced Induced flow percentage against coalescence time graph for for0% 0%

Figure 70. Non-induced flow individual percentage against coalescence time graph for 1.0% glycerol. Figure 71. Non-induced flow cumulative percentage against coalescence time graph for 5.0% glycerol.

11
Figure 72. Non-induced flow individual percentage against coalescence time graph for 5.0% glycerol.

Figure 73. Non-induced flow mean coalescence time against glycerol concentration.

12

3 Conclusion
In this work, the work of Wang et al () was extended such that non-induced flow coalescence method was introduced to compare the results obtained with induced flow. The induced flow method yielded very similar trends with the work of Wang et al. It was found that with increasing concentration of glycerol, the interfacial tension decreased and hence the surface free energy. Decreasing the interfacial tension by increasing the glycerol concentration on both methods showed opposite results. For non-induced method, the binary coalescence time increased and at 10% glycerol concentration v/v, zero coalescence was observed. For induced method, the binary coalescence time decreased favouring coalescence with increasing glycerol concentration. In another perspective, at 0% glycerol

concentration, the non-induced flow resulted in lower coalescence time than that of induced flow. This result was in good agreement with the vast literature such that and introduction of force or shear stress increases the coalescence time enhancing stability. It can be concluded generally that the glycerol acts as a demulsifier for induced flow method while an emulsifier when the flow was not induced. However, the results yielded by the non-induced flow remain sceptical due to human and experimental errors. Due to time limitation, further investigation prohibited from investigating repeatedly on the non-induced flow method. Hence, with this result, the work of Wang et al can remain substantially concrete while further repetitions should done on the non-flow induced method.

13

4 Future work
4.1 Experiment inefficiency and improvements
The design of the rig is a simple one but with various small degrees of inadequate design approaches. The surroundings also contribute to large variations of results obtained from the experiments. This section will detail the necessary improvements needed to be applied onto the rig and procedures if a more accurate reproducible work should be done to compare future results based on the similar basis.

4.1.1 Rig body


4.1.1.1 Thermal glue The rig body is made up of polyethylene plastic which is very inert and does not contribute any effects to the liquid systems. However, the rig is made by attaching the edges of the pieces of polyethylene plastics together using thermal glue. During the fabrication of the rig, the excess thermal glue was removed carefully without leaving any residual attached to the internal parts of the rig body. However, this was not achievable and there remained a small amount of excess dried thermal glue at the joined edges of the plastic pieces which are significantly visible to the naked eye. They appear to be flaky and light. This could be the reason for the appearance of very tiny solid particles suspended in the liquid mixture. As seen

before in the literature review, solid particles do appear to stabilize the emulsion droplets preventing them from coalescing. The thermal glue is shown in Figure. 4.1.1.2 Acetone and methanol contamination Apart from the thermal glue issue, there were visible stains from acetone and methanol contamination which were mistakenly used for cleaning the rig previously. However, it was not advisable as acetone left several white translucent patches of stains in the rig. It might have damaged the rig as the white patches were not

14

removable. Methanol is not advisable because of the chemical reactions with the polyethylene plastic pieces. These unknown stains may not be inert and could have potentially affected the coalescence experiment in many ways. 4.1.1.3 Protrusion cork The protrusion which is made of polyethylene plastic to partly immobilise the rig core is inserted and attached securely through a hole drilled at the centre bottom of the rig body using thermal glue. The protrusion has a hollow space running vertical through its centre. Hence the protrusion can be viewed as a solid annular cylinder attached to the bottom of the rig body at its bottom. Obviously the hollow space will drain the liquid mixture away and hence the cork was inserted at the bottom of the protrusion to cease any flow of liquid out of the rig. An inert rubber cork is shaped and inserted very tightly at the bottom of the hollow space in the protrusion. Despite this, it may have accumulated very minute amount of kerosene or glycerol. It was found that overnight kerosene will lead to higher stability of the kerosene droplets. The leftover glycerol or kerosene may deposit and being trapped between the tiny spaces between the cork and internal sides of the cylindrical protrusion. They may have altered in their physical properties overnight. Moreover, such trapped kerosene liquid is extremely difficult to remove once inserted tightly and securely into the bottom of the hollow space. During washing the hollow space with the inserted cork were strongly sprayed with RO water and the cork was not removed for a more considerable clean. The issue of thermal glue is also similar to that of Section 5.1.1.1. An improvement for this is a solid cylinder instead of the annular one. The solid cylinder will cease all potential leakage. However, the attachment of the solid cylinder requires thermal glue.

15

4.1.1.4 Slanting of rig core This is a visual problem for the camera to focus. The rig core is a cubical solid made out of polyethylene plastic with a vertical cylindrical hollow space at the centre for insertion of the protrusion. When fitted through the protrusion, the cube should lie flat horizontally on the bottom of the rig body. However this was not the case. It appeared that the rig core is tilted an angle less than 10o on the right. This problem was thought to be the cause of the unequal size fitting between the cylindrical hollow space of the rig core and the protrusion. More specifically, the diameters of the protrusion and the cylindrical hollow space of the rig core at near bottom are not similar. Obviously for the rig core to fit through the protrusion, the diameter of the protrusion should be less than 1% larger than the cylindrical hollow space of the rig core. However, it appeared that near bottom, the right side of the protrusion at near bottom was not fabricated at the diameter as top part; like an appearance of very small outgrowth that is invisible to the eye as shown in Figure.

The problem that this discrepancy in design can cause is the inaccuracy of image processing of the droplets to obtain the correct interfacial tension. Due to time limitation, there were no amendments made neither the protrusion nor the cylindrical hollow space of the rig core. Instead, when taking images or recording videos of the coalescence process, the camera position was adjusted to match the inclination of the rig core, and hence the inclination of the needles, so that the tip needles are correctly aligned horizontally and the height vertically. However, in the videos and images, the droplets might look a bit less than 90o vertically due to gravity and buoyancy effects.

16

Improvement is to eradicate the unequal shape of the protrusion with respect to the cylindrical hollow space of the rig core. 4.1.1.5 Rig cover and entry of experimental objects The purpose of the rig cover is to prevent the entry of unwanted solid particles during the experiment. However, throughout the whole experiment, the cover was left opened as shown in Figure. The reason for this is to allow the capillaries or tube and the metal temperature detector to pass into the rig body. However, the gap left by the position of the rig cover may introduce dust particles into the rig and contaminate the liquid mixture despite the fact that these dust particles may be light and only affects the top section of the liquid mixture. Larger solid or liquid particles may gain entry by accident such as coughing and sneezing. An improvement is to close the exposed top of the rig body with the rig cover completely. Three small holes can be drilled on top of the cover for the passage of the capillaries and metal temperature detector into the system. The size of the holes should be small enough to just fit these objects to reduce the chance of particles entering and contaminating the system. Sealing these holes can be difficult because these objects are washed often. 4.1.1.6 Mobility of capillaries and thermocouple The capillaries and thermocouple were both partly submerged in the liquid mixture. During experiment it is essential that these objects remain stagnant to go accordant with the basis of the experiment no external force to disrupt the coalescence process. From Figure, it can be seen that the capillaries and thermocouple were stable clinging onto the top edges of the rig body. However, from experience, these objects moved significantly due to relaxation, in this case the capillaries. The thermocouple is attached to a BLUTACK on the top wall of the rig body and due to the inelasticity of the BLUTACK, the thermocouple fell and

17

moved several times in the run. The BLUTACK contaminated the system nevertheless and the experiment has to be redone. To avoid such experimental slip, it is recommended that a special locking system with screws and locks that can be custom built on the rig body to secure these objects to cease any chance of mobility that can disturb the system. 4.1.1.7 Plates The plates were cut to prevent further mobilisation of the rig core. However, more precision and accuracy of the dimension were needed to immobilise the rig core in a correct form and not by tilting as shown in Figure. 4.1.1.8 Mixing Surfactants are added in by removing the rig body cover and slowly pouring the surfactants while stirring. It is not advisable to pour at a large rate; very small rate is needed to sufficiently mix well. The stirrer which is a metal spatula is washed and rubbed with methanol to remove stains and rinse with RO water. Removing the rig body cover introduces particles into the system.

4.1.2 Rig core


4.1.2.1 Placement of needles The rig core comes with two parts the rig core body and the rig core head as shown in Figure. On the top sides of the cubical rig core body, V-shaped or triangular grooves were cut off from the rig core body. The function of the grooves is to allow the cap needles to sit on them while the rig core head presses on them and screwed to fix and hold the positions of the needles securely. The grooves were cut accurately to say the least as shown in Figure. However, the problem arises when adjusting the needles.

18

The pair of needles was initially straight as shown in Figure and they were bent using hands on a strongly held vertical metal pole. Such method was very crude but due to limitations of the experimental sources, it was the only possible way to get as much similarly bent as possible. Due to limited amount of such needles, only the best few were chosen for the experiment. As shown in Figure, needles were bent differently at least. The diameters of the needles were checked using a vernier calliper instrument. Because the needles are not generally similar in terms of the bent areas, they were difficult to place in line with each other. The needles should be vertical parallel to each other and their tips should be horizontally in line with each other. Nevertheless such difficult task was always completed through different levels of screw tightness. Despite this, to the naked eye this may seemed to be aligned but if they were to be accurate measured their relative positions, they might be not. Such accurate position was to allow accurate determination of the coalescence time of the two droplets. In front of the camera view, apart from the needles being vertically parallel and the tips horizontally aligned, another criterion for accurate measurement of the coalescence behaviour is that when the droplets come in contact, the droplets should be seen pressed against each other side by side without one side behind of the other. Their circumference at the point of touching should be like Figure and not in Figure.

19

There were experimental sessions where the droplets were generated as shown in Figure which is not desirable. A major improvement for this is to replace or replicate a similar needle of the same shape and size. Another factor is the reproducible positions for every repetition of experiment. Because every component has to be dissembled for thorough cleaning, upon reassembly of the components, the general position which reflects the angle of inclination, distance between the two needles and height will be different at least to the one tenth of a centimetre. 1.1.1.1 Connectors The connectors from the rig core come from the needles. As shown in Figure, the connectors were not as tight as it seemed. Because the experiment requires periodical stirring when adding surfactants, to mix well, care has to be taken to avoid contacting the connectors. The connector to the needle is not of screw type; slide, slot, twist and tighten type as shown in Figure and hence they are very susceptible to the slightest disturbance introduce such as the spatula slightly contacted it and displaced its initial position which can cause potential leakage of the oil phase when loosen. It is rather difficult to tighten the loosen connector without introducing contaminants into the system. When such event occurred, the whole experiment is to be redone. The solution to such issue is to use a screw type connector which can be very secure. 1.1.1.2 Overall design of rig core The rig core design was simplified one but however, the fact that it has a many components linked to it made it more exposed to accidental errors. Perhaps to repeat the experiment using the same methodology, the design of the rig core has

20

to be significantly improved. For instance, instead of locking the positions of the needles by pressing and screwing the rig core head, a rig core can be designed such that the needles can be inserted through a fix hole within the rig core without any difficulty. 1.1.1.3 Capillaries The capillaries used are made of polyethylene and they are fragile and susceptible to sharp bents. Hence pressuring the capillaries should be avoided. One advantage of this tube is its inert material and its internal section is visible enough to detect undesirable solid particles from entering the system. The capillaries were more than 70 cm which is unnecessarily too lengthy and difficulty arises during cleaning because even after properly cleaned, the skin of the capillaries are easily contaminated upon contacting surfaces like the bench or laboratory coat. Hence more caution has to be taken when handling the capillaries because part of them was submerged into the liquid mixture. Any contamination from other particles such as sand or oil particles can ruin the consistency of the system.

1.1.2 Pumps
Due to lack of pump resources, two different infusion pumps were used instead of the same ones. From the pump calibration figures, it can be seen that at the same rate, the two different pumps generated two different pump rates. Calibration was necessary to find the true volumetric flowrate due to inefficiencies of the instruments. It is suggested that two new similar pumps to be purchased and the calibration from the two pumps should give equivalent true volumetric flowrate at a given setting. Often the pumps give obstacles for a clear coalescence process. This happened because upon the release of the coalesced droplets or unwanted droplet from each

21

needle due to unsuccessful coalescence, the pumps were simultaneously stopped to prevent the growth of droplet, but this did not happen. Even when the pumps had stopped, the growth of the droplets from the two needles remained continuous despite the growth rate was slow enough to allow the flow patterns to disperse away under the basis of time 5 minutes before making first contact. However, in the middle of the allowance time, say 2.5 minutes, the size of the droplets could be large enough to cause force of attraction and repulsion between them. However, despite the efforts in making changes to the pumps, such occurrence was inevitable throughout the whole experiment. 1.1.2.1 Controlling the growth rate The volumetric flowrate of the two pumps were set to the true volumetric flowrate of 0.5ml/hr and this corresponds to 0.63ml/hr and 0.68ml/hr for red pump and white pump respectively look at Figures. When generating the droplets, the size or specifically the volume of the droplets has to be equal for consistent measurement. However, difficulty arises when using the experimental pumps. Efforts were made to generate the droplets volume before contacting each other. If the one droplet is seen to be outgrowing the other, the flowrate generating the larger droplet is stopped and then switched on when the smaller droplet comes to the same size as the larger droplet. The method used in this is very inaccurate as it seemed to be. The droplets size were always monitored regularly throughout the experiment by observing the growth rate and the size of the droplets by placing a ruler horizontally on the computer screen which shows the live magnification of the droplets in the rig, and then move the horizontal ruler vertically and stop when the tip of one droplet is found. If the tip of the droplet is not aligned horizontally with this tip, then the flowrates are adjusted such that the tips are aligned horizontally as shown in Figure. Again this method show very little accuracy in determining the equality of both volumes. The volumetric flowrates must be equal to one another.

22

Fix volumes of droplets can be generated but such function is only available in the red pump but not on the white one. Solution is to get the same pumps preferably new. Overall the pumps should be control by a controller such that one switch will trigger the motors of the pumps and will induce equivalent flow at low flowrates. Previously a three way valve was used but had been a failure due to the unequal flowrates from two sides as shown in Figure.

1.1.3 Thermocouple
The fabrication of the thermocouple is complex and was completed successfully by Aime. The components of the thermocouple were of detectors as shown in Figure which are programmed to function as temperature measurer. Calibration was necessary to obtain correct temperature measurements using the instrument. However, throughout the experiment, the sensitivity of the temperature was very high as shown in Figure. The x- axis can be taken as per seconds while the y-axis is the temperature. The range of temperature variation was large from 19 oC- 28oC. And it was impossible for temperature to rise in an instant given that the room temperature did not vary as such. The sensitivity of the thermocouple was too large.

1.1.4 Lens
A more desirable lens is such that it can focus clearly on the interface with higher resolution.

23

1.2 Environment
1.2.1 Vibrations
The vibrations were caused by 4 objects, the 2 pumps, cold light and fume cupboard. The vibration from the pump is caused by the clicking of the motor that turns the screws which pushes the syringe to induce flow. The motor of the cold light also produces vibration. The fan in the fume cupboard produces the vibration. Due to the lack of benches, the pumps and cold light were put on the same bench as the rig. The fume cupboard is turned on continuously. These vibrations will nevertheless affect the surroundings of the droplets. An improvement for this would be to place the pumps and light on different and separate benches. The fan from the fume cupboard should be switched off when not using.

1.2.2 Temperature
The coalescence process is influenced by the change in temperature. Slight temperature change can change the interfacial tension and cause inconsistency with the coalescence behaviour. The only source of temperature gradient is the cold light. A transparent plastic glass should be placed in front of it to absorb the heat radiated from the light.

1.3 Data collection


To collect the binary coalescence time, for the first experiment where only touching was allowed and no flow induced. Such method was tedious because it is very difficult to define the point at which the surfaces touch and then wait for the coalescence to occur without inducing flow as shown in Figure. This method of data collection is very inaccurate. Moreover, in Section, the pumps had relaxation and

24

often generated slower growth rates even after the pumps were switched off. In Figure, the dark shadow which forms at this point is taken as the point of contact. When this occurs to the naked eye, the stopwatch begins to run until the bridging appears. Due to the limitation of the camera software, the camera was unable to record very high frame rates for a larger period of an event. For instance, at frame rate 6200pps, it can only record 3.589 seconds with maximum resolution. Maximum resolution and frame rate are desirable to obtain the point where the contact and bridging occur. However, throughout the experiment, the videos and images were recorded using 1000pps which allowed 22 seconds of recording. However for the second part of the experiment, this was simpler because the flow is induced. When the shadow starts to appear, the recording is started and the stopwatch starts until the droplets coalesced. It was not necessary for the exact time at which the shadow appears because the difference between the exact time at such event occurs and the point at which the observer notice its occurrence and simultaneously starts the stopwatch should be a fraction of a second. This issue is the same as the bridging and the full coalescence of the droplets. We humans can see Figure as a whole and stopped the time when this occurs but not Figure (bridging). The time difference between the two should also be a fraction of a second. Hence the accuracy in such cases is not necessary with respect to the coalescence time recorded. It is important that only the middle side of the surface touch of the droplets touch and not other parts.

1.4 Image processing


The quality of the image and videos are very important especially with regards to the formation of the bridge

25

1.5 Batch saturated


When carrying out the experiment for one day, it is recommended that it is important to use just one syringe batch enough for the experiment. 5mL was sufficiently enough to produce up to 15 droplets for 0, 0.5, 1.0, 5.0 and 10.0 %.

1.6 Overnight
It is recommended to carry out one whole batch of experiments before proceeding to repeat the second batch to perform reproducible data or result. The saturated RO water is prepared 12 hours before the experiment starts.

26

2 Bibliography
Aarts, D.G.A.L. & Lekkerkerker, H.N.W., 2008. Droplet coalescence: drainage, film rupture and neck growth in ultralow interfacial tension systems. Journal of Fluid Mechanics, 606., pp.275-94. Abdurahman, H.N., Abu Hassan, M.A. & Yunus, R.M., 2007. Characterization and demulsification of water-in-crude oil emulsions. Journal of Applied Sciences, 7(10), pp.1437-41. Abdurahman, H.N. & Yunus, R.M., 2009. Coalescence of Water Droplets in Waterin-Crude Oil Emulsions. International Journal of Chemical Technology, 1, pp.19-25. Abdurahman, H.N., Yunus, R.M. & Jemaat, Z., 2007. Chemical demulsification of water-in-crude oil emulsions. Journal of Applied Sciences, 7(2), pp.196-201. Abid, S. & Chesters, A.K., 1994. The drainage and rupture of partially-mobile films between colliding drops at constant approach velocity. International Journal of Multiphase Flow, 20, pp.613-29. Ali, M.F. & Alqam, M.H., 2000. The Role of Asphaltene, Resins and Other Solids in the Stabilization of Water in Oil Emulsions and its Effects on Oil Production in Saudi Oil Fields. Fuel, pp.1309-16. Allan, C.G.E. & Mason, S.G., 1961. The approach of gas bubbles to a gas/liquid interface. Journal of Colloid Science, 16(2), pp.150-65. Al-Mulla, A. & Gupta, R.K., 2000. Droplet coalescence in the shear flow of model emulsions. Rheol Acta, 39, pp.20-25. Andreas, J.M., Hauser, E.A. & Tucker, W.B., 1938. Boundary Tension by Pendant Drops. The Journal of Physical Chemistry, pp.1001-19. Anon., 1969. In Manual on Disposal of Refining Wastes. Washington D. C.: American Petroleum Institute. pp.5-15. Auflem, I.H., 2002. Influence of Asphaltene Aggregation and Pressure on Crude Oil Emulsion Stability. PhD Thesis. Trondheim, Norway: Norwegian University of Science and Technology. Aversana, P.D., Monti, R. & Gaeta, F.S., 1995. Marangoni flows and coalescence phenomena in microgravity. Pergamon, 16(7), pp.95-98.

27

Baldessari, F., Homsy, G.M. & Leal, L.G., 2007. Linear stability of a draining film squeezed between two appraoching droplets. Journal of Colloid and Interface Science, 307, pp.188-202. Bancroft, W.D., 1913. Theory of Emulsification. Journal of Physical Chemistry, pp.501-19. Ban, T., Kawaisumi, F., Nii, S. & Takahashi, K., 2000. Study of drop coalescence behaviour for liquid-liquid extraction operation. Chemical Engineering Science, 55, pp.5385-91. Bazhlekov, I.B., Chesters, A.K. & van de Vosse, F.N., 2000. The effect of the dispersed to continuous-phase viscosity ratio on film drainage between interacting drops. International Journal of Multiphase Flow, 26, pp.445-66. Becher, 1985. Encyclopedia of Emulsion Technology. New York: Marcel Dekker. Bfigura, 2007. Wikipedia. [Online] Available at:

http://en.wikipedia.org/wiki/Ostwald_ripening [Accessed 4 April 2010]. Bhardwaj, A. & Hartland, S., 1994. Dynamics of emulsification and demulsification of water in crude oil emulsions. Industrial Engineering Chemistry Research, 33, pp.1271-79. Bhardwaj, A. & Hartland, S., 1994. Kinetics of coalescence of water droplets in water-in-crude oil emulsions. Journal of Dispersion Science and Technology, 15, pp.133-46. Binks, B.P., 2002. Particles as surfactants - similarities and differences. Current Opinion in Colloids and Interface Science, 7(1), pp.21-41. Borges, B., Rondon, M., Sereno, O. & Asuaje, J., 2009. Breaking of water-in-crudeoil emulsions. 3. Influence of salinity and water-oil ratio on demulsifier action. Energy & Fuels, 23, pp.1568-74. Borrell, M. & Leal, L.G., 2008. Viscous coalescence of expanding low-viscosity drops; the dueling drops experiment. Journal of Colloid and Interface Science, 319, pp.263-69. Borrell, M. & Leal, L.G., 2008. Viscous coalescence of expanding low-viscosity drops; the dueling drops experiment. Journal of Colloid and Interface Science, 319, pp.263-69.

28

Bremond, N., Thiam, a.R. & Bibette, J., 2008. Decompressing Emulsion Drolets Favors Coalescence. Physical Review Letters, 100, p.024501. Burrill, K.A. & Woods, D.R., 1973. Film shapes for deformable drops at liquid-liquid interface III drop rest times. Journal of Colloid and Interface Science, 42, pp.35-51. Charles, G.E. & Mason, S.G., 1960. The coalescence of liquid drops with flat liquid/liquid interfaces. Journal of Colloid and Interface Science, 15, pp.236-67. Charles, G.E. & Mason, S.G., 1960. The mechanism of partial coalescence of liquid drops at liquid-liquid interfaces. Journal of Colloid and Interface Science, 15, pp.105-22. Chen, J.-D., 1985. A model of coalescence between two equal-sized spherical drops or bubbles. Journal of Colloid and Interface Science, 107(1), pp.209-20. Chen, C.T., Maa, J.R., Yang, Y.M. & Chang, C.H., 1998. Effects of electrolytes and polarity of organic liquids on the coalescence of droplets at aqueous-organic interfaces. Surface Science, 406, pp.167-77. Chen, D. & Pu, B., 2001. Studies on the binary coalescence model. II. Effects of drops size and interfacial tension on binary coalescence time. Journal of Colloid and Interface Science, 243, pp.433-43. Chen, D. & Pu, B., 2001. Studies on the binary coalescence model. II. Effects of drops size and interfacial tension on binary coalescence time. Journal of Colloid and Interface Science, 243, pp.433-43. Chen, G. & Tao, D., 2005. An experimental study of stability of oil-water emulsion. Fuel Processing Technology, 86, pp.499-508. Chesters, A.K., 1991. The modeling of coalescence processes in fluid-liquid dispersions: A review of current understanding. Chemical Engineering Research and Design, 69, pp.259-70. Chesters, A.K., 1991. The Modelling of Coalescence Processes in Fluid-Fluid Dispersions: A Review of Current Understanding. Chemical Engineering Research and Design, pp.259-70. Chesters, A.K. & Bazhlekov, I.B., 2000. Effect of insoluble surfactants on drainage and rupture of a film betwen drops interacting under a constant force. Journal of Colloid and Interface Science, 230, pp.229-43.

29

Chevaillier, J.P., Klaseboer, E., Masbernat, O. & Gourdon, C., 2006. Effect of mass transfer on the film drainage between colliding drops. Journal of Colloid and Interface Science, 299, pp.472-85. Cristini, V., Blawzdziewiez, J. & Loewenberg, M., 1998. Near contact motion of surfactant covered spherical drops. Journal of Fluid Mechanics, 366, pp.259-87. Crossdale, K.R. et al., 1999. Current and future hydrocarbons research and development. Technical report. K. R. Crossdale & Associates Ltd. Czarnecki, J. & Moran, K., 2005. On the Stabilization Mechanism of Water-in-Oil Emulsions in Petroleum Systems. Energy and Fuels, 19, pp.2074-79. Czarnecki, J. & Moran, K., 2005. On the stabilization mechanism of water-in-oil emulsions in petroleum systems. Energy & Fuels, 19, pp.2074-79. Dai, B. & Leal, L.G., 2008. The Mechanism of Surfactant Effects on Drop Coalescence. Physics of fluids, 20, p.040802. Danov, K.D. et al., 1993. Coalescence dynamics of deformable brownian emulsion droplets. Langmuir, 9, pp.1731-40. Davies, G.A., 1992. Mixing and coalescence phenomena in liquid-liquid systems. In J.D. Thornton, ed. Science and Practice of Liquid-Liquid Extraction. Oxford: Clarendon Press. Davis, R.H., Schonberg, J.A. & Rallison, J.M., 1989. The lubrication force between two viscous drop. Physics of Fluids A1, pp.77-81. Deng, S. et al., 2005. Destabilization of oil droplets in produced water from ASP flooding. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 252(23), pp.113-19. Dickinson, E., 1992. Interfacial interactions and the stability of oil in water emulsions. Pure & Applied Chemistry, 64(11), pp.1721-24. Dreher, T.M., Glass, J., O'Connor, A.J. & Stevens, G.W., 1999. Effect of rheology on coalescence rate and emulsion stability. AlChE Journal, 45, pp.1182-90. Edwards, D.A., Brenner, H. & Wasan, D.T., 1991. Interfacial Transport Process and Rheology. Boston: Butterworth-Hienmann.

30

EIA,

2010.

US

Energy

Information

Administration.

[Online]

Available

at:

http://www.eia.doe.gov/conference/2010/session3/eizember.pdf [Accessed Monday April 2010]. Evans, D.F. & Wennerstrom, H., 1999. The Colloidal Domain: Where Physics, Chemistry, Biology and Technology Meet. New York: Wiley-VCH. Fingas, M.F. & Fieldhouse, B., 1994. Studies of water-in-oil emulsions and techniques to measure emulsion treating agents. In 17th Arctic and Marine Oil Spill Program (AMOP), Technical Seminar. Ottawa, Ontario, 1994. Fingas, M.F. et al., 1994. Laboratory effectiveness testing of water-in-oil emulsion breakers. In P. Lane, ed. The Use of Chemicals in Oil Spill Response. Philadelphia: ASTM STP 1252. Fordham, S., 1948. On the calculation of Surface Tension from Measurements of Pendant Drops. In Proceedings of the Royal Society. Mathematical, Physical and Engineering Sciences. London, 1948. Royal Society Publishing. Gaaseidnes, K. & Turbeville, 1999. Separation of Oil and Water in Oil Spill Recovery Operations. Journal of Pure Applied Chemistry, pp.95-101. Ghosh, P., 2004. Coalescence of air bubbles at air-water interface. Chemical Engineering Research and Design, 82, pp.849-54. GIMP, 2001-2010. GIMP. [Online] Available at: http://www.gimp.org/ [Accessed 17 September 2010]. Giribabu, K. & Ghosh, P., 2007. Adsorption of nonionic surfactants at fluid-fluid interfaces: Importance in the coalescence of bubbles and drops. Chemical Engineering Science, 62, pp.3057-67. Gnyloskurenko, S.V., Byakova, A.V., Raychenko, O.I. & Nakamura, T., 2003. Influence of wetting conditions on bubble formation at orifice in an inviscid liquid. Transformation of bubble shape and size. Colloids and Surface A: Physicochemical and Engineering Aspects, 218, pp.73-87. Goodall, D.G., Gee, M.L. & Stevens, G.W., 2002. An imaging reflectometry study of the effect of electrolyte on the drainage and profile of an aqueous film between an oil droplet and a hydrophilic silica surface. Langmuir, 18, pp.4729-2735.

31

Greene, M.R., Hammer, P.A. & Olbricht, W.L., 1994. The Effect of Hydrodynamic Flow Field on Colloidal Stability. Journal of Colloid and Interface Science, pp.23246. Griffin, W.C., 1954. Calculation of HLB values of non-ionic surfactants. Journal of Society of Cosmet Chemistry, pp.249-56. Groeneweg, F., van Voorst Vader, F. & Agterof, W.G.M., 1993. The dynamics stability of w/o emulsions prepared from vegetable oil. Chemical Engineering Science, 48(2), pp.229-38. Guido, S. & Simeone, M., 1998. Binary Collision of Drops in Simple Shear Flow by Computer-Assisted Video optical Microscopy. Journal of Fluid Mechanics, pp.1-20. Hartland, S. & Wood, S.M., 1973. Effect of applied force on the approach of a drop to a fluid-liquid interface. AlChE Journal, 19, pp.871-76. Hodgson, T.D. & Lee, J.C., 1969. The effect of surfatants on the coalescence of a drop at an interface-I. Journal of Colloid and Interface Science, 30, pp.94-108. Hodgson, T.D. & Woods, D.R., 1969. The effect of surfactants on the coalescence of a drop at an interface. II. Journal of Colloid and Interface Science, 30(4), pp.42946. Hodgson, T.D. & Woods, D.R., 1969. The effect of surfactants on the coalescence of a drop at an interface-II. Journal of Colloid and Interface Science, 30, pp.429-46. Hung, J., Ortega, D., Castillo, J. & Acevedo, S., 2007. Studies of kinetics of coalescence of water-in-furrial crude oil emulsions by confocal microscopy. Journal of Dispersion Science and Technology, 28, pp.717-23. Hu, Y.T., Pine, D.J. & Leal, L.G., 2002. Drop Deformation, Breakup, and Coalescence with Compatibilizer. Physics of Fluids, pp.484-89. Ivanoc, I.B., Danov, K.D. & Kralchevsky, P.A., 1999. Flocculation and coalescence of micron-size emulsion droplets. Colloids and Surfaces !: Physicochemical Engineering Aspects, 152, pp.161-82. Ivanov, I.B. & Dimitrov, D.S., 1988. Thin Liquid Films: Fundamentals and Applications. New York: Marcel Dekker. p.379.

32

Ivanov, I.B., Dimitrov, D.S., Somasundaran, P. & Jain, R.K., 1985. Thinning of films with deformable surfaces: Diffusion-controlled surfactant transfer. Chemical Engineering Science, 40(1), pp.137-50. Jeelani, S.A.K. & Hartland, S., 1993. Effect of velocity fields on binary and interfacial coalescence. Journal of Colloid and Interface Science, 156, pp.467-77. Kabalnov, A.S., Pertzov, A.V. & Shchukin, E.D., 1987. Ostwald ripening in emulsions: I. Direct observations of Ostwald ripening in emulsions. Journal of Colloid and Interface Science, 118(2), pp.590-97. Kang, W.L. et al., 2004. The action mechanism of demulsifiers at model O/W interfacial film. Acta Physico - Chimica Sinica, 20(2), pp.194-98. Kilpatrick, P.K. & Spiecker, P.M., 2001. Chapter 30. In J. Sjoblom, ed. Encyclopedic Handbook of Emulsion Technology. New York: Marcel Dekker. pp.707-30. Kim, J. & Longmire, E.K., 2009. Investigation of binary drop rebound and coalescence in liquids using dual-field PIV technique. Experiments in Fluids, 47, pp.263-78. Koh, A., Gillies, G., Gore, J. & Saunders, B.R., 2000. Flocculation and coalescence in oil-in-water poly(dimethylsiloxane). Journal of Colloid and Interface Science, 227, pp.390-97. Kumar, K., Nikolov, A.D. & Wasan, D.T., 2001. Mechanisms of stabilization of water-in-crude oil emulsions. Industrial Engineering Chemistry Research, 40, pp.3009-14. Kumar, K., Nikolov, D. & Wasan, D.T., 2002. Effect of film curvature on drainage of thin liquid films. Journal of Colloid and Interface Science, 256, pp.194-200. Leal, L.G., 2004. Flow Induced coalescence of Drops in a Viscous Fluid. Physics of Fluids, 16(6), pp.1833-51. Lee, J.C. & Hodgson, T.D., 1968. Film flow and coalescence. I. Basic relations, film shape and criteria for interface mobility. Chemical Engineering Science, 23, pp.1375-97. Leunissen, M.E. et al., 2007. Electrostatics at the oil-water interface: Stability and order in emulsions and colloids. Proceedings of the National Academy Sciences of the USA, 104(8), pp.2585-90.

33

Li, D., 1994. Coalescence between two small bubbles or drops. Journal of Colloid and Interface Science, 163, pp.108-19. Li, J. & Gu, Y., 2005. Coalescence of oil-in-water emulsions in fibrous and granular beds. Separation and Purification Technology, 42, pp.1-13. Likhterova, N.M., Kovalenko, V.P. & Lebedev, V.V., 2003. Khim. Tekhnol. Topl. Masel. pp.24-28. Li, D.M. & Slattery, J.C., 1988. Experimental support for analyses of coalescence. AlChE Journal, 34, pp.862-64. Lobo, L., 2002. Coalescence during emulsification. Journal of Colloid and Interface Science, 254, pp.165-74. Loewenberg, M. & Hinch, E.J., 1997. Collision of Two Deformable Drops in Shear Flow. Journal of Fluid Mechanics, pp.299-315. Mackay, G.D.M. & Mason, S.G., 1961. The Marangoni effect and liquid/liquid coalescence. Nature, 191(488). Mackay, G.D.M. & Mason, S.G., 1963. Some effects of interfacial diffusion on the gravity coalescence of liquid drops. Journal of Colloid and Interface Science, 18, pp.674-83. Mackay, G.D.M. & Mason, S.G., 1963. The gravity approach and coalescence of fluid drops at interfaces. Canadian Journal of Chemical Engineering, 41, pp.203-12. Makhonin, G.N., Petrov, A.A. & Borisov, S.I., 1979. Surface-Active Components of Petroleum Emulsion Stabilizers. Chemistry and Technology of Fuels and Oils, pp.3841. Mana-Zloczower, I., 1994. Mixing and Compounding of Polymers - Theory and Practice. Munich: Hanser Publishers. Marangoni, C.G.M., 1865. Sull expansiome dell goccie di un liquido galleggianti sulla superficie di altro liquido. PhD Thesis. Pavia. Marrucci, G., 1969. A theory of coalescence. Chemical Engineering Science, pp.97585.

34

McLean, J., SPiecker, P.M., Sullivan, A. & Kilpatrick, P.K., 1998. The Role of Petroleum Asphaltenes in the Stabalization of Water-in-Oil Emulsions. In In Structure and Dynamics of Asphaltenes. New York: Plenum Press. McNaught, A.D. & Wilkinson, A., 1997. IUPAC Compendium of Chemical

Terminology. 2nd ed. Mikula, R.J. & Munoz, V.A., 2000. Characterization of demulsifiers. In L.L. Schramm, ed. Surfactants: Fundamentals and applications in the petroleum industry. 1st ed. Cambridge: Cambridge University Press. pp.51-79. Miller, D.J. & Bohm, R., 1993. Optical studies of coalescence in crude oil emulsions. Journal of Petroleum Science and Engineering, 9, pp.1-8. Mills, O.S., 1953. British Journal of Applied Physics, p.247. Misak, M.D., 1968. Equations for Determining 1/H versus S Values in Computer Calculations of Interfacial Tension by the Pendent Drop Method. Journal of colloid and Interface Science, pp.141-42. Mitra, T. & Ghosh, P., 2007. Binary coalescence of water drops in organic media in presence of ionic surfactants and salts. Journal of Dispersion Science and Technology, 28, pp.785-92. Mukerjee, P. & Mysels, K.J., 1971. Critical micelle concentrations of aqueous surfactant systems. In NSRDS-NBS 36. Washington DC: US Government Printing Office. Mullin, J., 2006. Bureau of Ocean Energy Management, Available [Accessed 4 Regulation and at: April

Enforcement. 2010].

[Online]

http://www.boemre.gov/tarprojectcategories/behavior.htm

Nandi, A., Khakhar, D.V. & Mehra, A., 2001. Coalescence in surfactant stabilized emulsions subjected to shear flow. Langmuir, 17, pp.2647-55. Nandi, A., Mehra, A. & Khakkar, D.V., 2005. Coalescence in a surfactant-less emulsion under simple shear flow. AlChE Journal, 52(3), pp.885-94. Nemer, M.B. et al., 2004. Hindered and Enhanced Coalescence of Drops in Stokes Flow. Physical Review Letters, p.114501.

35

Nielsen, L.E., Wall, R. & Adams, G., 1958. Coalescence of liquid drops at oil-water interfaces. Journal of Colloid and Interface Science, 13, pp.441-58. Paunov, V.N., 2008. University of Hull. [Online] Available at:

www.hull.ac.uk/scg/paunov/paunov06536-9.pdf [Accessed 4 April 2010]. Petroleum, B., 2007. [Online] Available at:

http://www.bp.com/liveassets/bp_internet/globalbp/globalbp_uk_english/reports_a nd_publications/statistical_energy_review_2008/STAGING/local_assets/downloads/ pdf/statistical_review_of_world_energy_full_review_2008.pdf [Accessed 2010]. Pharma, April 2010]. Pichot, R., Spyropoulus, F. & Norton, I.T., 2010. O/W emulsions stabilised by both low molecular weight surfactants and colloidal particles: The effect of surfactant type and concentration. Journal of Colloid and Interface Science, p.Accepted Manuscript. Poznyshev, G.N., 1982. Stabilization and Breaking of Emulsions [in Russian]. Pu, B. & Chen, D., 2001. Studies on the binary coalescence model. I. Jumping coalescence phenomenon. Journal of Colloid and Interface Science, 235, pp.1-3. Rondon, M., Bouriat, P. & Lachaise, J., 2006. Breaking of water-in-crude oil emulsions. 1. Physicochemical phenomenology of demulsifier actio. Energy & Fuels, 20, pp.1600-04. Rondon, M. et al., 2008. Breaking of water-in-crude oil emulsions. 2. Influence of Asphaltene concentration and diluent nature on demulsifier action. Energy & Fuels, 22, pp.702-07. Rother, M.A., Zinchenko, A.Z. & Davis, R.H., 1997. Buoyancy-Driven Coalescence of Slightly Deformable Drops. Journal of Fluid Mechanics, pp.117-48. Saboni, A., Gourdon, C. & Chesters, A.K., 1995. Drainage and rupture of partially mobile films during coalescence in liquid-liquid systems under a constant interaction force. Journal of Colloid and Interface Science, 175, pp.27-35. S., 2009. Sigmoid Pharma. [Online] Available [Accessed at: 4

http://sigmoidpharma.com/dynamicdata/ProprietaryTechnology.asp

36

Saboni, A., Gourdon, C. & Chesters, A.K., 1995. Drainage and Rupture of Partially Mobile Films during Coalescence in Liquid-Liquid Systems under a Constant Interaction Force. Journal of Colloid and Interface Science, pp.27-35. Sacanna, S., Kegel, W.K. & Philipse, A.P., 2007. Thermodynamically stable pickering emulsions. Physical Review Letters, 98, p.158301. Schmitt, V. & Leal-Calderon, F., 2004. Measurement of the coalescence frequency in surfactant-stabilized concentrated emulsions. Europhysics Letters, 67(4), pp.662-68. Sjoblom, J. et al., 2003. Our Current Understanding of Water-inCrude Oil Emulsions. Recent Characterization Techniques and High Pressure Performance. Advances in Colloid and Interface Science, pp.100-02. Slattery, J.C., 1990. Interfacial Transport Phenomena. New York: Springer-Verlag. Smoluchowski, M.V., 1917. Versuch einer mathematischen theorie der

koagulationskinetik kolloider losungen. Z Physical Chemistry, 92, pp.129-68. Solutions, O.S., 2002. Oil Spill Solutions. [Online] Available at:

http://www.oilspillsolutions.org/controlledburning.htm [Accessed 4 April 2010]. Stauffer, C.E., 1965. Journal of Physical chemistry, 69, p.1933. Stergios, G., Yiantsios, G. & Davis, R.H., 1991. Close Approach and Deformation of Two Viscous Drops due to Gravity and Van Der Waals Forces. Journal of Colloid and Interface Science, pp.412-33. Sullivan, A.P., Zaki, N.N., Sjoblom, J. & Kilpatrick, P.K., 2007. The stability of water-in-crude and model oil emulsions. The Canadian Journal of Chemical Engineering, 85, pp.793-807. Sundararaj, U. & Macosko, C.W., 1995. Drop Breakup and Coalescence in Polymer Blends - the Effects of Concentration and Compatibilization. Macromolecules, 28, p.2647. Tadros, T.F., 1996. Correlation of viscoelastic properties of stable and flocculated suspensions in thickness. Journal of American Chemical Society, pp.3074-78. Tadros, T.F., 2005. Applied Surfactants : Principles and Applications.

37

Thomson, J., 1855. On certain curious motions observable at the surfaces of wine and other alcoholic liquors. Philosophical Magazine Series 4, 10(67), pp.330-33. Venugopal, B.V. & Wasan, D.T., 1983. Encyclopedia of Emulsion Technology. 2nd ed. New York: Marcel Dekker. Vrij, A., 1964. Light scattering by soap films. Journal of Colloids and Interface Science, 19, pp.1-27. Vrij, A. & Overbeek, J.T.G., 1968. Rupture of thin liquid films due to spontaneous fluctuations in thickness. Journal of American Chemical Society, 90, pp.3074-78. Wang, W., Gong, J., Ngan, K.H. & Angeli, P., 2009. Effect of glycerol on the binary coalescence of water drops in stagnant oil phase. Chemical Engineering Research and Design, 87, pp.1640-48. Xia, L., Lu, S. & Cao, G., 2004. Stability and Demulsification of Emulsions Stabalized by Asphaltenes or Resins. Journal of Colloid Interface Science, pp.50406. Yang, S.-M., Leal, L.G. & Kim, Y.-S., 2002. Hydrodynamic interaction between spheres coated with deformable thin liquid films. Journal of Colloid and Interface Science, 250, pp.457-65. Yeo, L.Y., Matar, O.K., Perez de Ortiz, E.S. & Hewitt, G.F., 2003. Film drainage between two surfactant-coated drops colliding at constant approach velocity. Journal of Colloid and Interface Science, 257, pp.93-107. Yoon, Y., 2006. Flow-Induced coalescence of Two Deformable Drops. Ph. D. Thesis. Santa Barbara: University of California. Yoon, Y., Hsu, S. & Leal, L.G., 2007. Experimental Investigation of the Effects of Copolymer Surfactants on Flow-Induced Coalescence of Drops. Physics of Fluids, 19, p.023102. Zapryanov, Z., Malhotra, A.K., Aderangi, N. & Wasan, D.T., 1983. Emulsion Stability : An Analysis of the Effects of Bulk and Interfacial Properties on Film Mobility and Drainage Rate. International Journal of Multiphase Flow, p.105. Zdravkov, A.N., Peters, G.W.M. & Meijier, H.E.H., 2003. Film drainage between two captive drops: PEO-water in silicon oil. Journal of Colloid and Interface Science, 266, pp.195-201.

38

39

Вам также может понравиться