+ All Categories
Home > Documents > T H E U N I V E R S I T Y O F T U L S A THE GRADUATE...

T H E U N I V E R S I T Y O F T U L S A THE GRADUATE...

Date post: 09-Mar-2018
Category:
Upload: buicong
View: 215 times
Download: 1 times
Share this document with a friend
218
T H E U N I V E R S I T Y O F T U L S A THE GRADUATE SCHOOL HORIZONTAL PIPE SEPARATOR (HPS © ) EXPERIMENTS AND MODELING by Ciro Andrés Pérez A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Discipline of Petroleum Engineering The Graduate School The University of Tulsa 2005
Transcript

1

T H E U N I V E R S I T Y O F T U L S A

THE GRADUATE SCHOOL

HORIZONTAL PIPE SEPARATOR (HPS©)

EXPERIMENTS AND MODELING

by Ciro Andrés Pérez

A dissertation submitted in partial fulfillment of

the requirements for the degree of Doctor of Philosophy

in the Discipline of Petroleum Engineering

The Graduate School

The University of Tulsa

2005

iii

ABSTRACT

Pérez, Ciro A. (Doctor of Philosophy in Petroleum Engineering).

Horizontal Pipe Separator (HPS©). Experiments and modeling

Directed by Professors Ovadia Shoham and Ram S. Mohan (205 pp., Chapter 6)

(329 words)

The objective of this study is to investigate experimentally and theoretically the

developing region of oil-water flow in horizontal pipes. The study aims at using the

developing region of the pipe as an oil-water separator (Horizontal Pipe Separator,

HPS©).

An experimental HPS© facility has been designed and constructed, to enable

measurements of local parameters in oil-water flow in the developing region of the flow

in a 3.75-in.-ID 19.33-ft-long acrylic pipe. Special instrumentation was developed for

acquiring the local parameters data, namely, local velocity profiles; water cut profiles and

droplet size distribution. Experimental data were acquired for mixture velocities of 0.44

and 0.58 ft/s, and water cuts of 10, 30, 50 and 70%, measured at two metering stations,

located at 7.5 ft and 13.5 ft from the inlet, respectively. The data were acquired for a

concentric inlet with and without a mixer, and for three different outlet configurations.

Also, inlet flowrates as well as the water cut in both the oil and water outlet were

iv

measured as functions of the split ratio. For all experimental runs, the flow did not reach

fully developed flow conditions.

A model is developed for the prediction of the flow evolution in the developing

region of the HPS©. The model comprises two sub-models: one-dimensional flow of

three layers (hydrodynamic sub-model) and population balance coalescence theory

(coalescence sub-model). The three layers are, from top to bottom (for water-continuous

flow at the inlet): pure oil, packed dispersion of oil in water and loose dispersion of oil in

water. For oil-continuous flow the three layers are loose dispersion of water in oil, packed

dispersion of water in oil and clear water. Linear velocity and water cut profiles were

assumed for the intermediate (packed dispersion) layer. Average and minimum water cut

of the intermediate (packed dispersion) layer are required as input. The results of the

model match fairly well the experimental data, with respect to layer height development;

velocity and water cut profiles and overall droplet size distributions.

v

ACKNOWLEDGEMENTS

I really want to give my deepest gratitude to my advisor, Dr. Ovadia Shoham, for

his support and confidence during the development of this study. I also want to thank Dr.

Ram Mohan, my co-advisor for his help and reviews during the different phases of this

study. Thanks are due also to Dr. Luis Gómez, Dr. Shoubo Wang and Dr. Gene Kouba,

for their valuable suggestions and support. I am very grateful to The University of Tulsa,

and to the Tulsa University Separation Technology Project (TUSTP) for the financial

support and opportunity to accomplish this endeavor. I would like to thank all the TUSTP

members and graduate students for the time we invested sharing ideas, and for the

friendship they showed during this time, especially to Dr. Nólides Guzman, Mr. Carlos

Avila and Dr. Carlos Torres. I am especially grateful to Mrs. Judy Teal for her help, and

to Oscar Escobar and Rafael Rivas. Thanks to Marisabel Herrera, Jose Alaña, Mariela

Lander, and all the people that made this whole experience richer. Also, thanks are due to

the “LABCEM” Laboratory (specially to Nathaly Moreno, Andrés Tremante and Frank

Kenyery) and to the “Departamento de Termodinamica y Fenomenos de Transporte”,

both at the Universidad Simón Bolívar, in Caracas, Venezuela, for all the help they gave

me to accomplish this milestone. Finally, all my gratitude to my family, that helped me so

much. I would like to dedicate this work to m y parents Benigno and Fidela, my sisters

Maria Eglee, Carmen Alicia and Simone, and my brother Pedro.

vi

TABLE OF CONTENTS

Page

ABSTRACT....................................................................................................................... iii ACKNOWLEDGEMENTS.................................................................................................v TABLE OF CONTENTS................................................................................................... vi LIST OF TABLES............................................................................................................. ix LIST OF FIGURES .............................................................................................................x CHAPTER 1: INTRODUCTION ......................................................................................1 CHAPTER 2: LITERATURE REVIEW ..........................................................................4 2.1 Two-Phase Fully Developed Liquid-Liquid Flow .........................................4 2.1.1 Flow Patterns ......................................................................................5 Flow Patterns Classification and Flow Pattern Maps ......................7 Flow Pattern Prediction..................................................................15 2.1.2 Pressure Drop ...................................................................................18 2.2 Liquid-Liquid Developing Flow Region.......................................................19 2.2.1 Effects of Inline Mixing .....................................................................21 2.2.2 Effect of Pre-Mixing ..........................................................................21 2.3 Measurement of Local Parameters in Oil-Water Flow ..............................23 2.3.1 Velocity..............................................................................................23 2.3.2 Local Holdup.....................................................................................23 2.3.3 Local Droplet Size Distribution ........................................................24 2.3.4 Local Continuous Phase Measurement.............................................24 2.4 Coalescence/Breakup and Droplet Size Distribution..................................24 2.4.1 Droplet Coalescence .........................................................................25 2.4.2 Droplet Breakup................................................................................30 2.4.3 Probability Density Functions ..........................................................30 Continuous-Size Distribution.........................................................31 2.4.4 Sauter Mean Diameter ......................................................................33 2.5 Outlet Studies in Horizontal Pipes ...............................................................34 2.6 Use of Horizontal Pipes as Separators .........................................................35

vii

CHAPTER 3: EXPERIMENTAL PROGRAM .............................................................36 3.1 Experimental Facility ....................................................................................36 3.1.1 Storage and Metering Section ...........................................................37 3.1.2 Test Section .......................................................................................39 3.1.3 Local Measurement Instrumentation ................................................43 3.1.4 Gas-Oil-Water Separation Section ...................................................45 3.1.5 Data Acquisition System ...................................................................46 3.1.6 Working Fluids..................................................................................47 3.2 Experimental Test Matrix .............................................................................49 3.2.1 Velocity Profiles at Vertical Plane, and Velocity Surfaces ...............50 3.2.2 Water Cut Profiles.............................................................................54 3.2.3 Layer Height......................................................................................57 3.2.4 Droplet Size Distribution Profiles.....................................................58 3.2.5 Pressure Drop ...................................................................................67 3.2.6 Outlets Performance .........................................................................70 CHAPTER 4: MODELING..............................................................................................74 4.1 Hydraulic Sub-Model ....................................................................................74 4.1.1 Number of Layers ..............................................................................74 4.1.2 Layer Mixture Properties ..................................................................77 4.1.3 Mathematical Formulation ...............................................................79 Taitel et. al. (1995) 3-layered model..............................................79 4.2 Coalescence Sub-Model .................................................................................85 4.2.1 Physical Phenomena .........................................................................85 4.2.2 Assumptions.......................................................................................87 4.2.3 Mathematical Formulation ...............................................................88 Estimation of the number of collisions per unit volume per unit time .........................................................................................88 Estimation of the number of coalescing collisions per unit volume per unit time ......................................................................89 4.3 Closure Rules..................................................................................................91 4.3.1 Estimation of the Settling Velocity ....................................................91 4.3.2 Estimation of the Velocity and Water Cut Profiles in the Packed Layer ....................................................................................92 4.3.3 Estimation of the Local Droplet Size Distribution in the Packed Layer ....................................................................................98 Evolution of the Local Distribution Parameters Along the Separator ......................................................................................106 4.3.4 Coalescence Estimation Procedure ................................................106 4.4 Calculation Procedure .................................................................................108 CHAPTER 5: RESULTS AND DISCUSSION .............................................................111 5.1 Comparison of Layers Height Evolution ...................................................111

viii

5.2 Comparison of Velocity and Water Cut Profiles ......................................115 5.3 Comparison of Droplet Size Distribution Evolution in Packed Layer....119 5.4 Comparison of Droplet Size Distribution as a Function of the Height in Packed Layer...........................................................................................123 CHAPTER 6: CONCLUSIONS AND RECOMMENDATIONS ...............................129 NOMENCLATURE ........................................................................................................134 REFERENCES ................................................................................................................138 APPENDIX I: LOCAL MEASUREMENT SYSTEMS............................................145 APPENDIX II: LAYER HEIGHT COMPARISON BETWEEN LOCAL WATER CUT AND PHOTOGRAPHIC METHODS MEASUREMENTS .......................................................176 APPENDIX III: LOCAL VELOCITY MEASUREMENT .........................................180 APPENDIX IV: CALCULATION OF THE VELOCITY AND WATER CUT SLOPES FOR LINEAL VELOCITY PROFILE APPROXIMATION ...........................................................................197 APPENDIX V: DROPLET SIZE MEASUREMENT ................................................205

ix

LIST OF TABLES

Table 2.1 Classification of Gas-Liquid Flow Patterns (Ishii, 1975) ................................. 14

Table 2.2 Classification of Liquid-Liquid Flow Patterns (Kurban, 1997)........................ 14 Table 3.1 Properties of Water-Phase ................................................................................ 47

Table 3.2 Properties of Oil-Phase ..................................................................................... 48

Table 3.3 Average WC in Packed Dispersion Layer (7.5 ft and 13.5 ft).......................... 55

Table 3.4 Average WC in Loose Dispersion Layer (7.5 ft and 13.5 ft)............................ 55

Table 3.5 Dimensionless (h/D) Height of Packed Dispersion Layer-Loose Dispersion

Layer Boundary (7.5 ft) ........................................................................................... 57

Table 3.6 Dimensionless (h/D) Height of Packed Dispersion Layer-Loose Dispersion

Layer Boundary (13.5 ft) .......................................................................................... 57

Table 3.7 Log-Normal Distribution Fitting Parameters for Cumulative Distributions

in Figure 3.21 ............................................................................................................ 65

Table 3.8 Log-Normal Distribution Fitting Parameters for Cumulative Distributions

in Figure 3.22 ............................................................................................................ 65

Table 5.1 Values of b Used to Adjust the Local Droplet Size Distribution in Model ....124

x

LIST OF FIGURES

Page

Figure 2.1 Relative Location of Flow Patterns for Light Oil (μO<20 cP) and Water Flow

With Same Density in a Mandhane (1947) Flow Pattern Map. (from Charles et al.,

1961) ........................................................................................................................... 8

Figure 2.2 Relative Location of Flow Patterns for Oil (65 cP) and Water Flow With Same

Density in a Mandhane Flow Pattern Map (From Charles et al., 1961)..................... 8

Figure 2.3 Flow Pattern Map for Water and Oil Flow, With Oil Viscosity of 21.7 mPa*s

in a 39.4 mm ID Pipe. (after Guzhov et al., 1973) ................................................... 10

Figure 2.4 Horizontal Oil-Water Flow Patterns. (after Trallero, 1995)............................ 10

Figure 2.5 Experimental Flow Pattern Map Using Superficial Velocities as Coordinates

(after Trallero, 1995)................................................................................................. 12

Figure 2.6 Experimental Flow Pattern Map Using Mixture Velocity and Input Water Cut

as Coordinates (after Trallero, 1995) ........................................................................ 12

Figure 2.7 ZNS and ZRC Transition Boundaries ............................................................. 17

Figure 2.8 Stratified-Stratified Dispersed Flow Boundary ............................................... 17

Figure 2.9 Schematic of Trallero (1995) Inlet Mixer ....................................................... 20 Figure 3.1 Experimental Facility ...................................................................................... 38

Figure 3.2 Storage and Metering Section ......................................................................... 39

Figure 3.3 HPS© Test Section........................................................................................... 39

Figure 3.4 HPS© Inlet Section .......................................................................................... 40

xi

Figure 3.5 KOMAXTM Static Mixer Spool Upstream of the Inlet Section....................... 41

Figure 3.6 HPS© Outlet Configurations............................................................................ 41

Figure 3.7 Location of Local Measurement Ports and Pressure Tap Ports....................... 42

Figure 3.8 Experimental Test Matrix Shown on Steady State Flow Pattern Map........... 49

Figure 3.9 Velocity Profiles at Vertical Plane (vM=0.44 ft/s) ........................................... 51

Figure 3.10 Velocity Profiles at Vertical Plane (vM=0.58 ft/s) ......................................... 51

Figure 3.11 Velocity Contours at 7.5 ft from Inlet (vM=0.44 ft/s) .................................... 52

Figure 3.12 Velocity Contours at 13.5 ft from Inlet (vM=0.44 ft/s) .................................. 52

Figure 3.13 Velocity Contours at 7.5 ft from Inlet. (vM=0.58 ft/s)................................... 53

Figure 3.14 Velocity Contours at 13.5 ft. from Inlet. (vM=0.58 ft/s)................................ 53

Figure 3.15 Water Cut Profiles at Vertical Plane (vM=0.44 ft/s) ...................................... 56

Figure 3.16 Water Cut Profiles at the Vertical Plane. (vM=0.58 ft/s) ............................... 56

Figure 3.17 Droplet Size Distribution Profiles at Vertical Plane (vM=0.44 ft/s) .............. 59

Figure 3.18 Droplet Size Distribution Profiles at Vertical Plane (vM=0.58 ft/s) .............. 60

Figure 3.19 d32 Profiles at Vertical Plane (vM=0.44 ft/s) .................................................. 61

Figure 3.20 d32 Profiles at Vertical Plane (vM=0.58 ft/s) .................................................. 61

Figure 3.21 Overall Droplet Size Distribution in Packed Dispersion Layer (vM=0.44 ft/s)

................................................................................................................................... 63

Figure 3.22 Overall Droplet size Distribution in Packed Dispersion Layer (vM=0.58 ft/s)

................................................................................................................................... 64

Figure 3.23 Pressure Drop Along HPS© for Different Water Cuts, Without Mixer......... 69

Figure 3.24 Pressure Drop Along HPS©, With Mixer ...................................................... 69

Figure 3.25 Oil Cut at Oil Outlet (vM=0.44 ft/s) ............................................................... 72

xii

Figure 3.26 Oil Cut at Oil Outlet (vM=0.58 ft/s) ............................................................... 73 Figure 4.1 Three Layer Developing Flow. (vM=0.44 ft/s, 50% WC, without Mixer, 14 ft

from Inlet) ................................................................................................................. 75

Figure 4.2 Force Balances Over Each of the Three Layers. (after Taitel et. al., 1995).... 80

Figure 4.3 Coalescence in Simple Shear Flow due to Velocity Gradient......................... 86

Figure 4.4 Schematic of Proposed Velocity and WC Profiles in the Packed Dispersion

Layer ......................................................................................................................... 93

Figure 4.5 Cumulative Frecuency of Selected Droplet Diameters as Function of Height

Inside the HPS© (70%WC, vM=0.44 ft/s, w/mixer, 7.5 ft from the inlet) ................. 99

Figure 4.6 Cumulative Frequency of Selected Droplet Diameters as Function of Height

Inside the HPS© (70%WC, vM=0.58 ft/s, w/mixer, 7.5 ft from the inlet) ................. 99

Figure 4.7 Procedure of Assignment of Local Droplet Size Distributions as Function of

Height (nb=5).......................................................................................................... 104

Figure 4.8 Calculation Procedure Flowchart .................................................................. 110

Figure 5.1. Comparison of Model Predictions and Experimental Data for Layer Heights

Evolution (vM=0.44 ft/s)..........................................................................................113

Figure 5.2. Comparison of Model Predictions and Experimental Data for Layer Heights

Evolution (vM=0.58 ft/s).......................................................................................... 114

Figure 5.3 Comparison of Model Predictions and Experimental Data for 30% WC.

Mixture Velocities vM=0.44 and 0.58 ft/s ............................................................... 116

Figure 5.4 Comparison of Model Predictions and Experimental Data for 50% WC.

Mixture Velocities vM=0.44 and 0.58 ft/s ............................................................... 117

xiii

Figure 5.5 Comparison of Model Predictions and Experimental Data for 70% WC. Mix.

Velocities vM=0.44 (a) and 0.58 (b) ft/s .................................................................. 118

Figure 5.6. Comparison of Model Predictions and Experimental Data for Droplet Size

Distribution Change between Metering Stations (vM=0.44 ft/s) ............................. 120

Figure 5.7 Comparison of Model Predictions and Experimental Data for Droplet Size

Distribution Change between Metering Stations (vM=0.58 ft/s) ............................. 121

Figure 5.8 Comparison of Model Adjusted and Experimental Measured Droplet Size

Distributions at 7.5 ft from Inlet, vM=0.44 ft/s........................................................ 125

Figure 5.9 Comparison of Model Adjusted and Experimental Measured Droplet Size

Distributions at 7.5 ft from Inlet, vM=0.58 ft/s........................................................ 126

Figure 5.10 Comparison of Model and Experimental Measured Droplet Size

Distributions at 13.5 ft from Inlet, vM=0.44 ft/s...................................................... 127

Figure 5.11 Comparison of Model and Experimental Measured Droplet Size

Distributions at 13.5 ft from Inlet, vM=0.58 ft/s...................................................... 128

1

CHAPTER 1

INTRODUCTION

The actual world’s energy demand requires further advances in the knowledge of

the generation, production and development of proven and new energy sources. To this

end, the nuclear industry, and more recently the petroleum industry, have been driven to

study the flow of two or three phases through production and processing facilities. These

studies are aimed at the challenges of defining the flow pattern, pressure gradient, phase

volume fractions and separation efficiency of these multiphase flows.

As the need for energy production increases, the requirement of production of

more energy at lower costs is pursued. The gained knowledge of the flowing phenomena

in multiphase flow enables optimization of the different components of the entire

production and processing infrastructure.

The requirements of oil-water separation in more challenging environments,

especially in sub-sea production, give rise to the need of alternative separation

technologies that can decrease deployment expenses, and increase the robustness of the

energy production process.

2

The natural segregation of crude oil and water due to density difference can occur

not only in gravity based vessel-type separators (widely used by the petroleum industry),

where the fluids have large residence time, but also when flowing through pipes, if the

flowing conditions are favorable for flow segregation.

Thus, the use of pipes as separators is especially suitable for sub-sea applications.

The ease of installation and simplicity of operation of pipe separators ensure reliable

performance of the entire production system. The proposed Horizontal Pipe Separator

(HPS©)1 is a simple concept: a pipe spool with appropriate geometry promoting natural

separation of the phases under favorable flow conditions. However, not many studies

have been published on HPS©, especially addressing the developing flow along it. Thus,

the objectives of this study are:

1- Study the behavior of oil-water mixtures in horizontal pipes;

2- Develop a mechanistic model that predicts separation efficiency for given fluids,

geometry and flow rates;

3- Compare/refine the model with data obtained in this study and from literature;

4- Develop a computational code based on the developed model..

This dissertation is divided into six chapters. Chapter 1, the current one, is this

introduction. The description of the following five chapters follows:

1 HPS© – Horizontal Pipe Separator – Copyright, The University of Tulsa, 1999

3

Chapter 2 presents a literature review on the topics related to oil-water flow in

pipes and to droplet coalescence mechanisms. Chapter 3 gives a description of the

experimental facility, the instrumentation used, the experimental matrix and testing

procedures; finally the measured data are presented. Chapter 4 provides details of the

proposed model for the prediction of the flow evolution in the developing region of the

HPS©. A comparison between the model predictions and the experimental data is given in

Chapter 5, along with the discussion of the results. Finally, Chapter 6 provides the

conclusions and the recommendations.

4

CHAPTER 2

LITERATURE REVIEW

There is a limited availability of experimental data for liquid-liquid flow (as

compared to gas-liquid flow), despite two peaks in the rate of publication of liquid-liquid

studies: the first peak occurred in the early 1960's (Charles et al., 1961; etc.), and the

second in the past 10 years. Still today, many of published models for liquid-liquid flows

represent extensions of gas-liquid models rather than an original methodology.

This chapter presents literature review on liquid-liquid oil-water flow phenomena

in pipes as follows: (1) Fully developed liquid-liquid flow; (2) Liquid-liquid developing

flow region; (3) Measurement of local parameters in oil-water flows; (4)

Coalescence/break up and droplet size distributions; (5) Outlets studies in horizontal

pipes; and (6) Use of horizontal pipes as separators.

2.1 Two Phase Fully Developed Liquid-Liquid Flow

This section presents previously published studies on fully developed liquid-

liquid flow, including flow pattern definitions and pressure drop phenomena.

5

2.1.1 Flow Patterns

In two-phase flow in pipes, the deformable interface between the immiscible

phases can exist in different shapes or spatial distributions for different flow conditions.

These configurations are commonly called flow patterns.

For gas–liquid systems, the flow patterns are considered as functions of the

following variables (Shoham, 2006): (a) operational parameters, namely, the gas and

liquid flowrates, (b) geometrical variables, including pipe diameter and inclination angle,

and (c) the physical properties of the two phases, i.e., gas and liquid densities, viscosities

and the surface tension. These same parameters are thought to control the flow pattern

phenomena for liquid-liquid mixtures flowing in pipes (i.e., Trallero, 1995; Kurban,

1997). More recent studies such as Angeli (1996), Soleimani (1999) and Shi (2001)

suggested that the flow pattern configuration is also a function of the pipe material, and

of the presence of additional components (such as surfactants) in the mixture. Most of

these studies were executed on small-diameter pipes, namely, 1-inch and 2-inch nominal

diameters.

These flow patterns are usually presented in the form of two-dimensional plots,

or flow pattern maps. The importance of the flow pattern maps is that within each pattern

the flow has certain similar hydrodynamic characteristics. The knowledge of these

characteristics simplifies the problem of building a hydrodynamic flow model for the

6

phenomena into the construction of hydrodynamic sub-models appropriate for each flow

pattern.

In order to differentiate between the different flow patterns, several different

methods have been developed and applied for gas-liquid and liquid-liquid flow pattern

detection.

Visual observation is one of the most common methods to identify flow patterns

in multiphase flow. The identification is done by using a photographic/video technique to

view the flow through the wall of a transparent tube. High-speed photography must be

used to capture the flow patterns at high fluid mixture velocities. However, the

disadvantage of the visual observation method is the difficulty to observe the internal

structure of the flow clearly (Hewitt et al., 1997). This problem can be mitigated by other

visualization alternatives such as X-ray photography.

Since photographic methods are somewhat subjective and not generally reliable,

more objective and quantitative detection methods have been developed. Conductivity

probes (Angeli, 1996; and Trallero, 1995) provide a more precise method of

investigating the spatial distribution of two phases across the cross section of the tube

(Angeli and Hewitt 1998). Valle and Kvandal (1995) used local sampling to measure the

liquid fraction at points across the cross section of a pipeline. Gamma ray densitometers

(Elseth, 2001), and High Frequency Impedance Probes (Lovick 2004) are other devices

used for measuring the local distribution of two phases. All of these alternative tech-

7

niques provide valuable additional information for defining flow patterns in a more

objective way.

Flow Patterns Classification and Flow Pattern Maps

From the initial studies of Russel et al. (1959), Russel & Charles (1959), Charles

et al. (1961) and Charles & Redberger (1962), different but related flow patterns had

been defined. Charles et al. (1961) observed the flow patterns that occurred during the

flow of equal density oil and water mixture in a horizontal, 1-inch ID cellulose acetate

butyrate pipe. Three oils with dynamic viscosities of 6.29, 16.8 and 65 mPa*s were used.

A schematic of the resulting flow pattern maps plotted in terms of the superficial

velocities of the two liquids, are shown in Figure 2.1 for the two lighter oils and in Figure

2.2 for the heavier one. The following flow patterns were defined in this work:(a) Water

droplets in oil, (b) Water bubbles in oil, (c) Oil in water concentric flow, (d) Water slugs

in oil, (e) Oil slugs in water, (f) Oil bubbles in water and (g) Oil droplets in water. It

should be noted that the absence of density difference between the two liquids resulted in

a symmetric flow about the pipe axis. Thus, the stratified flow pattern was not observed.

The flow patterns for the 6.29 and 16.8 mPa*s viscosity oils (Figure 2.1) were almost

identical, with only one oil continuous flow pattern. The flow patterns for the heavier oil

(Figure 2.2) were similar to those already observed, except at the low water velocities,

where a succession of different flow patterns with oil as the continuous-phase occurred,

as shown in Figure 2.2. Charles et al. (1961) attributed this difference to the fact that the

most viscous oil wetted the pipe wall more than the other two oils.

8

Figure 2.1 Relative Location of Flow Patterns for Light Oil (μO<20 cP) and Water Flow

With Same Density in a Mandhane (1947) Flow Pattern Map. (from Charles et al., 1961)

Figure 2.2 Relative Location of Flow Patterns for Oil (65 cP) and Water Flow With Same

Density in a Mandhane Flow Pattern Map. (From Charles et al., 1961)

0.01

0.1

1.0

10. Water droplets in oil

Oil in water concentric

Oil slugs in water

Oil bubbles in water

Oil drops in water

Superficial water velocity [ft/s]

Supe

rfic

ial o

il ve

loci

ty [f

t/s]

0.05 0.1 1.0 10.

Water drops in oil

Oil in water concentric

Oil slugs in water

Oil bubbles in water

Oil drops in water

Superficial water velocity [ft/s]

Supe

rfic

ial o

il ve

loci

ty [f

t/s]

Water slugs in oil

Water bubbles in oil

0.01

0.1

1.0

10

0.05 0.1 1.0 10.

9

Russell et al. (1959), working in a similar pipe with water and oil (viscosity 18

mPa*s and density 834 kg/m3), observed similar flow patterns to those reported by

Charles et al. (1961) but with the asymmetry imposed by the density difference. The

annular flow pattern did not occur at all, while for a wide range of velocities the stratified

flow pattern was present.

Using similar flow pattern classifications to the ones used by Charles et al.

(1961), other flow pattern maps for horizontal flow of oil and water were developed, but

with mixture velocity and water volume fraction as axes. For example, Guzhov et al.

(1973) presented such a flow pattern map, for water and oil with oil viscosity of 21.7

mPa*s and density 896 kg/m3 at 20°C in a horizontal, 39.4 mm ID pipe, which schematic

is shown in Figure 2.3. Also, Arirachakaran et al. (1989) published a similar map for

water and oil, with oil viscosity of 84 mPa*s at 21°C in a 39.3 mm ID pipe.

Since these studies, experimental and phenomenological models have evolved,

and slightly different flow pattern classifications have been presented. There have been a

convergence on similar classifications during the last ten years, as can be seen in the

following examples: Trallero (1995), Nadler and Mewes (1997), and Kurban (1997).

Trallero (1995) acquired experimental data in a two 2-inch nominal 51-ft-long

pipe facility, connected with a U bend. Experimental fluids were mineral oil (Cristex AF-

M 31) and tap water, with density ratio of 0.85, viscosity ratio of 29, and interfacial

tension of 36 dynes/cm (at 78º F).

10

Figure 2.3 Flow Pattern Map for Water and Oil Flow, With Oil Viscosity of 21.7 mPa*s

in a 39.4 mm ID Pipe. (after Guzhov et al., 1973)

Figure 2.4 Horizontal Oil-Water Flow Patterns (after Trallero, 1995)

Stratified Flow

Water drops and oil layer

Oil drops

Water drops

Intermittent flow

Oil drops and water layer

0 0.5 1. Volume fraction of water

Mix

ture

vel

ocity

[m/s

]

1.

2.

3.

11

The author proposes a flow pattern classification for oil-water flows where six

flow patterns were identified and classified into two categories:

(1) "segregated" flow with two sub-regimes of stratified and stratified with

mixing at the interface;

(2) "dispersed" flow with two sub-regimes of water dominated dispersed flow and

oil dominated dispersed flow.

Figure 2.4 shows the “segregated” flow patterns (on the left hand side), and the

“dispersed” flow patterns (on the right hand side). The author presents the experimental

results using both the Charles et al. (1961) (Figure 2.5) and Guzov et al. (1973) (Figure

2.6) coordinate systems for the flow pattern maps.

Nadler and Mewes (1997) presented a slightly different classification of flow

patterns (the authors added an extra flow pattern: dispersion of water in oil-dispersion of

oil in water-pure water flow pattern). Their facility consisted of a 48-m-long, 59-mm (2.3

inch)-ID test section. Mineral oil and tap water were used, with a density ratio of 0.85,

and a viscosity ratio from 28 to 35, in the range of 18ºC to 30ºC. They plotted their

results in both flow pattern map types.

12

Figure 2.5 Experimental Flow Pattern Map Using Superficial Velocities as Coordinates

(after Trallero, 1995)

Figure 2.6 Experimental Flow Pattern Map Using Mixture Velocity and Input Water Cut

as Coordinates (after Trallero, 1995)

13

Finally, Kurban (1997) presented a flow pattern classification based on the study

of Ishii (1975) for gas-liquid flows. Experimental results were acquired using two

facilities: 1) 8-m-long stainless steel and acrylic pipes of 1-inch nominal diameter, using

EXXSOL D-80 mineral oil and tap water (oil-water density Ratio 0.8, viscosity ratio

1.6, interfacial tension 0.017 N/m, operating temperature 25ºC), and 2) 42-m-long, 77.90-

mm (3 inches) diameter, stainless steel pipe, using water and Shell TELLUS 22 mineral

oil (Density Ratio 0.865, viscosity ratio 45, with no interfacial tension nor operating

temperature reported).

The work of Ishii (1975) provides a classification of two-phase gas-liquid flows

based on the interfacial structures and topology of each phase. The three main categories

were separated flows, mixed or transition flows and dispersed flows, as shown in Table

2.1.

Kurban combined the categories of Ishii (1975), with the experimental

observations obtained from the Imperial College TOWER and WASP rigs (Kurban,

1997) and those reported by Guzhov et al. (1973), Oglesby (1979) and Arirachakaran et

al. (1989), as given by Table 2.2. Kurban (1997) proposes that although the phases for all

the studies were oil and water, such classifications should be valid for any two-phase

flow of two immiscible liquids. He presented his results in terms of dimensionless

variables (namely the generalized Lockhart-Martinelli parameter, non-dimensional wall

14

Table 2.1

Classification of Gas-Liquid Flow Patterns (Ishii, 1975)

Category Flow Regime Configuration

Stratified Liquid layer below gas with a planar interface

Separated

Annular Gas core and liquid film

Slug Gas pockets in liquid

Mixed or transition

Annular with Entrainmnent

Gas core with droplets and liquid film with gas bubbles

Bubbly Gas bubbles in liquid continuous phase

Dispersed

Droplet Liquid droplets in gas

Table 2.2

Classification of Liquid-Liquid Flow Patterns (Kurban, 1997)

Category Flow Regime Configuration

Stratified Water layer in oil with a near planar interface

Separated

Core-annular Oil core-water annular film. Circular, and non pipe-concentric interface

Intermittent Phases alternately occupying the pipe as a free and as a dispersed phase

Stratified with dispersion

Layers of a dispersion with a free phase

Mixed or transition

Annular with dispersion

Oil core with water droplets and water film with oil droplets

Water in oil Water droplets in oil continuous phase

Dispersed

Oil in water Oil droplets in water continuous phase

15

and interfacial shear stresses and the superficial friction factors) in order to provide

generalization of the results.

Flow Pattern Prediction

Different mechanistic models have been proposed for the prediction of different

flow pattern boundaries. An example is the Brauner and Moalen Maron (1992 a,b)

criteria for the transition from one flow regime to the other, as presented next. In this

model the considered transitions are mainly the boundaries between stratified, annular

and stratified-dispersed flow regimes.

For the stratified flow pattern, Brauner and Moalen Maron (1992 a,b) presented a

temporal stability analysis of the governing continuity and momentum equations and

considered the conditions under which these equations constitute a well-posed initial

value problem. This analysis produces two transition lines, the zero neutral stability

(ZNS) and the zero real characteristics (ZRC) line. These lines are shown in Figure 2.7,

as functions of the superficial velocities of the two phases for a particular oil-water

system. For this case, the upper layer is oil and the lower is water. The authors stated that

the ZNS line represents the transition from stratified-smooth to stratified-wavy flow

pattern, while the ZRC line represents an upper boundary for the existence of the wavy-

stratified configuration, beyond which other flow patterns exist. According to their

analysis the area identified by the ZRC line (corresponding to the stratified flow pattern),

diminishes in size when the density difference between the two phases decreases, the

16

viscosity difference between phases increases, and the tube diameter decreases. The

departure from the stratified flow pattern can lead either to annular or stratified-dispersed

flow.

In annular flow, Branuer and Moalen Maron (1992 (b)) assumed a certain ratio of

wall to core phase that can produce large interfacial waves, capable of blocking the core

space and leading to slug flow, and derived an equation for the transition line between

these two regimes. They found that this transition line was not sensitive to fluid

properties and tube diameter.

The stratified-dispersed flow transition on the other hand, may exist as a sub-

division of stratified flow, when one phase flowrate is both considerably smaller than the

other phase flowrate and is flowing in the form of entrained droplets. As a result of

buoyancy forces, these droplets tend to concentrate at the top or the bottom of the pipe,

depending on whether the droplets are lighter or heavier than the continuous phase,

respectively. If the buoyancy force exceeds the surface tension force, the authors

suggested that the droplets will merge together to form a continuous layer.

A transition line based on the equation of the balance of these two forces can then

be derived. One such line for a particular oil-water system is shown in Figure 2.8.

Brauner and Moalen Maron (1992 (b)) analysis showed that a decreasing density or

viscosity difference or an increasing surface tension will result in a larger stratified-

dispersed region. Also, as the tube diameter decreases, the stratified-dispersed flow

pattern is more likely to occur.

17

Figure 2.7 ZNS and ZRC Transition Boundaries

(after Brauner and Moalen Maron, 1991 b)

Figure 2.8 Stratified-Stratified Dispersed Flow Boundary

(after Brauner and Moalen Maron, 1991 b)

Superficial oil velocity [m/s]

Supe

rfic

ial w

ater

vel

ocity

[m/s

]

ZNS

ZRC

ZRC

ReW=1500

vW= vO

10-5

10-4

10-3

10-2

10-1

100

10-3 10-2 10-1 100 101

Superficial oil velocity [m/s]

Supe

rfic

ial w

ater

vel

ocity

[m/s

]

ZRC

ZRC

Stratified Dispersed

10-3 10-2 10-1 100 101 10-5

10-4

10-3

10-2

10-1

18

2.1.2 Pressure Drop

Most of the published oil-water flow studies are related to pressure drop gradient,

as this is one of the most important parameters in pipeline design. The interest in early

investigations of the pressure gradient in liquid-liquid flow originated from the idea of

injecting water into the pipeline as a drag reducing agent to reduce the pumping power

requirement. Clark and Shapiro (1949) reported that injection of 7-24% water in the oil

pipeline reduced the pressure gradient by factors from 7.8 to 10.5 in laminar flow. The

authors reported that the maximum pressure reduction occurred when 8-10% water was

injected into heavy crude. In general, the reduction factor depends on the ratio of the oil

to water viscosity (Russel et. al., 1959 and Charles et al., 1961).

In recent years, there has been an increasing need to evaluate the pressure gradient

for oil-water flow originating totally from production well streams or from old fields (i.e.,

without injection of extra water). For the latter case, the interface between the oil and the

underlying aquifer, in water-driven reservoirs, becomes close to the production well or

near the zone where secondary recovery takes place. Under these conditions, dispersions

and emulsions can occur, causing an increase in the pressure gradient (Pal, 1987).

Moreover, the mixtures can exhibit non-newtonian flow behavior, especially when the oil

phase presents natural surfactants (Pal, 1987). Many researchers have attempted to

estimate the pressure gradient as function of both the flow parameters and the given flow

pattern, with model mainly based on 1-D flow geometry, employing either the two fluids,

or drift flux models. Khor (1998) presented a comprehensive review of closure

19

relationships used in liquid-liquid and gas-liquid-liquid stratified pipe flow, for a two

fluid, 1-D model similar to that given by Brauner and Moalen Maron (1989). Comparing

with experimental data, Khor (1998) recommended a sub-set of the closure relationships

that gave the better fitting for both the layer heights and the pressure gradient in the pipe.

2.2 Liquid-Liquid Developing Flow Region

The models for flow pattern and pressure drop prediction in liquid-liquid flow

presented in section 2.1 correspond to steady-state flow conditions. Under steady state

flow conditions; all flow parameters (i.e., local hydrodynamic flow parameters as

velocity or turbulent energy dissipation; or mixture parameters as the mixture interfacial

area concentration) at any location of the cross-sectional area of the pipe are the same

along the pipe. So, the occurrence of stratified smooth flow implies the complete

segregation of oil and water phases, being this the starting condition for the prediction of

the stratified to non-stratified flow transition boundary. As a result, most of the published

experimental studies related to this transition boundary were carried out using low

viscosity oils that allowed the attainment of fully developed flow condition over a short

distance. Also, the inlet sections were designed in such a way that pre-mixing of the

phases was minimized (i.e., Trallero, 1995 (Figure 2.9); also Nadler and Mewes, 1997

and Khor, 1998).

In the developing flow region, the hydrodynamic flow conditions and the mixture

interfacial area concentration per unit volume changes from given inlet conditions

20

towards the steady state flow conditions of the system. As a consequence, the application

of the flow pattern prediction methods shown in the previous section is of limited use.

These models should be used only for predicting the expected steady-state flowing

configuration that a system will reach after some developing length. Gas-liquid pipe

flows usually show short developing lengths (due to the high density difference between

the phases), but liquid-liquid pipe flow developing length can be very long due to the

lower phase density ratios.

Figure 2.9 Schematic of Trallero (1995) Inlet Mixer

There are much less published studies on liquid-liquid two-phase developing flow

than for developed flow conditions. Some of these studies are discussed in the following

sections.

21

2.2.1 Effects of Inline Mixing

Nadler and Mewes (1997) used a 59-mm diameter Perspex pipe for their study.

They observed a reduction of the pressure drop gradient along the pipe on their runs.

They suggested that the measured reduction of the pressure drop gradient was caused by

the development of the flow pattern from stratified at the inlet to dispersion at the outlet,

due to the wetting of the pipe wall by the continuous phase. The results demonstrated that

the higher the viscosity of the oil phase, greater the length required for the flow to

develop, and greater was the pressure drop gradient. At low velocities, stratification was

maintained, so no change of pressure gradient along the pipe was observed. In all tests,

the highest values of the pressure drop gradient were obtained near the inversion point.

2.2.2 Effect of Pre-Mixing

Soleimani (1999) carried out experimental oil-water tests in 25.4-mm ID stainless

steel pipe. The objective of this study was to investigate the effect of pre-mixing of the

flow upstream of the test section. He also reported the results of using a de-swirling unit

downstream of the inlet mixer and before the inlet section.

The viscosity ratio used in his experiment was much smaller (less than 2) than the

one used by Nadler and Mewes (1997). Also, the pressure gradient was used to study the

development of the flow. From the results reported by the Soleimani (1999) for the use of

a single mixer, the pressure gradient decreases as the flow develops for lower flowrates,

22

while the flow segregates and rearranges to a stratified flow configuration. For high

mixture velocities (vM greater than 2 m/s, under dispersed flow regime), the author

reports the effects of installing multiple mixers at the inlet. The extra mixing induced on

the fluids by these mixers increased the pressure drop gradient, indicating the occurrence

of smaller droplets in the mixture. Also, the pressure drop gradient decreased in pipe

sections at larger distance from the inlet due to the segregation of the phases. The steady-

state flow pattern for this velocity was segregated flow, as was evident from the local

hold up analysis presented by the author at 8 m from the inlet. This work did not report

local droplet size distributions.

Lovick (2004) also studied the developing water cut profile in oil-water pipe flow,

and reported velocity profiles and droplet size distribution measurements. Tap water and

EXXOL D140 oil (with density: 828 kg/m3, viscosity: 6 cp and interfacial tension: 39.6

mN/m with tap water) were used as fluids, flowing into two sections of stainless steel

pipe, 38-mm (1.5-inch) ID, 8-m long, connected by a U turn. The experimental facility

could be inclined at ± 10° from the horizontal, and the installation of 1 m acrylic

visualization spool was possible at any location. Settling of oil-in-water dispersions along

the pipe is reported at low velocities of 1 m/s. Slower coalescence is observed, compared

with the previous experiments of Soleimani (1999). The author explains that this

phenomenon is due to the higher viscosity of the oil used in his study (6 cp vs. 1.4 cp).

The author also reports the droplet size distribution and dispersed-phase velocity profiles

at 8 m from the inlet.

23

2.3 Measurement of Local Parameters in Oil-Water Flow

The need for more information on oil-water flow characteristics in liquid-liquid

flow in pipes has prompted studies on local flow characteristics such as the ones carried

out at the Imperial College. Kurban et al. (1995) reported local hold up measurements,

and the measurement of the maximum droplet sizes using conductance probes. Since

then, some other measurement techniques have been reported. Following is a summary of

published studies on local measurements:

2.3.1 Velocity

- Velocity measurement by pitot tubes and/or isokinetic sampling (Khor et al.

(1996), at the Imperial College, and Vedapuri et al. (1997), Shi et al., (1999,

2000), at Ohio University).

- Hot wire anemometry (Farrar et al. (1995), Farrar and Bruun (1996); for vertical

upward flow).

- LDV in the main vertical plane, used to measure stratified flow characteristics, by

Elseth (2001, at NTNU).

- Dual High Frequency Impedance Probe (Lovick 2004)

2.3.2 Local Holdup

- Isokinetic sampling (Khor, 1998), Vedapuri et al. (1997), Shi (2001).

- Single High Frequency Impedance Probe (Angeli 1996, Soleimani, 1999, Lovick

2004).

- Local nuclear densitometry (Elseth, 2002)

24

2.3.3 Local Droplet Size Distribution

- Image analysis (in dispersed flow, Karabelas 1977)

- Laser backscattering and laser diffraction (In dispersed flows: Simmons and

Azzopardi, 2001; and Angeli and Hewitt 2000 a).

- Dual High Frequency Impedance Probe (Lovick, 2004)

- Hot wire anemometry (Farrar and Bruun, 1996)

2.3.4 Local Continuous Phase Measurement

- Conductivity probes (Trallero, 1995; Angeli, 1996; Lovick 2004)

2.4 Coalescence/Breakup and Droplet Size Distribution

As mentioned before, in liquid-liquid two phase flow in pipes, usually one of the

phases disperses as droplets in the other phase, or even the two phases can flow

simultaneously as continuous phases, with some amount of each phase dispersed in the

other, in the form of droplets.

The droplets of the dispersed phase may interact, resulting in droplet coalescence.

Also, the dispersed phase interaction with the continuous phase can lead to breakage of

the larger droplets into smaller ones. In steady state turbulent dispersed flow conditions,

the coalescence of droplets can balance the break up process, reaching equilibrium.

25

Information about coalescence, breakup and droplet size distribution definitions

will be presented in the following sections.

2.4.1 Droplet Coalescence

The coalescing phenomenon can be divided in two steps: collision and film

drainage (Coulaloglou and Tavlarides, 1977).

While studying the collision phenomenon, Prince and Blanch (1990) state that

droplet collisions can occur by different reasons:

- Turbulent collision caused by the effect of the fluctuating turbulent velocity of the

continuous phase on the droplet trajectories.

- Laminar collision caused by velocity gradients in bulk velocity profiles.

- Buoyancy driven collision caused by the difference in bubble rise velocities due

to bubble characteristics and geometry. This phenomenon is most important in

low velocity, vertical flow.

Also, Brownian collisions can occur (Friedlander, 2000) due to Brownian

movement. However, the Brownian effects are usually important for droplet sizes smaller

than the ones that can be subjected to gravitational seggregation in oil-water systems.

26

Laminar collision models take in account the diameter of the droplets involved,

the concentration per unit volume of the droplets of the droplet sizes involved in the

collision, and the overtaking velocities due to shear flow. The pioneering study was given

by Smoluchowski (1915, in German), and presented in a more recent publication by

Friedlander (2000). The collision rate per unit time per unit volume of droplets of

diameters di and dj, in a two-dimensional shear flow field, in rectangular coordinates is

given by:

⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎠⎞

⎜⎝⎛ +=

dydvddCnCnN jijiji

3, )(

61 ................................................................................(2.1)

In this expression Ni,j is the number of collisions between the droplets of

diameters di and dj per unit volume per unit time, Cn is the droplet concentration per unit

volume of each droplet diameter, and dydv is the shear rate. Eq. 2.1 does not have a

dispersed phase concentration restriction, as long as the droplet trajectory is assumed

linear, and the droplets are spherical.

Turbulent collision models assume that the collision mechanism is analogous to

particle collision in ideal gas (Prince and Blanch, 1990). Under this assumption it is

possible to estimate the collision rate as a function of the bubble sizes, concentrations and

velocities, as follows:

5.0222 )''()(16 jijiji vvddCnCnN +⎟

⎠⎞

⎜⎝⎛ +=

π .......................................................................(2.2)

27

In this expression, iv 2' and jv 2' are the mean square fluctuation velocities of the

droplets due to the turbulence. These velocities are assumed equal to that of the turbulent

eddies of the turbulence inertial sub-range of the same size. They can be obtained through

the estimation of the energy dissipation due to turbulence in a homogeneous turbulent

flow (Rotta 1972), as given by:

666.0666.02' 4.1 ii dv ε= ........................................................................................................(2.3)

Where ε is the local turbulent energy dissipation, and the eddy length is

considered the same as the diameter of the bubble, di. This estimation is obtained for low-

concentration dispersions.

Only a fraction of the collisions may lead coalescence due to the existence of a

thin film between the two adjacent droplets that needs to drain for the coalescence to

occur. The interface of the droplet is deformed at the point of contact and the thin liquid

film of the continuous phase gradually drains out. The film at the boundary between the

two droplets eventually collapses, but only when the film is very thin (Oolman and

Blanch, 1986; Chesters, 1991). However, during the drainage process, velocity

fluctuations may provide sufficient energy to the droplets to produce their separation.

As mentioned before, the rate of coalescence depends on the efficiency as well as

the frequency of the collisions, which increases with increasing dispersed phase

28

concentration (Coulaloglou and Tavlarides, 1977). These authors suggested that the

collision efficiency, λ, could be expressed as:

λi,j =exp-(tD-i,j /tC-i,j) ........................................................................................................(2.4)

where tD-i,j, is the continuous-film drainage time (sec), and tC-i,j, is the contact time

between the colliding i, j droplets (sec). In stirred vessels the drainage time is given as a

function of the continuous-phase viscosity and density, interfacial tension, droplet sizes,

agitation rate, and impeller size. The contact time is given as a function of droplet sizes,

and the agitation rate and impeller size.

Oolman and Blanch (1986) and Chesters (1991) suggested that there are three

factors influencing the coalescence process: (1) the external flow field, which determines

the frequency, interaction forces and duration of collisions; (2) the internal flow field in-

volved in drainage of the residual film between the droplets; and finally (3) the

destabilization of very thin films by colloidal forces, leading to film rupture.

The coalescing time is a function of multiple parameters such as: droplet

diameter, phase properties (viscosity of the phases, density of the phases, surface tension)

and contact forces. It also depends on the presence of salt concentration gradients in the

continuous phase as well as the presence of surfactants (Oolman and Blanch, 1986).

29

Chesters (1991) divides the estimation of different coalescing times as functions

of the different boundary conditions the film fluid encounters at the film-droplet interface

while draining. These are:

a) Fully mobile interface: Here, the draining fluid of the film slips over the

droplet surface, so that the droplet does not cause any drag on the film

flow outside the droplet gap. Thus the drainage of the film is only a

function of the film fluid response to deformation (viscosity dependant)

and acceleration (inertia dependant). This interface is characteristic of

the drainage of a liquid film between gas bubbles.

b) Immobile interface: In this model, a no-slip boundary condition is

considered between the liquid film and the droplet surface, and the

velocity at the droplet surface is zero. The interface can be non-

deformable (solid particles as dispersed phases) or deformable

(dispersed phase with a very high viscosity, or when the droplet surface

is saturated with surfactants).

c) Partially mobile interface: An intermediate condition between the

previous two cases. The drainage of the film fluid is dragged to a

certain degree due to viscous internal recirculations on the droplets that

are coalescing, caused by shear at the interface film-droplet.

30

2.4.2 Droplet Breakup

Droplet breakup may occur due to turbulent eddy-droplet interactions, or by

subjecting the droplets to elongational flow fields. Hinze (1955) presented a fundamental

analysis on droplet break-up under different flow configurations. Coulaloglou &

Tavlarides (1977) present an equation similar to Eq. 2.2 for estimating the number of

interactions between droplets and turbulent eddies under turbulent flow conditions. This

topic will not be further developed, as the present study does not consider breakup, under

the assumption that for low phase velocities and Reynolds numbers, the turbulent break-

up is not important and can be neglected.

2.4.3 Probability Density Functions

The study of particle size change in dispersion as function of time usually requires

the use of probability density function for characterization of the droplet size population.

Probability density functions can be built as function of the particle diameter or volume,

(as in separation studies in the petroleum and aerosol industries) or as function of the

particle weight (as in solid particles analysis in the cement industry). These definitions

are different, but related as the particle volume is proportional to the cube of the

diameter, and the weight is proportional to the density of the dispersed phase. The

following definitions will use the particle volume as the distribution parameter.

31

Continuous-Size Distribution

The most used continuous probability density functions used for droplet size

distribution analysis in pipe flow are:

a) Standard Distribution:

⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎠⎞

⎜⎝⎛ −

−=2

21exp

21)(

σμ

πσVVf ...............................................................................(2.5)

where the particle average droplet volume (μ), and the standard deviation (σ) are given

by:

%50VVMED ==μ .............................................................................................................(2.6)

MEDVV %84=σ .......................................................................................................................(2.7)

This distribution usually does not fit properly experimental droplet size

distribution data in liquid-liquid flow.

b) Log-Normal Distribution

The Log-Normal distribution comes from the substitution of the variable in a

normal distribution by the log of the variable. Then the probability density function

becomes:

32

⎟⎟

⎜⎜

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛ −−=

2

0

0

0

ln21exp

21)(

σμ

πσV

VVf ......................................................................(2.8)

and the cumulative frequency distribution is:

∫ ⎟⎟

⎜⎜

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛ −−=

V

VVdV

VF0

2

0

0

0

)(ln21exp

21)(

σμ

πσ............................................................(2.9)

Note that the integral can be solved through the use of the error function (erf)

definition:

⎥⎦

⎤⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛ −+=

0

0ln1

21)(

σμV

erfVF ...................................................................................(2.10)

The mean of a distribution corresponds to the value at which F(x)=0.5, value that

can be obtained from Eq. 2.10 only when the argument of the erf is zero. Thus, the mean

of a Log-Normal distribution is:

0μ =ln(VMedian) ..............................................................................................................(2.11)

as indicated by Crowe et al. (1998). The standard deviation can be obtained from a plot

of the data in a log-probability scale or from the equation:

33

MedianVV %84

0 ln=σ ..............................................................................................................(2.12)

c) Rosin-Rammler Distribution

This distribution is defined as a cumulative distribution function only, not having

a probability density function. The advantage of this distribution is its ease of use.

))/(exp()(1 δaVVF −=− ..........................................................................................(2.13)

F is the cumulative volume fraction of the droplets that have volumes smaller

than V (or 3)6( dπ , as function of the droplet diameter), and a and δ are the adjustable

parameters of the distribution.

2.4.4 Sauter Mean Diameter

Another definition used in the analysis of liquid-liquid dispersions is the Sauter

Mean Diameter. It is defined as the ratio between particle cumulative volume and particle

overall surface area (defined in sprays and atomization literature). Note that this

definition causes the Sauter Mean Diameter to shift towards larger diameters with a small

increase of the frequency of large diameter droplets.

34

For continuous distributions:

∫=

MAX

MAX

d

d

dddfd

dddfdd

0

2

0

3

32

)()(

)()(..................................................................................................(2.14)

And for discrete distributions

=

==TOT

TOT

n

iii

n

iii

dn

dnd

1

2

1

3

32 .............................................................................................................(2.15)

2.5 Outlet Studies in Horizontal Pipes

There are no published studies on the design of an outlet of horizontal pipe

separators. For inclined pipes, Haheim (2001) suggests the use of multiple draining holes

along the pipe, drilled along the top and the bottom of the outlet section, under the

hydraulic requirement of promoting equal liquid drainage of the phases through the holes.

35

2.6 Use of Horizontal Pipes As Separators

A Russian patent (SU 1809911 A3, published in 1993, in Russian) describes the

use of a large diameter pipe as a separator for a tight emulsion of oil and water,

previously treated with a de-emulsifier. Overall experimental results were tabulated, but

no mathematical modeling is given.

Sontvedt and Gramme (1998) filled the World Patent (WO 98/41304) with Norsk-

Hydro as the agent. This patent describes the use of horizontal pipes as wellbore

separators. For patent purposes, experiments were conducted on a 0.78-m ID steel pipe.

Oil and water were injected at vM= 0.6 m/s, with water cut such that the operating

conditions fell at the stratified smooth and stratified with mixing flow pattern boundary

of the reported system flow pattern map. The experimental fluids were water and North

Sea Light crude (ρ=776 kg/m3, μ=1 cp). From the patent information can be inferred that

the estimation of phase segregation was made due to dispersed phase droplet trajectory

calculation, as given by an example with another set of experimental conditions. Through

back-calculation, it is possible to estimate that the developing lengths for their test section

and fluids were of the order of tens to hundreds of meters.

36

CHAPTER 3

EXPERIMENTAL PROGRAM

This chapter presents the most important experimental results, along with a

discussion of their relevance to the flow behavior in the Horizontal Pipe Separator

(HPS©). The complete processed experimental results are included in the Appendixes III

and V.

A description of the experimental facility and test matrix will be given first. This

will be followed by the experimental data acquired.

3.1 Experimental Facility

The three-phase oil-water-gas flow loop located in the College of Engineering and

Natural Sciences Research Building at the North Campus of The University of Tulsa was

utilized in this study. This indoor facility enables experimental investigations throughout

the year. The oil-water-gas indoor facility is a fully instrumented state-of-the-art flow

loop, capable of testing single separation equipment or combined separation systems.

Figure 3.1 shows a photo of the experimental facility. The experimental setup consists of

four major sections: storage and metering section, HPS© test section, oil-water-gas

37

separation section and data acquisition system. A brief description of these components is

presented next.

3.1.1 Storage and Metering Section

As is shown in Figure 3.2, oil and water are stored in separate tanks, each of 400

gallons capacity. Each tank is connected to two pumps that are equipped with return

lines. The first pump is a model 3656, size 1x2-8, of cast iron construction with a bronze

impeller, John Crane Type 21 mechanical seal, and 10 HP motor rotating at 3600 rpm

nominal. It delivers 25 gpm @ 108 psig. The second pump is a model 3656, size 1.5x2-

10, of cast iron construction with a bronze impeller, John Crane Type 21 mechanical seal,

and 25 HP motor rotating at 3600 rpm nominal. It delivers 110 gpm @ 150 psig.

The liquids are pumped from the storage section to the metering section where the

flow rates, densities and temperatures are measured. The metering section is comprised of

pressure transducers, temperature transducers, control valves and state-of-the-art

Micromotion® Coriolis mass flow meters. The water and oil flow rates are controlled

using Fisher control valves mounted in the water and oil lines, respectively. Both the

water and oil pipelines have check valves mounted in the lines downstream of the control

valves to avoid back flow. The flow rates and densities of both water and oil are

measured using the Micromotion® mass flow meters. The oil and water are combined in a

mixing-tee to obtain oil-water mixture. A static mixer, in series with the mixing tee, is

available to promote mixing of the two liquids.

38

Figure 3.1 Experimental Facility

A SULLAIR LS-100 40H compressor with working capacity of 0-1560 scfm (at

up to 125 psig delivery pressure) is used to supply compressed air for operating the

control valves and pressurizing the 3-phase separator. The compressor also supplies the

air to the flow loop. The airflow rate is controlled by a gas control valve and metered

using a Micromotion® mass flow meter. The air and liquid streams are combined at a

mixing tee. Check valves, located downstream of each feeder line, are provided to

prevent back flow.

39

Figure 3.2 Storage and Metering Section

3.1.2 Test Section

Figure 3.3 shows a photo of the HPS© test section. The HPS© body is a 3.75-in.-

ID, 19-ft 8-in. long transparent PVC pipe, built from pipe spools, with 2-inch nominal

inlet and outlet pipes.

Figure 3.3 HPS© Test Section

40

The HPS© is equipped with multiple inlet arrangement, as shown in Figure 3.4.

However, the present study utilized only the inlet concentric with the HPS© body. Figure

3.5 shows the static mixer installed upstream of the inlet to promote mixing. Test runs

with and without the mixer indicated that the use of the mixer isolates the inlet mixture

flow conditions from the “mixing history” of the flow upstream of the mixer, namely,

mixing due to the flow through the mixing tee and piping components previous to the

HPS© inlet.

Figure 3.4 HPS© Inlet Section

41

Figure 3.5 KOMAXTM Static Mixer Spool Upstream of the Inlet Section

a) Fishbone Outlet

b) Straight Tee Outlet c) Vessel Outlet

Figure 3.6 HPS© Outlet Configurations

42

Figure 3.6 displays the three outlet configurations tested in this study, namely, a

fishbone, straight tee and vessel outlets. Manual gate valves at the HPS© outlets control

the split between the oil and water rich stream flow rates.

A differential pressure transducer is installed in the HPS©, connected with an

array of tubes and valves that enables the measurement of the pressure drop along the

separator at various locations. Two metering ports, at 7.5 and 13.5 ft from the inlet, allow

the installation of instrumentation for local velocity profile and video acquisition. Figure

3.7 shows the location of the different pressure taps and local measurement

instrumentation ports.

Figure 3.7 Location of Local Measurement Ports and Pressure Tap Ports

(Lengths in inches)

Pressure Tap Locations

79 ½” 39”

56 ¾”

24 ½”

90” (l/D=24)

72” (l/D=19.2)

HPS Inlet

Oil Rich Outlet

Water Rich Outlet

Local measurement ports location

43

3.1.3 Local Measurement Instrumentation

The following instrumentation/measurement methods were used for acquiring

local measurements:

a) Velocity: Flushed Pitot Tube

b) Water Cut: Isokinetic Sampling

c) Droplet Size Distribution: Borescope and Video Image Proccessing

A brief description of the different methods is given below. Detailed information,

including calibration and operating procedures are included in Appendix AI.

a) Flushed Pitot Tube

A pitot tube was built for the measurement of local oil-water mixture velocity

profiles. The pitot was built using both 3/16-in OD and 5/16-in OD brass tubing. The

pitot tube is connected to a RosemountTM differential pressure transducer, calibrated

between 0 and 1-in of water. It is equipped with a continuous flushing circuit, with two

rotameters for controlling the flushing flowrate. The fluid used for flushing is fed from a

5-gallons oil tank (kept at 40 psia using hydraulic pressurization), through tee

connections at the differential pressure meter ports.

44

The operating principle of the flushed pitot tube is similar to the standard pitot

tube. However, while measuring, the pitot is fed with a flushing fluid that drains through

the pitot ports, flushing the pitot tube internally. This continuous flushing avoids

contamination of the fluid inside the pitot by invasion of fluids from the HPS© flow.

Thus, changes in the measured differential pressure due to gravitational and capillary

effects from the invading non-flushing phase are avoided. The use of a similar

arrangement for air-water flow measurements was reported by Lahey (1987) in water-

continuous flows, while studies with a similar approach for oil-water flow have not been

previously published.

The flushing flowrates were set as a compromise between effective flushing and

possible disturbance of the HPS© flow field. Also, a method is devised for the zero

flushing flowrate calibration, as explained in Appendix AI.

b) Isokinetic Sampling Tube

A sampling tube was built for isokinetic sampling purposes. The tube can be

installed inside the HPS© utilizing the same mechanism used for the pitot tube. The tube

is connected to the inlet of a sampling vessel, initially filled with clean water. This

sampling vessel has a discharge to the atmosphere through a rotameter. The sampling

flowrate is controlled with the rotameter to ensure isokinetic sampling conditions.

45

c) Borescope

A borescope-digital video system was installed inside the HPS©, for obtaining

digital video of the droplets flowing through a specific location. Later, the frames of these

videos were processed for obtaining local droplet size distributions.

3.1.4 Gas-Oil-Water Separation Section

The outlets of the HPS© flow into a downstream three-phase separator. The three-

phase separator operates at 5 psig and has a capacity of 528 gal. It consists of three

compartments: in the first compartment the oil-water mixture is stratified whereby the oil

flows into the second compartment through floatation. In the second compartment, there

is a level control system that activates a control valve discharging the oil into the oil

storage tank. The water flows from the first compartment into the third compartment

through a channel below the second compartment. In this compartment, too, there is a

level control system, allowing the water to flow into the water storage tank. The gas, if

present, is separated in the 3-phase separator and is discharged through a separate outlet

to the atmosphere.

46

3.1.5 Data Acquisition System

Three control valves are mounted in the gas, oil and water metering sections, to

control the inlet gas, oil and water flow rates, respectively. The experimental loop is

equipped with various metering devices, and pressure and temperature transducers. All

output signals from the sensors; transducers, and metering devices are collected at a

central panel.

A "virtual instrument" interface is developed using the LabVIEW® application

program. It integrates measurements, data acquisition, and interactive data processing and

analysis for the feedback control, as well as data and results display. It also provides

accurate and interactive control and display of measured and analyzed variables. The

control of all functions and data acquisition settings is conveniently provided through the

virtual instrument's "front-panel" interface. The LabVIEW® application program provides

variable sampling rates. In this study, the sampling rate was set at 2 or 5 Hz.

A calibration procedure, employing a high-precision pressure pump, is performed

on each pressure transducer at a routine schedule, to guarantee the precision of

measurements.

47

3.1.6 Working Fluids

The working fluids used in this study are tap water and mineral oil (Tulco Tech

80). A small amount of red colored dye (Automatik Red) was added to the mineral oil in

order to improve flow visualization between the phases. Both are marketed by a local

company (Tulco Oils Inc.). Typical properties of the working fluids are shown in Table

3.1 and Table 3.2.

The criteria for selecting this oil are as follows: low emulsification, fast

separation, appropriate optical characteristics, non-degrading properties, and non-

hazardous. The temperature in the flow loop varied between 70º F and 80º F during the

entire experimental investigation.

Table 3.1

Properties of Water-Phase

Density (ρ), @ 70° F 1.0 ± 0.003 g/cm3

Viscosity (μ) @ 70° F 1.25 ± 0.15 cP

48

Table 3.2

Properties of Oil-Phase

Typical Properties ASTM Test Method Tulco Tech 80

Viscosity, SUS

@ 100ºF

@ 200ºF

D2161

85

38

Viscosity, cP

@ 100ºF

@ 200ºF

13.6

2.8

Viscosity, cSt

@ 104ºF

D 445

15.6

Gravity, ºAPI

Specific Gravity @ 60ºF

Pounds/ Gallons

D 287

D1298

33.7

0.8571

7.14

Viscosity Gravity Constant

Flash Point, ºF

Pour Point, ºF

D 2501

D 92

D 97

0.81

365

10

Aniline Point, ºF

Refractive Index @ 68ºF

Molecular Weight

Volatility, 22hrs @ 225ºF, wt%

D 611

D 1218

D 2502

D 972

225

1.469

330

2.0

Distillation, ºF

IBP

95 %

D 2887

535

832

49

3.2 Experimental Test Matrix

The objective of this study is to investigate the developing flow region of

oil/water mixture flow in pipes, such as that occurring in the HPS©, towards forming a

segregated flow condition, for medium to high oil cuts. Thus, the experimental test matrix

was chosen inside the steady-state, stratified flow pattern region. Trallero (1995) model

was used to develop the flow pattern map for the HPS© conditions, as shown in Figure

3.8. Based on the stratified region, and on the performance of the downstream 3-phase

separator and instrumentation, the experimental test matrix was defined as shown in the

figure. As can be seen, eight combinations of vSO and vSW were chosen, corresponding to

water cuts of 10, 30, 50 and 70%, and mixture velocities of vM =0.44 ft/s and vM =0.58

ft/s. Next, the most important data sets are presented.

0.01

0.1

1

10

0.1 1 10v OS [ft/s]

v WS

[ft/s

]

Disp. Oil In Wat.

Disp. Oil/Wat. & Wat.

Disp. Wat. In Oil

Disp. Wat. In Oil & Oil

Disp. Wat. In Oil & Disp. OilIn Wat.Stratified

Stratified & Mixing

10%WC

30%WC

50%WC

70%WC

Figure 3.8 Experimental Test Matrix Shown on Steady State Flow Pattern Map

(after Trallero, 1995)

50

3.2.1 Velocity Profiles at Vertical Plane and Velocity Surfaces.

Figure 3.9 (a,b,c,d) and Fig. 3.10 (a,b,c,d) show the measured velocity profiles at

the vertical plane for both metering stations (7.5 and 13.5 ft from the inlet) for the

experimental test matrix. Figure 3.11 (a,b,c,d) and Fig. 3.12 (a,b,c,d) show the contour

plots of the velocity profiles, for vM=0.44 ft/s, interpolated from the data. Figure 3.13

(a,b,c,d) and Fig. 3.14 (a,b,c,d) show similar contour plots at the same locations but for

vM=0.58 ft/s. As can be seen the velocity profile transforms from nearly parabolic (for

10% WC) to a shear-type profile as the water cut increases, due to the change of

continuous phase (10% is oil-continuous flow, and all other inlet WC are water

continuous flow, as will be shown later) and the lubrication effect of the settling water.

Higher velocities are found at the high water concentration zones at the bottom, while the

oil tends to flow at lower velocity at the top. The difference between the velocity profiles

at the two metering stations is small, and within the experimental error, indicating that the

flow is momentum-developed for the local mixture conditions, except for 10% WC,

where the difference between the profiles is due to water droplets settling.

Although Figs. 3.9 to 3.14 do not show discontinuities in the velocity profile, for

10% WC, the pitot tube measurement method experiences a discontinuity in the

measurement, due to capillary pressure effects in the pitot ports. As this discontinuity is a

measurement-related phenomenon, and not actual hydrodynamic phenomenon, the

discontinuity was eliminated applying both overall and oil-phase mass balances. Still, the

location of the discontinuity is a direct measurement of the height of the

51

a) 10%WC

10%WC, v M=0.44ft/sw/mix, θ=0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft13.5 ft

b) 30%WC

30%WC, v M=0.44 fts/s, w/mix, θ=0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght

[in]

7.5 ft13.5 ft

c) 50%WC

50%WC, v M=0.44 ft/s, w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft13.5 ft

d) 70%WC

70%WC, v M=0.44 ft/s, w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft13.5 ft

Figure 3.9 Velocity Profiles at Vertical

Plane (vM=0.44 ft/s)

a) 10%WC

10%WC v M= 0.58 ft/s w/mixer, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft13.5 ft

b) 30%WC

30%WC, v M=0.58 ft/s w/mix, θ =0º

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft13.5 ft

c) 50%WC

50%WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft13.5 ft

d) 70%WC

70%WC v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft13.5 ft

Figure 3.10 Velocity Profiles at Vertical

Plane (vM=0.58 ft/s)

52

a) 10% WC

b) 30% WC

c) 50% WC

d) 70% WC

Figure 3.11 Velocity Contours at 7.5 ft from Inlet (vM=0.44 ft/s)

a) 10% WC

b) 30% WC

c) 50% WC

d) 70% WC

Figure 3.12 Velocity Contours at 13.5 ft from Inlet (vM=0.44 ft/s)

53

a) 10% WC

b) 30% WC

c) 50% WC

d) 70% WC

Figure 3.13 Velocity Contours

at 7.5 ft from Inlet. (vM=0.58 ft/s)

a) 10% WC

b) 30% WC

c) 50% WC

d) 70% WC

Figure 3.14 Velocity Contours

at 13.5 ft. from Inlet. (vM=0.58 ft/s)

54

oil-continuous water-continuous interface, and can be used in the analysis as the criterion

for the location of the boundary between these two layers

3.2.2 Water Cut Profiles

Figure 3.15 (a,b,c,d) and Figure 3.16 (a,b,c,d) show the water cut profiles

measured at the 7.5 and 13.5 ft metering locations for the experimental matrix conditions.

As can be seen, the water tends to settle at the bottom of the pipe, resulting in a small

difference in the measured water cut between both metering stations. At 10% WC the

concentration profile changes between the metering stations, indicating slow settling of

water droplets. The fact that the water cut is not zero at the oil rich layer at any data set

indicates that the flow is in the developing region. (under steady state flow, the flow

pattern should be stratified, with complete segregation of the phases)

For layer-average water cut estimations, the flow in the HPS© is divided into three

layers. The boundary between the oil-continuous and the water-continuous regions can be

defined through the velocity profile estimation (Sec. 3.2.1). The boundary between the

packed dispersion and the loose dispersion layer is be set at local WC equal to 75% in

this investigation. With this information, the average WC of each layer can be calculated

through numerical integration of the WC and velocity profiles within each layer. Tables

3.3 and 3.4 present the average WC in both the packed dispersion and the loose

dispersion layer. Note that the average water cut of the packed dispersion layer increases

as the overall flowrate increases, and as the inlet water cut increases (for the same overall

flowrate). Also, note that the water cut of the loose dispersion layer is usually 80% or

55

higher, except at very low WC at the inlet, condition that also shows a very small water

layer thickness. From Figure 3.15 and Figure 3.16, the minimum WC at the packed

dispersion layer top is between 15% to 20%.

Table 3.3

Average WC in Packed Dispersion Layer (7.5 ft and 13.5 ft from inlet)

a) Average Water Cut Values, 7.5 ft from inlet

Mix Vel [ft/s] 10%WC 30%WC 50%WC 70%WC 0.44 17±8 27 ±4 21 ±2 27 ±3 0.58 n/a 30 ±2 36 ±4 34 ±9

b) Average Water Cut Values, 13.5 ft from inlet Mix Vel [ft/s] 10%WC 30%WC 50%WC 70%WC

0.44 52 ±25 22 ±6 18 ±2 31 ±3 0.58 n/a 26 ±50 29 ±2 38 ±5

Table 3.4

Average WC in Loose Dispersion Layer (7.5 ft and 13.5 ft from inlet)

a) Average Water Cut Values, 7.5 ft from inlet

Mix Vel [ft/s] 10%WC 30%WC 50%WC 70%WC 0.44 57±14 61±10 81±3 91±3 0.58 n/a 52±2 89±4 91±3

b) Average Water Cut Values, 13.5 ft from inlet Mix Vel [ft/s] 10%WC 30%WC 50%WC 70%WC

0.44 52 ±25 78 ±10 73 ±2 91 ±4 0.58 n/a 62 ±75 81 ±8 90 ±4

56

a) 10%WC

10% WC, v M=0.44 ft/s w/mix , θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft13.5 ft

b) 30%WC

30% WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft13.5 ft

c) 50%WC

50% WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft13.5 ft

d) 70%WC

70%WC, vM=0.44 ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft13.5 ft

Figure 3.15. Water Cut Profiles at

Vertical Plane (vM=0.44 ft/s)

a) 10%WC

10%WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft13.5 ft

b) 30%WC

30% WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft13.5 ft

c) 50%WC

50% WC, v M =0.58 ft/s w/mix θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft13.5 ft

d) 70%WC

70%WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft13.5 ft

Figure 3.16. Water Cut Profiles at the

Vertical Plane. (vM=0.58 ft/s)

57

3.2.3 Layer Height

Based on the layer boundary height definitions given in Sec 3.2.1 and 3.2.2, it is

possible to determine the heights of the different layers in the HPS©. The results are given

in Table 3.5 and Table 3.6, in dimensionless form (i.e., divided by the HPS© diameter).

The changes between heights at different metering locations are within the uncertainty of

the measurement, indicating that the flow might be a developed one. However, from the

previous knowledge that there is mixing of the phases, when the steady-state flow pattern

should be stratified smooth, shows that the flow is in developing conditions.

Table 3.5

Dimensionless (h/D) Height of Packed Dispersion Layer-Loose Dispersion Layer

Boundary (7.5 ft from inlet)

Dimensionless heights, 7.5 ft from inlet

Mix Vel [ft/s] 10%WC 30%WC 50%WC 70%WC 0.44 0.08 ±0.03 0.13 ±0.03 0.20 ±0.03 0.33 ±0.03 0.58 n/a 0.10 ±0.03 0.20 ±0.03 0.34 ±0.03

Table 3.6

Dimensionless (h/D) Height of Packed Dispersion Layer-Loose Dispersion Layer

Boundary (13.5 ft from inlet)

Dimensionless heights, 13.5 ft from inlet

Mix Vel [ft/s] 10%WC 30%WC 50%WC 70%WC 0.44 0.14 ±0.03 0.20 ±0.03 0.26 ±0.03 0.29 ±0.03 0.58 n/a 0.16 ±0.03 0.20 ±0.03 0.32 ±0.03

58

3.2.4 Droplet Size Distribution Profiles

Figure 3.17 and Figure 3.18 present the experimental data utilizing three droplet

size distribution parameters: minimum diameter (dMIN), median diameter (dMED, or d50)

and maximum diameter (dMAX). Figure 3.19 and Figure 3.20 present the change of the

local Sauter Diameter (d32) at the different metering stations. As can be seen, at 10% WC

the droplet diameter increases towards the bottom of the HPS©, indicating oil-continuous

flow conditions. For all other conditions, the droplet size increases towards the top of the

HPS©, indicating water-continuous flow. Note also that the average droplet size is smaller

at vM=0.58 ft/s, as compared to 0.44 ft/s, indicating that the static mixer causes more

mixing and droplet break-up at higher mixture velocities.

The local Sauter diameter increases slightly between the two metering stations

(specially for vM=0.44 ft/s), demonstrating that the flow is not fully developed, as

suggested by the previous figures. Thus, it is clear that the momentum development

length (length required to attain a length-independent local velocity and phase

distribution profiles in a pipe section) is much smaller than the interfacial area

development length (length required for attaining a local interfacial area per unit volume

profile constant at any location of the pipe cross sectional area).

59

a)

10%WC, v M=0.44 ft/s w/mixer, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in] dmin, 7.5 ft

d50, 7.5 ftdmax, 7.5 ftdmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ft

b)

30% WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 7.5 ftd50, 7.5 ftdmax, 7.5 ftdmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ft

c)

50% WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 7.5 ftd50, 7.5 ftdmax, 7.5 ftdmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ft

d)

70% WC, v M=0.44 ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 2000 4000 6000Droplet Diameter [μm]

Hei

ght [

in] dmin, 7.5 ft

d50, 7.5 ftdmax, 7.5 ftdmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ft

Figure 3.17 Droplet Size Distribution Profiles at Vertical Plane (vM=0.44 ft/s)

59

60

a)

10%WC, v M=0.58 ft/s, w/mix, θ=0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 7.5 ftd50, 7.5 ftdmax, 7.5 ftdmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ft

b)

30% WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in] dmin, 7.5 ft

d50, 7.5 ftdmax, 7.5 ftdmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ft

c)

50% WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 7.5 ftd50, 7.5 ftdmax, 7.5 ftdmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ft

d) 70% WC, v M=0.58 ft/s

w/mix, θ =0°0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in] dmin, 7.5 ft

d50, 7.5 ftdmax, 7.5 ftdmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ft

Figure 3.18 Droplet Size Distribution Profiles at Vertical Plane (vM=0.58 ft/s)

60

61

(a) 10%WC

10% WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

d32, 7.5 ft

d32, 13.5 ft

(b) 30%WC

30% WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

d32, 7.5 ftd32, 13.5 ft

(c) 50%WC

50% WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

d32, 7.5 ftd32, 13.5 ft

(d) 70%WC

70% WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

d32, 7.5 ftd32, 13.5 ft

Figure 3.19 d32 Profiles at Vertical Plane (vM=0.44 ft/s)

(a) 10%WC

10% WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

d32, 7.5 ftd32, 13.5 ft

(b) 30%WC

30% WC, v M=0.58 ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

d32, 7.5 ftd32, 13.5 ft

(c) 50%WC

50% WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

d32, 7.5 ftd32, 13.5 ft

(d) 70%WC

70% WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

d32, 7.5 ft

d32, 13.5 ft

Figure 3.20 d32 Profiles at Vertical Plane (vM=0.58 ft/s)

62

Fig. 3.21 presents the overall droplet size distribution of the packed dispersion

layer measured at the two metering stations, namely, 7.5 and 13.5 ft, respectively, for

vM=0.44 ft/s. Fig. 3.22 presents similar results for vM=0.58 ft/s. In both figures, the overall

distributions are shown for 30, 50 and 70% WC. The plots are that of cumulative

frequency expressed as a function of the droplet diameter. The continuous lines represent

fitting of the droplet size distribution using Log-Normal probability curves as given by

Eq. 2.9, where the data is fitted as function of the droplet volume. For Log-Normal

distribution fitting parameter estimations, all lengths are given in meters.

The obtained fitting parameters for the droplet size distributions are shown in

Table 3.7 and Table 3.8 These distributions will be used as indication of the development

of the flow, and the distribution at 7.5 ft will be used as input to the model for the

coalescence process prediction along the HPS©.

From Figure 3.21 and Figure 3.22, it can be observed that the distribution exhibits

small changes at higher mixture velocities and water cuts. Also, the changes of the

droplet size distribution are small, comparable to the instrument and calculation

procedure uncertainty in most of the cases. This indicates that the length of the

experimental facility is not enough to have a measurable change on some of the

experiments. Even, in Figure 3.22 can be observed that the measured droplet size

distributions appears to be smaller at 13.5 ft than at 7.5 ft (even the measurements are still

inside the uncertainty band). This is due to the error of the numerical integration of the

velocity profile at higher velocities and steeper profiles and the small droplet size change.

63

a)

Packed Layer Droplet Size Dist. Development30%WC, v M=0.44 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

Cum

. Fre

quen

cy

Experimental 7.5 ft

Experimental 13.5 ft

Log-Normal 7.5 ft

Log-Normal 13.5 ft

b)

Packed Layer Droplet Size Dist. Development50%WC, v M=0.44 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

Cum

. Fre

quen

cy

Experimental 7.5 ft

Experimental 13.5 ft

Log-Normal 7.5 ft

Log-Normal 13.5 ft

c)

Packed Layer Droplet Size Dist. Development70%WC, v M=0.44 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

Cum

. Fre

quen

cy Experimental 7.5 ft

Experimental 13.5 ft

Log-Normal 7.5 ft

Log-Normal 13.5 ft

Figure 3.21 Overall Droplet Size Distribution in Packed Dispersion Layer (vM=0.44 ft/s)

64

a)

Packed Layer Droplet Size Distrib. Development 30%WC, v M=0.58 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

Cum

. Fre

quen

cy

Experimental 7.5 ft

Experimental 13.5 ft

Log-Normal 7.5 ft

Log-Normal 13.5 ft

b)

Packed Layer Droplet Size Dist. Development50%WC, v M=0.58 ft/s, w/mixer

00.10.20.30.40.50.60.70.80.9

1

0 1000 2000 3000 4000Diameter (μm)

Cum

. Fre

quen

cy

Experimental 7.5 ft

Experimental 13.5 ft

Log-Normal 7.5

Log-Normal 13.5 ft

c)

Packed Layer Droplet Size Dist. Development70%WC, v M=0.58 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

Cum

. Fre

quen

cy Experimental 7.5 ft

Experimental 13.5 ft

Log-Normal 7.5 ft

Log-Normal 13.5 ft

Figure 3.22 Overall Droplet size Distribution in Packed Dispersion Layer (vM=0.58 ft/s)

Packed Layer Droplet Size Dist. Development

65

Table 3.7

Log-Normal Distribution Fitting Parameters for Cumulative Distributions (Eq. 2.9)

in Figure 3.21

Mixture Superficial Velocity= 0.44 ft/s

7.5 ft 13.5 ft

WC μ0 σ0 μ0 σ0

30% -21.07 0.86 -20.635 0.735

50% -21.8 0.85 -21.436 0.84

70% -21.92 0.79 -22.05 1.1

Table 3.8

Log-Normal Distribution Fitting Parameters for Cumulative Distributions (Eq. 2.9)

in Figure 3.22

Mixture Superficial Velocity= 0.58 ft/s

7.5 ft 13.5 ft

WC μ0 σ0 μ0 σ0

30% -22.33 0.84 -22.79 1.0

50% -22.84 0.73 -23.36 1.1

70% -22.95 0.55 -23.05 0.6

66

As a summary of the experimental results presented, the following statements can

be made on the hydrodynamic flow behavior in the HPS©:

1- The observed flow patterns show local mixing of the phases. So the flow patterns

are not reaching stratified smooth flow conditions (as indicated in Figure 3.8),

demonstrating that the flow is under developing conditions.

2- Smaller droplet sizes were seen at higher mixture velocity of vM=0.58 ft/s at 7.5 ft

from the inlet. This shows that the average mixture interfacial area at the pipe

inlet is a function of the flowrate due to the static mixer at the inlet.

3- Dispersed-phase segregation happens few diameters from the inlet when the

mixture was water-continuous (24 diameters or less for the given experimental

conditions). No total dispersed phase segregation was observed for oil-continuous

flow, even at the total experimental facility length (64 diameters).

4- There is little difference between the velocity profiles measured at the two

metering stations (7.5 and 13.5 ft from inlet). However, the droplet size

distributions changed, especially for vM=0.44 ft/s. This confirms that the flow is

not fully developed, and also indicates that the hydraulic development is much

faster than the interfacial area concentration development.

5- The droplet distribution characteristic diameters (minimum, median and max

diameter) show dependence with the height, indicating gravitational segregation

of the droplet population in the HPS© vertical plane.

6- Less change in droplet size distribution between the two metering stations was

observed at higher flowrates, as the residence time decreases.

67

3.2.5 Pressure drop

Figure 3.23 and Figure 3.24 show the experimental results for the pressure drop

along the HPS©, with and without the static mixer at the inlet, respectively (In these plots,

the results shown for 100% oil flow pressure drop corresponds to calculated ones, as was

not possible to measure these parameters experimentally due to the pressure drop

metering system design). The observed flow patterns were quite similar with and without

mixer, but the observed droplet sizes were much bigger without the mixer than with the

mixer installed. Also, the droplets without the mixer were not spherical, but deformed, so

no calculation of their diameter was attempted.

From the figures can be observed that the pressure drop per unit length is much

smaller without mixer (Figure 3.23) as compared to with mixer (Figure 3.24). This leads

to the fact that the pressure drop is smaller under conditions of lower interfacial area

concentration, compared to the ones obtained at higher interfacial area concentration.

For both cases the pressure drop increases (as compared to pure water flow

results) when the water concentration decreases from 100% down to a certain value, and

then markedly decreases, even below pure oil flow predictions. This pressure drop per

unit length reduction was previously reported (i.e., Pal, 1987). The value at which the

maximum pressure drop occurs when no mixer is installed is around 50% WC (Figure

3.23), and decreases when the mixer is installed to around 30% WC (Figure 3.24).

68

When estimating the mixture inversion point WC value by considering that the

phase inversion occurs at the same WC where the maximum pressure drop per unit length

in the pipe flow for a given mixture flowrate is measured, the obtained results are

somehow contradictory: from Figure 3.24 (a), this criterion indicates the flow should be

in oil-continuous conditions below 30% WC. However, from Figure 3.15 (b) the

minimum local water cut found in the upper part of the pipe is lower (around 10% to

15%), and the Figure 3.17 (b) indicates that the flow is water continuous as the droplet

size increases towards the top of the pipe, indicating that the fluid on the dispersed phase

has smaller density than the continuous fluid. So, the estimation of the inversion point

through pressure gradient measurement cannot be applied under developing and/or

segregated flow conditions.

A summary of the pressure drop results is:

1- Pressure drop increases as the interfacial area concentration at inlet increases

2- Pressure drop can be smaller than that of pure phase results.

3- At the inversion point, critical water cut concentration obtained through the

pressure drop analysis can be different from the local measured ones acquired in

this study.

69

a)

0.00.20.40.60.81.01.21.41.61.82.0

0 5 10 15 20Length [ft]

dP [i

n H

2O]

Pure WaterPure Oil90%WC70%WC50%WC30%WC10%WC

v M=0.44 ft/s no mixer

b)

0.00.20.40.60.81.01.21.41.61.82.0

0 5 10 15 20

Length [ft]

dP [i

n H

2O]

Pure WaterPure Oil90%WC70%WC50%WC30%WC10%WC

v M=0.58 ft/s no mixer

Figure 3.23. Pressure Drop Along HPS© for Different Water Cuts, Without Mixer.

c)

v M=0.44 ft/s, w/mixer

0.00.20.40.60.81.01.21.41.61.82.0

0 5 10 15 20Length [ft]

dP[in

ch H

2O]

Pure WaterPure Oil90%WC70%WC50%WC40%WC30%WC25%WC20%WC

d)

v M=0.58 ft/s w/mixer

0.00.20.40.60.81.01.21.41.61.82.0

0 5 10 15 20Length [ft]

dP [i

nch

H2O

]

Pure WaterPure Oil90%WC70%WC50%WC40%WC30% WC20%WC10%WC

Figure 3.24 Pressure Drop Along HPS©, With Mixer

69

dP [i

nch

of H

2O]

dP [i

nch

of H

2O]

dP [i

nch

of H

2O]

dP [i

nch

of H

2O]

70

3.2.6 Outlets Performance

Figure 3.25 and Figure 3.26 show the experimental results for the oil cut at the oil

outlet for the three different outlet designs, as a function of the split ratio between the

outlets, for vM=0.44 and 0.58 ft/s, respectively. The split ratio is defined as the ratio of the

flowrate through the oil outlet to the total inlet flow rate. The measured oil cuts are

compared with the oil cut at the oil outlet at complete (hypothetical) separation, and with

the oil cut that would occur at the oil outlet if the outflow splitting is made through the

use of a hypothetical horizontal plane dividing the pipe cross section into upper and lower

channels at 13.5 ft from the inlet. For the second case, for a given height the outlet

flowrates (and the corresponding split ratio) are obtained by numerical integration of the

experimental velocity and water cut profiles of the cross sectional across in both

channels.

The results show that the experimental oil cut measured at the HPS© oil outlet is

equal or slightly higher than the one obtained from the horizontal plane-velocity profile

integration method, and all are lower than the complete hypothetical separation oil cut

curves. This indicates that the separation efficiency is not a strong function of the

proposed outlet designs in spite of having very different geometries, but rather depends

on the conditions inside the separator, upstream of the outlets.

71

A summary of the results follows:

1- The overall separation is a strong function of the hydrodynamic flow behavior in

the HPS©, and a weak function of the tested outlet configurations.

2- The outlet designs show good efficiency, as the obtained water cut at the oil outlet

results were similar to the expected ones occurring when splitting the flow inside

the HPS© with a horizontal plane.

72

a)

00.10.20.30.40.50.6

0.70.80.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1Split Ratio

Oil

Cut

Integration Upstream Outlets

Complete separation OC

Fishbone Outlet OC

Straight Outlet OC

Vessel Outlet OC

WC=10%v M=0.44 ft/sWith mixer

b)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1Split Ratio

Oil

Cut

Integration Upstream OutletsComplete separation OCFishbone Outlet OCStraight Outlet OCVessel Outlet OC

WC=30%v M=0.44 ft/sWith mixer

c)

0

0.10.20.3

0.40.5

0.6

0.70.8

0.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1Split Ratio

Oil

Cut

Integration Upstream OutletsComplete Separation OCFishbone Outlet OCStraight Outlet OCVessel Outlet OC

WC=50%v M=0.44 ft/sWith mixer

d)

00.10.20.30.40.50.6

0.70.80.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1Split Ratio

Oil

Cut

Integration Upstream Outlets

Complete Separation OC

Fishbone Outlet OC

Straight Outlet OC

Vessel Outlet OC

WC=70%v M=0.44 ft/sWith mixer

Figure 3.25 Oil Cut at Oil Outlet (vM=0.44 ft/s)

72

73

a)

00.10.20.30.40.50.60.70.80.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Split Ratio

Oil

Cut

Integration Upstream OutletsComplete separationFishbone Outlet OCStraight Outlet OCVessel Outlet OC

10%WCv M=0.58 ft/sWith Mixer

b)

00.10.20.30.40.50.60.70.80.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Split Ratio

Oil

Cut

Integration Upstream OutletsComplete separationFishbone Outlet OCStraight Outlet OCVessel Outlet OC

30%WCv M=0.58 ft/sWith Mixer

c)

00.10.20.30.40.50.60.70.80.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Split Ratio

Oil

Cut

Integration Upstream OutletsComplete separation OCFishbone Outlet OCStraight Outlet OCVessel Outlet OC

50%WCv M=0.58 ft/sWith Mixer

d)

00.10.20.30.40.50.60.70.80.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Split Ratio

Oil

Cut

Integration Upstream OutletsComplete separationFishbone Outlet OCStraight Outlet OCVessel Outlet OC

70%WCv M=0.58 ft/sWith Mixer

Figure 3.26 Oil Cut at Oil Outlet (vM=0.58 ft/s)

73

74

CHAPTER 4

MODELING

This chapter presents the model developed for the prediction of the flow behavior

in the developing region of the HPS©. The model comprises of two sub-models, namely,

the hydrodynamic sub-model and the coalescence sub-models; and a set of closure rules

that relate the two sub-models. These are presented in the next sections.

4.1 Hydraulic Sub-Model

4.1.1 Number of layers

Figure 4.1 demonstrates the segregation phenomena occurring in the HPS©. At the

inlet, the fluids might be well mixed and distributed over the entire cross sectional area of

the separator. After some length, gravitational segregation occurs, and a layered flow (up

to three per each continuous-phase at the inlet) develops. These are the fundamental

phenomena that will be associated with the hydraulic sub-model.

75

Figure 4.1 Three Layer Developing Flow. (vM=0.44 ft/s, 50% WC, without Mixer, 14 ft

from Inlet)

The following assumptions are made in the development of the hydraulic sub-

model:

a) A single continuous-phase is assumed at the separator inlet, either oil or water.

This assumption provides a conservative estimate of the developing length.

b) The droplet size distribution of the dispersed-phase at the inlet is assumed to be

known.

c) The mixture flows in the HPS© inlet is a homogeneous mixture.

d) Up to three different layers will develop along the HPS©. Assuming water as the

continuous phase, these layers (from top to bottom of the HPS© cross-sectional

area) are:

d.1) Pure oil, or water in oil emulsion (w/o). Under assumption (a), the initial

flowrate of this layer is zero. The flowrate of this layer increases along the

HPS© owing to mass transfer from the middle layer due to droplet

coalescence. Water might be present in the layer only if a double-dispersion

occurs at the inlet of the HPS© (i.e., water in oil in water). If this condition

76

applies, it is also assumed that the water droplets inside the oil droplets are

small and stable enough to stay homogeneously dispersed in the oil over the

entire HPS© length.

d.2) Oil-in-water (o/w) packed dispersion. This layer is formed through the

creaming of the oil droplets in the oil-in-water dispersion fed into the HPS©.

The initial flowrate of this layer is zero, and the layer flowrate increases along

the HPS© through dispersion settling. After settling, the oil droplets in this

layer coalesce with each other, increasing the average size of the droplet size

distribution in the layer, as the mixture flows along the HPS©. Droplets that

grow above a limiting maximum size are assumed to instantaneously

coalescence with the upper layer interface, feeding oil flow into the upper

layer, and decreasing the flowrate of this layer. The average WC of this layer

can be estimated from experimental data, or (if no data is known) assumed

similar to the one for maximum mono-dispersed sphere packaging, namely,

25% WC. This assumed value is based on the compromise between the

droplet dispersion due to bulk layer flow and mixing, and the droplet packing

due to the low-velocity flow through the HPS© and the improved packaging of

multiple-droplet diameter dispersions.

d.3) Loose oil in water dispersion. This is the initial flow condition at the inlet of

the HPS© (i.e. homogeneous dispersion of oil in water). As the mixture flows

at low velocity along the HPS©, the oil droplets settle, increasing the middle

layer flowrate. As a result, the flowrate of this lower layer decreases, as well

as its oil concentration. At some length from the inlet this lower layer

77

becomes mostly a clean water layer. If this layer flows under turbulent flow

regime, some amount of small diameter oil droplets will stay dispersed due to

turbulence effects. This amount of oil is assumed a function of the droplet size

distribution and the turbulent dissipation energy.

Note that this structure can also be used for the description of oil-continuous -

water-in-oil dispersion flow in the HPS©, by interchanging the continuous and dispersed

phases.

4.1.2 Layer Mixture Properties

For all layers, it is assumed that the layer density can be obtained by applying a

local volumetric average, assuming local no slippage between the phases in the layer. A

set of assumptions for each layer follows:

The following assumptions are made for upper layer:

a) The WC is constant, and can be assumed as zero or a small value (smaller than

the value at the inversion point).

b) The mixture is assumed Newtonian, and the mixture viscosity is estimated using

an effective viscosity model (Brinkmann, 1952 or similar methods).

For the middle layer, the following assumptions apply:

a) A constant water cut is estimated for he entire layer from experimental data. If not

data is known, a value is assumed, i.e WC is 25%. This value is somewhat smaller

78

than the maximum packing of same diameter spheres due to the existing droplet

size distribution.

b) No slippage between phases is considered.

c) The mixture is assumed Newtonian, and the mixture viscosity is estimated using

an effective viscosity model (Brinkmann, 1952; Pal, 1987; or similar methods).

The following assumptions apply for the lower layer:

a) The water cut evolves along the HPS© from the value at the inlet to a maximum

one. This maximum value can be smaller than 100%, if some of the inlet

dispersed-phase has droplets with small diameters that can stay homogeneously

dispersed due to turbulent mixing.

b) All oil droplets flow from the bottom of the pipe, floating upward towards the

dense packed layer-loose packed layer interface. The vertical velocity component

can be obtained from a drag-buoyancy force balance. The axial velocity of the

droplets along the pipe is the same as the continuous-phase velocity (assuming no

slippage in the axial direction). These droplets will be considered large enough for

not being affected by local turbulence.

c) Dilute dispersions are assumed.

d) No turbulent, added mass or Basset forces are considered.

e) Plug velocity profile is assumed in the lower layer.

79

4.1.3 Mathematical Formulation

The previously given assumptions allow the characterization of the mixture flow

in the HPS© as a developing segregated three layer flow. With these assumptions, the

three-layer model of Taitel et al. (1995) originally developed for gas-oil-water flow is

modified for oil-packed dispersion-loose dispersion flow, and applied for the

determination of the two layer boundary heights and the pressure drop estimation. The

HPS© is discretized into segments, and a stepwise calculation method is used, where the

level of each layer is calculated from the segment inlet conditions, based on the settling

and/or coalescence that occur in the current segment. Initially, the height of each layer is

calculated for the segment inlet flowrates; then mass transfer due to buoyancy is

determined, and the height is re-calculated for the new layer flowrates. This is repeated

until convergence is obtained in both height and settling. Next, the coalescence occurring

in the segment is estimated, the mass transfer between layers due to coalescence is

calculated, and the layers heights are recalculated. These new heights and flowrates are

used as inlet conditions for the next HPS© segment. In the next section, the hydraulic sub-

model based on Taitel et al. (1995) three-layer model is presented.

Taitel et at. (1995) 3-layered model

This model is based on a three-layer stratified flow configuration in a pipe. The

model assumes that the flow is in steady-state, one-dimensional and fully developed

conditions. Under these assumptions, the momentum equation for each layer can be

written as (refer to Figure 4.2):

80

Figure 4.2 Force Balances Over Each of the Three Layers. (after Taitel et. al., 1995)

hO

hW

AO (px+dx)

Water layer

Dense Packed layer

Oil layer

AW (px+dx)

AM (px+dx)

τO SO dx

τi,O-M Si,O-M dx

AW (px)

AM (px)

AO (px)

τM SM dx

τi,M-W Si,M-W dx

τi,M-W Si,M-W dx

τW SW dx

ρW AW dx

ρM AM dx

ρO AO dx

g

Direction of flow

τi,O-M Si,O-M dx

Oil Layer:

Dense Packed Layer:

Loose Packed Layer:

81

0sin,, =−+−− −− βρττ WWWMiWMiWWW

W gASSdx

dpA .................................................. (4.1)

0sin,,,, =−+−−− −−−− βρτττ MMMOiMOiWMiWMiMMM

M gASSSdx

dpA ......................... (4.2)

0sin.. =−−−− −− βρττ OOMOiMOiOOO

O gASSdx

dpA ..................................................... (4.3)

The shear stresses at the pipe wall are given by:

2

2WW

WWv

τ = ............................................................................................................. (4.4)

2

2oo

oov

τ = .............................................................................................................. (4.5)

2

2MM

MMv

τ = ........................................................................................................... (4.6)

And the interfacial shear stresses between the layers are:

2)()(,

,,WMWMWMi

WMiWMi

vvvvf

−−= −

−−

ρτ ............................................................... (4.7)

2)()(,

,,MOMOMOi

MOiMOi

vvvvf

−−= −

−−

ρτ .................................................................. (4.8)

82

The friction factor expressions are functions of the rheology of the fluid in the

layer. Assuming Newtonian flow, the Blasius correlation can be used for the friction

factor calculations, as suggested by Taitel et. al.(1995), namely:

no Cf −= Re ...................................................................................................................(4.9)

Where C=16, n=1 for laminar flow; and C=0.046, n=0.2 for turbulent flow.

The geometrical parameters can be expressed as functions of the two non-

dimensional layer heights, Dhh OO =~ and Dhh WW =

~ , as follows:

[ ]212

)1~2(1)1~2()1~2(cos4

)~( −−−+−= −WWWWW hhhDhA ........................................(4.10)

[ ]212

)1~2(1)1~2()1~2(cos4

)~( −−−+−= −OOOOO hhhDhA .......................................... (4.11)

⎟⎟⎠

⎞⎜⎜⎝

⎛−−= )~()~(

4)~,~(

2

OOWWOWM hAhADhhA π .................................................................(4.12)

)1~2(cos)~( 1 −= −WWW hDhS .........................................................................................(4.13)

)1~2(cos)~( 1 −= −OOO hDhS ..........................................................................................(4.14)

)~()~()~,~( OOWWOWM hShSDhhS −−= π .........................................................................(4.15)

2, )1~2(1)~( −−=− WWWMi hDhS ...................................................................................(4.16)

83

2, )1~2(1)~( −−=− OOMOi hDhS ....................................................................................(4.17)

These equations are solved using the procedure suggested by Taitel et.al. (1995),

which is described next. Adding the Eqs. 4.1 and 4.2, and assuming that pW=pM=p, the

following expression is obtained:

0sin,, =⎟⎟⎠

⎞⎜⎜⎝

⎛++

−+

+⎟⎟⎠

⎞⎜⎜⎝

⎛++

−− −− βρρτττ

gAA

AAAA

SAA

SSdxdp

MW

MMWW

MW

MOiMOi

MW

MMWW ................... (4.18)

Now, assuming also pO=p, the pressure gradient from Equations 4.3 and 4.18 can

be eliminated, resulting in the following equation:

0sin11,, =⎟⎟

⎞⎜⎜⎝

⎛−

++

−⎟⎟⎠

⎞⎜⎜⎝

⎛+

++

+⎟⎟⎠

⎞⎜⎜⎝

⎛++

−− βρρρ

τ

τττ

gAA

AAAAA

S

AS

AASS

OMW

MMWW

OMWMOiMOi

O

OO

MW

MMWW

...................... (4.19)

From the previous assumptions, the pressure gradient from equations 4.1 and 4.3

can be eliminated, resulting in:

0sin)(,,,, =−−++

+−

−−−− βρρττ

ττ

gA

SAS

AS

AS

OWW

WMiWMi

O

MOiMOi

O

OO

W

WW

............................................... (4.20)

84

Substituting Eqs. 4.4 through 4.17 into Eqs. 4.19 and 4.20, these equations

become functions of the layer flowrates, pipe geometry, fluids physical properties, and

the non-dimensional heights Dhh OO =~ and Dhh WW =

~ (Taitel & Dukler (1976),

Taitel, et. al. 1995 and Shoham (2006)). As only the non-dimensional heights are

unknowns, the following solution procedure applies for solving for these heights:

a) Assume as a guess a small value of Oh~

b) For this assumed value, solve for Wh~ from Eq. 4.20. Only one solution of Wh~ will

be found for horizontal or downward flow.

c) With these values of Oh~ and Wh~ , Eq. 4.19 is checked. If not satisfied, a higher

value of Oh~ is used and b) is repeated. Bisection numerical method can be used

for obtaining the roots.

d) After finding the Oh~ and Wh~ pair that are roots of Eq. 4.19, looking for additional

roots is required, assuming higher values of Oh~ until reaching the top of the pipe.

As indicated by Taitel et. al. (1995), multiple roots can be found for upward flow,

and following the same approach given by the author, the pair of Oh~ , Wh~ that

shows the smallest Wh~ is considered to be the stable one.

85

4.2 Coalescence Sub-Model

4.2.1 Physical Phenomena

As the dispersion settles, the droplets make contact with their neighbors. The film

of the continuous-phase between the droplets thins out and for long enough contact times

the droplets coalesce into larger ones.

From Figs. 3.9 and 3.10, it can be seen that the velocity profile of the packed

dispersion layer in the vertical plane can be assumed as linear, increasing towards the

bottom of the pipe for water-continuous flow conditions. Thus, it is assumed that the

packed layer behaves like a simple shear flow (actually, at least in the HPS© vertical

plane this is valid) along the separator axis. Note that from Figs. 3.15 and 3.16 it can be

seen that the water cuts can also be assumed, as a first approximation, linear functions of

the height.

For simple shear flow, Friedlander (2000), reports the Smoluchowsky (1915)

model for droplet collision for simple shear flow conditions. This model considers the

number of collisions per unit time as a function of the velocity gradient, the local phase

concentration, and the sizes of the droplets in the control volume. As shown

schematically in Fig. 4.3, droplet “j” collides with droplet “i” due to the velocity gradient,

droplet diameters and relative location.

86

Figure 4.3 Coalescence in Simple Shear Flow due to Velocity Gradient

The local distributions of droplet sizes are measured in this study in the vertical

plane only. Thus the variation of droplet sizes at different angles from the vertical is

unknown. However, due to the low velocities used in this study, and also due to the

gravity settling, it is reasonable to assume that the droplet size distribution is not a strong

function of the angle from the vertical plane, but rather of the height of the layer inside

the HPS©. Thus, it is assumed that the droplet size distribution measured at the vertical

plane is similar along the chord of the cross-section, at the same height.

Figures 3.17, through 3.20 show that there is some degree of segregation of the

droplet sizes in the vertical plane of the HPS©. Hence, selective coalescence occurs, as a

droplet of a given size might not be able to coalesce with all other droplets of different

sizes due to a concentration gradient occurring in the HPS© section. Note that the

concentration of the largest droplets is expected to be high near the oil-packed dispersion

boundary. However, it is not expected that the large droplets are surrounded by a high

di

dj

h

U(h)

87

concentration of small droplets, as the small droplets are more able to disperse in the

entire packed layer cross-sectional area. Thus, the proposed model must take into account

the effect of the stratified segregation of the droplets in the packed layer, as will be

discussed next.

4.2.2 Assumptions

The following assumptions are considered for modeling the coalescence process

in the HPS©:

a) Coalescence occurs only in the middle (packed) layer.

b) Only coalescence is taken into account in the prediction of the droplet size

distribution (no break-up is considered).

c) The coalescence rate is a function of the number of droplet collisions per unit

volume per unit time, multiplied by a coalescence efficiency. This approach has

been considered in previous studies (i.e., Prince and Blanch, 1990 and Chesters,

1991)

d) Laminar flow conditions are assumed (simple shear flow) in the middle layer.

e) A model is used for estimation of the overall droplet size distribution segregation

in the packed dispersion layer cross sectional area of the HPS©.

f) For coalescing contact time estimation purposes, the following assumptions

apply:

f.1) For estimation of the linear velocity profile in the packed layer, the velocity at

the top of the packed layer is assumed to be the same as the average velocity

88

of the upper layer. The velocity at the bottom of the packed layer is calculated

from a mass balance on the layer. This assumption allows the use of the

Smoluchowski’s equation (Friedlander, 2000) for predicting impact frequency

in shear flow.

f.2) Linear concentration profile in the packed layer is assumed. The minimum

water cut occurs at the top of the packed layer, which is assumed to be 15%,

an average value of the ones given by the experimental data on this

investigation (Chapter 3). The maximum water cut occurs at the bottom of the

packed layer, which can be calculated from a mass balance over the layer.

4.2.3 Mathematical Formulation.

Estimation of the number of collisions per unit volume per unit time.

The flow is assumed to be stratified, and all properties are assumed as functions

of the height only. Thus, the coalescence is a function of the number of droplets, their

diameter as well as the shear rate (assuming simple shear flow), as given in Eq. 2.2:

⎟⎠⎞

⎜⎝⎛+=

dhdvddCnCnN jijiji

3, )(

61 ................................................................................(4.21)

where, as defined previously, Ni,j is the number of collisions per unit time per unit

volume of droplets i and j, Cn is the respective droplet concentration per unit volume, d

is the diameter of the droplets, and dv/dh is the shear rate of the continuous-phase.

89

Estimation of the number of coalescing collisions per unit volume per unit time.

The number of collisions that lead to coalescence is given by the total number of

collisions times an efficiency function (Colaloglou and Tavlarides, 1977) that relates the

actual contact time jiCt ,− and the required time jiDt ,− for drainage of the film between the

droplets, namely:

)exp( ,,, jiCjiDji tt −−−=λ ..............................................................................................(4.22)

For shear flow, the contact time can be estimated as the inverse of the shear rate:

⎟⎟⎟

⎜⎜⎜

⎛=−

dhdv

t jiC1

, ............................................................................................................(4.23)

The required time for drainage is a function of the time required for inertial

drainage of the film due to collision (function of inertia), and viscous drainage (function

of film and dispersed-phase viscosities; also affected by interfacial additives and

impurities) (Oolman and Blanch, 1986). Chesters (1991) proposed this expression for

semi-mobile interfaces:

90

⎟⎟⎠

⎞⎜⎜⎝

⎛−

⎟⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜⎜

⎟⎟⎠

⎞⎜⎜⎝

⎛=−

IF

ji

jiDjiD hh

d

Ft 11

425.1

,

5.0,

,

πσ

πμ..................................................................................(4.24)

For this expression the following assumptions are made:

a) The contact force between the droplets is proportional to the deformation of the

droplets. Assuming the proportionality constant to be equal to one:

σπ jiji dF ,, = .................................................................................................................(4.25)

b) For droplets of different sizes, an equivalent diameter can be used, as follows

(Prince and Blanch, 1990):

1

,11

21

⎟⎟⎠

⎞⎜⎜⎝

⎛+=

jiji dd

d ....................................................................................................(4.26)

In these equations, the initial distance between droplet surfaces where film-

thinning hydrodynamics become important (hI), and the distance where interface

instabilities produces interface breakage and coalescence (hF) are needed. From the

literature (Oolman and Blanch, 1986 and Prince and Blanch, 1990) a typical value of hI is

between 10-3 and 10-4 m., while for hF (Chesters and Hofman, 1982, Prince and Blanch,

1990, Chesters, 1991) it is of the order of 10-8 m.

91

Note that these criteria are local, and for applying them, the following information

is needed: (1) Shear rate, (2) Local WC, (3) Local dispersion flowrate and (4) Local

droplet size distribution.

So, a set of closure rules that relate the hydraulic and the coalescence sub-models

are required to define these pseudo-local flow conditions. These rules are described in the

next section.

4.3 Closure Rules

This section presents the set of closure rules that relates the hydraulic and

coalescence sub-models. For the coalescence closure rules, as coalescence is assumed to

occur only in the packed layer, the rules apply to the packed layer only, unless otherwise

is stated.

4.3.1 Estimation of the Settling Velocity

The settling velocity of the droplets in the dispersed flow layer is estimated by

applying a simple force balance between drag and gravity forces. For a droplet of a given

diameter d, these forces are:

( ) 3

6dgF DCB

πρρ −= ................................................................................................(4.27)

92

22

8vdCF CDRDR ρπ

= ..................................................................................................(4.28)

and:

Re2424

== CC

DR vdC μ

ρ................................................................................................(4.29)

Equating FB=FDR, and using Eq. 4.29, results in:

( )DCC

Sgdv ρρ

μ−=

18

2

..................................................................................................(4.30)

This Stokes terminal velocity is assumed as the velocity in the vertical direction.

No slip is assumed in the axial direction.

4.3.2 Estimation of the Velocity and Water Cut Profiles in the Packed Layer

Determination of the local velocity and WC gradients requires an estimation of

the slopes of the velocity and water cut profiles. In this study it is assumed that the

velocity and water cut profiles are two-dimensional, and change linearly with the height.

Under these conditions, it is possible to analytically calculate the mixture and water

flowrates through the pipe cross-sectional area by integration. The resultant expressions

can be used for estimating the velocity and the WC slopes. This is developed next.

The cross-sectional area of a circle can be determined as:

93

hdhDhdSAD

i~)1~2(1~ 1

0

22

0∫∫ −−== ...........................................................................(4.31)

Assuming the velocity and water cut profiles in the packed dispersion layer as

linear, as shown schematically in Figure 4.4, one obtains:

)~~()~~)(~(

)~( ,WO

OMOi hh

hhhddvvhv

−+= − .................................................................................(4.32)

and

)~~()~~()~(

)~( ,WO

OMOi hh

hhhddWCWChWC

−+= − ....................................................................(4.33)

Figure 4.4 Schematic of Proposed Velocity and WC Profiles

in the Packed Dispersion Layer

h~h~Oh~

Wh~

( )hv ~ ( )hWC ~MOiv −, MOiWC −,

94

The overall flowrate of the packed layer through its cross-sectional area can be

determined by integration of the velocity profile (Eq. 4.32), as follows:

hdhD

hhhhhddv

v

hdhDhvQO

W

O

W

h

hOW

O

MOih

h

~ )1~2(1~~

)~~)(~(~ )1~2(1)~(

~

~

22

,~

~

22 ∫∫⎟⎟⎟⎟

⎜⎜⎜⎜

−−⎟⎟⎟⎟

⎜⎜⎜⎜

+

=⎟⎠⎞⎜

⎝⎛ −−=

....(4.34)

Also, the overall flowrate of the packed layer can be estimated from the average

velocity and the pipe cross-sectional area of flow, namely:

( ) O

W

h

hhhhDvQ

~

~212 )1~2(1)1~2()1~2(cos

41

−−−+−−= −π ........................................(4.35)

Equating Eqs. 4.34 and 4.35, the following expression is obtained

O

W

O

W

h

h

h

h OW

OMOi

hhhDv

hdhhh

hhhddvvD

~

~

212

~

~

2,

2

)1~2(1)1~2()1~2(cos41

~ )1~2(1)~~(

)~~)(~(

⎟⎠⎞⎜

⎝⎛ −−−+−−

=⎟⎟⎠

⎞⎜⎜⎝

⎛−−⎟

⎟⎠

⎞⎜⎜⎝

−+

−∫

π

..............................................(4.36)

From this expression, it is possible to obtain the velocity slope that preserves the

volumetric balance of the layer.

95

Similar expressions for the estimation of the water flowrate through the pipe

cross-sectional area can also be obtained. Thus, the water flowrate through a cross-

section delimited by the layer boundaries can be estimated as:

hdhDhvQO

W

h

hWW

~))1~2(1)~((~

~

22∫ −−= ...........................................................................(4.37)

As no local slippage between the phases is considered, the mixture and water-

phase velocities are related through the Water Cut. Thus, Eq. 4.37 changes to:

∫⎟⎟⎟⎟

⎜⎜⎜⎜

−−⎟⎟⎟⎟

⎜⎜⎜⎜

−+⎟⎟

⎟⎟

⎜⎜⎜⎜

−+

=−−

O

W

h

hOW

O

MOi

WO

O

MOi

W hdhD

hhhhhddv

v

hhhhhddWC

WC

Q~

~

22

,,~)1~2(1

~~)~~)(~(

~~)~~)(~( ..........(4.38)

The water flowrate on the packed dispersion layer also can be obtained from the

average water velocity on the packed layer as:

( ) O

W

h

hW hhhDWCvQ

~

~212 )1~2(1)1~2()1~2(cos

41

−−−+−−= −π ................................(4.39)

Solving for the water flowrate in Eqs. 4.38 and 4.39

96

( )

O

W

O

W

h

h

h

hOW

H

MOi

OW

O

MOi

hhhDWCv

hdhD

hhhhhddv

v

hhhhhddWC

WC

~

~

212

~

~

22

,,

)1~2(1)1~2()1~2(cos41

~ )1~2(1~~

)~~(~~~

)~~)(~(

⎟⎠⎞⎜

⎝⎛ −−−+−−

=⎟⎟⎟⎟

⎜⎜⎜⎜

−−⎟⎟⎟⎟

⎜⎜⎜⎜

−−

−⎟⎟⎟⎟

⎜⎜⎜⎜

−+

−−

π

..................(4.40)

From this expression, it is possible to obtain a value for the WC slope that

preserves the phase mass balance.

Eqs. 4.36 and 4.40 have analytical solutions. After integration and solving for the

slopes (details are given in Appendix IV), the following results are obtained:

The velocity slope is given by:

O

W

O

W

O

W

h

h

h

h

h

h

F

FF

hddv

~

~

~

~

~

~

1

23~

−= ....................................................................................................(4.41)

where:

⎟⎟⎟⎟

⎜⎜⎜⎜

−−⎟⎠⎞

⎜⎝⎛ −

⎟⎟⎟⎟

⎜⎜⎜⎜

−+

−−

−= 2

32

21

22

)~~(31~

21

)1~2arcsin(81

)~~)(1~2(41

)~~(21 hhh

h

hhh

hhDF O

OW

............................ (4.42)

97

⎟⎟⎟⎟

⎜⎜⎜⎜

⎟⎟⎟⎟

⎜⎜⎜⎜

−+

−−= Hv

h

hhhDF

)1~2arcsin(41

)~~)(1~2(21

221

2

2 .........................................................................(4.43)

vh

hhhDF

⎟⎟⎟⎟

⎜⎜⎜⎜

−+

−−=

)1~2arcsin(41

)~~)(1~2(21

321

2

2 ................................................................................(4.44)

The water cut slope can be solved from the following equation:

O

W

O

W

O

W

h

h

h

h

h

h

G

GG

hddWC

~

~

~

~

~

~

1

23~

−= ............................................................................................... (4.45)

where:

⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜

⎟⎟⎠

⎞⎜⎜⎝

−−⎟⎟

⎞⎜⎜⎝

⎛−

−⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

−⎟⎠⎞

⎜⎝⎛ −+

−+

−−

⎟⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜⎜

⎟⎟⎠

⎞⎜⎜⎝

−⎟⎠⎞

⎜⎝⎛ −−+

⎟⎟⎠

⎞⎜⎜⎝

−⎟⎠⎞

⎜⎝⎛ −

+

⎟⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜⎜

⎟⎟⎠

⎞⎜⎜⎝

−⎟⎠⎞

⎜⎝⎛ +−+

⎟⎟⎠

⎞⎜⎜⎝

−⎟⎠⎞

⎜⎝⎛ −

=

223

2

23

2,

21

2

22

,

22

,

2

)~~(

~)~~(

2

~

)~~(~~~

85~2

)~~(32

)~~)(~21(

)~~(

~~21

325~

21

~~~

21

21

)1~2arcsin(

)~~(

~

645~

41~

41

~~21~

41

1

WO

WOOMOi

WO

WOOO

WO

MOiO

WOOO

WO

MOiO

hhhddvhhh

hhhhhddvhv

hh

hhh

hhhddvhh

hhv

h

h

hhhddvhh

hhv

h

DG .......................(4.46)

98

⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜

−−

+

⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜

−⎟⎟⎠

⎞⎜⎜⎝

⎛−−

+

+−−

⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜

−⎟⎟⎠

⎞⎜⎜⎝

⎛−+

=

23

2,

,

,

21

2

,

,

2

)~~()~()~~(3

2

)1~2arcsin(

)~~(

~

4

~

81

41

)~~)(~21(

)~~(

~

2

~

41

21

2

hhhddvWChh

h

hhhddvWC

h

WCv

hhh

hhhddvWC

h

WCv

DG

MOiWO

WOMOi

O

MOiO

WOMOi

O

MOiO

................................(4.47)

)()1~2arcsin(

41

)~~)(1~2(21

321

2

2 WCvh

hhhDG

⎟⎟⎟⎟

⎜⎜⎜⎜

−+

−−= .......................................................................(4.48)

From these expressions, values of water cut and velocity slopes are calculated,

and the two-dimensional velocity and Water Cut profiles can be built. These are needed

for calculating the local droplet concentration and coalescence rates in Eq. 4.21 and 4.23,

respectively, as functions of the height.

4.3.3 Estimation of the Local Droplet Size Distribution in the Packed Layer

From Figs 3.17 and 3.18, it is evident that there is a change in the concentration of

droplets of the same diameter as a function of the height in the HPS©. Figures 4.5 and 4.6

99

show the experimental cumulative frequency of the population of specific droplet size

intervals traveling inside the HPS© as a function of the height, for different diameter

intervals at the given experimental conditions.

70%WC, v M =0.44 ft/s w/mixer

at 7.5 ft from inlet

0

0.2

0.4

0.6

0.8

1

0.00 0.75 1.50 2.25 3.00 3.75Height from HPS bottom [in]

F 900-1000 microns

1500-1600 microns

2000-2100 microns

Figure 4.5 Cumulative Frequency of Selected Droplet Diameters as Function of Height

Inside the HPS© (70%WC, vM=0.44 ft/s, w/mixer, 7.5 ft from the inlet)

70%WC, v M =0.58 ft/s, w/mixer,

7.5 ft from the inlet

0

0.2

0.4

0.6

0.8

1

0.00 0.75 1.50 2.25 3.00 3.75Height from the HPS bottom [in]

F

600-700 microns900-1000 microns1500-1600 microns2000-2100 microns

Figure 4.6 Cumulative Frequency of Selected Droplet Diameters as Function of Height

Inside the HPS© (70%WC, vM=0.58 ft/s, w/mixer, 7.5 ft from the inlet)

100

The following phenomena can be observed from the experimental data:

1- Droplets of a given size do not travel at the same height, but are dispersed

through a height range.

2- When analyzing the height where larger droplets of a given distribution

travel along the pipe, the median diameters of the droplet distribution

corresponding to these droplets appear to occur at a higher location from

the bottom than that of the smaller ones.

The hydraulic sub-model gives no information on the local droplet distribution in

the packed layer. Thus, an alternative procedure is implemented to distribute the droplets

along the height of the packed layer. The description of this procedure is given in the

following paragraphs.

Based on the phenomenon described earlier, an ad-hoc expression is presented for

determination of the local droplet size distribution probability frequency as a function of

the population droplet volumes flowing in the packed dispersion layer, where the

population is arranged in volume bins. (A volume bin is defined as the volume range that

is used to quantify the frequency (or the actual number) of droplets with a volume

between the lower and upper limits of the bin). This expression includes some fitting

parameters that account for the effect of the local lift and turbulence dispersion on the

packed droplets, not calculable from the hydraulic sub-model.

101

It is expected that droplets of bigger diameters tend to stay at the top of the pipe,

while smaller ones tend to be sparsely distributed, or stay near the bottom, when they

cannot fit between the larger ones. Thus, this model considers that a given droplet mixes

with the others following a distribution based on the droplet volumes, as follows:

))(ln(

)))((ln())(ln())(ln(

exp2

,j

i

ij

iiji dV

dVbdVdV

cnn⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛ −−

= ...............................................................(4.49)

This expression gives the number of droplets from the diameter j bin that travel

together with droplets of the diameter i bin per unit time. The dispersion of the droplet

population of a given diameter around the other diameters is governed by the parameter

b, while c is a scaling factor, which is a function of each droplet diameter, and is used to

ensure that the maximum cumulative frequency of the distribution of droplets of a

diameter j in all other droplet bins is equal to one.

Note that the values of b are bounded between two values that represent two

specific mixing conditions:

102

b = 0: Completely segregated flow: Each droplet size travels separately from the

others.

b →∞: Competely mixed flow. The droplet size distribution is the same at any

location in the packed dispersion zone of a given pipe area section.

The parameter ci can be calculated using the following expression:

∑=

⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛ −−

=

nb

j j

i

ij

i

dV

dVbdVdV

c

1

2

))(ln(

))(ln())(ln())(ln(

exp

1 ..................................................................(4.50)

where nb is the total number of bins of the droplet size distribution discretization.

After defining how the droplets mix with each other, a volumetric flow balance is

carried out, assuming that the droplets flow as sub-layers in the packed-dispersion layer.

There are as many sub-layers as bins of the population discretization. The droplet

population flowing through each of this sub-layer is obtained by adding the droplets

contained in a given bin as estimated from Eq. 4.49, and assigning these to the

corresponding layer. Then, through an overall oil volumetric balance on the packed layer

it is possible to calculate the height from the bottom of the pipe of all sub-layer.

103

The procedure of defining the local distribution from the overall distribution when

experimental data are available follows (refer to Fig. 4.7, where the number of bins used

is nb =5):

a) Check that the packed layer has already a discrete droplet size distribution, or at

least has a continuous overall droplet size distribution, as given in Eq. 2.9. The

Log-Normal distribution (Eq. 2.9) was found to better fit the experimental data.

b) If the initial overall distribution in the packed layer is not discrete, it is discretized

into a given number of droplet volume bins (nb). The range of each volume bin is

calculated as 1/nb times the volume of a 5000 microns sphere. Then, the number

of droplets per unit time of each bin can be estimated through the following

equation:

∫∫

+

+=

)(

)()(

;)(),(

1

1)(

0)(

i

i

MAXii

dV

dVdV

PDLOdVdV dVVF

dVVVF

Qn .......................................................................(4.51)

where n is the number of droplets per unit time flowing in the packed dispersion

layer in the bin defined between the volumes V(di) and V(di+1).

c) Having the overall droplet size distribution discretized into bins, an initial value

of b is assumed (near to zero).

d) ci are calculated for all bins using Eq. 4.50.

104

1) Steps (a) and (b)

2) Steps (c) to (f)

3) Step (g)

Figure 4.7 Procedure of Assignment of Local Droplet Size Distributions as Function of

Height (nb = 5)

V(d)

F

j sub-layers

i bins

hj

hO HPS Q(V5)

Q(V4) Q(Vj)

Eq. 4.51

105

e) A matrix of nb x nb is built, where the results of Eq 4.49 are fed into the columns.

f) The rows of this matrix are assumed to be the local distribution of the packed sub-

layer j. The oil volumetric flow of each j sub-layer is found by adding the volume

of the droplets per unit time given in the row, using the following equation:

)(;, )( jOi

iji QdVn =∑ ....................................................................................................(4.52)

g) A mass balance is carried out on the packed dispersion layer cross-section, from

the largest bin (nb-th bin) to the i-th bin, looking for the height in the HPS© which

transports an oil mass flow equal to the summation of oil flow from the i-th bin to

the nb-th bin (as given in Eq. 4.53). At this height, local velocity and water cut are

recorded, and are to be used for residence time and coalescence calculation

purposes.

)(;)1(;)(;

~

~

22 ...~)1~2(1))~(1)(~( jOnbOnbO

h

h

QQQhdhDhWChvO

j

+++=−−− −∫ .......................(4.53)

h) The obtained local droplet size distribution as function of the height is compared

with the experimental results.

i) The value of b is incremented and steps (d) through (h) are repeated, until

properly fitting of the experimental data (i.e. applying least square method to fit

some of the experimental data: (1) fitting the model distribution to a particular set

106

of diameters, from the information given by figures like Figs. 4.5, 4.6; or (2)

fitting the basic droplet diameter information of the distribution (dMIN, dMEDIAN and

dMAX)).

Evolution of the Local Distribution Parameters Along the Separator.

The value of b is a function of the flowing conditions in the HPS©, and it might

change as the coalescence of the dispersed phase occurs. However, as this value is

expected to be stronger function of the overall flowrates through the HPS© than of the

phase distribution, the present model will assume b constant for the whole coalescing

process for the given inlet conditions.

4.3.4 Coalescence Estimation Procedure

Now it is possible to relate the flowing characteristics with the predictions of Eq.

4.21. For a given sub-layer k, the flowrate, the water cut, the velocity, and the inlet

distribution can be calculated from the previous sections. From this information, the

concentration of droplets i per unit volume can be estimated as:

)1(,

,,

kkO

kiki WCQ

nCn

−= ...................................................................................(4.54)

Next, Eq. 4.21 can be evaluated. The number of coalescences can be determined

from the following expression:

107

( ) ⎟⎟⎠

⎞⎜⎜⎝

⎛ −

⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛+⎟⎟

⎞⎜⎜⎝

⎛−

= jiD

jiCt

jijikMk

kOk e

dhdvddCnCn

vdx

WCQ

NC ,

,

3

,

,

61

)1(τ ..................................(4.55)

where NC is the number of coalescences of droplets of diameters di and dj in the k-th

sublayer, dx is the length of the HPS© segment, kMvdx , is the residence time of the fluids

in this segment and )1(, kkO WCQ − is the sub-layer mixture flowrate.

After estimating the coalescence in all sub-layers, the results are applied to the

overall distribution and the process is repeated (generation of sub-layers, coalescence,

and application of results to the overall distribution) for all time steps in a segment, and

for all segments until the HPS© outlet is reached.

108

4.4 Calculation Procedure

A step-by-step calculation procedure is given below. Refer to Fig. 4.8 for the

corresponding flowchart.

1- The HPS© is divided into a number of segments (given by the user).

2- The number of bins for the droplet size distribution is chosen. The given overall

droplet size distribution at inlet is discretized with this bin distribution (Eq. 4.51),

using the overall oil flow at the inlet.

3- The flow in the HPS© is assumed homogeneously dispersed (oil in water, all of it

flowing in the loose dispersion layer).

4- Initially, all the oil droplets are assumed to be located at the bottom of the HPS©. This

gives a conservative approach for the required settling time.

5- For the given HPS© segment, the hydraulic sub-model is run to define inlet layer

thickness and layer velocities (Eqs. 4.19 and 4.20).

6- For the calculated layer velocity, the residence time in the segment is calculated for

all bins in the loose dispersion layer, and the distance traveled by each bin from the

bottom of the HPS© is determined.

7- The vertical distance traveled by each bin from the inlet to the outlet of the segment is

compared with the height of the boundary between the dispersed and packed layers. If

the distance traveled is larger than the height of the boundary, settling is assumed to

occur for that given bin. The bin mass is then transferred from the loose dispersion

109

layer to the packed dispersion layer, including some amount of water, to comply with

the water cut assumption for the packed layer. Flowrates in the packed and loose

dispersion layers are modified to reflect this migration.

8- Step 5 and 7 are repeated until convergence is reached.

9- After convergence, the coalescing sub-model is initiated: shear rate (Eq. 4.41) and

water cut slope (Eq. 4.45) of packed dispersion layer are determined

10- Sub layers are generated and their droplet size composition, flowrates and heights are

determined (Eqs. 4.49, 4.52, 4.53 and 4.54)

11- Coalescence is calculated in each sub-layer (Eq. 4.55).

12- If droplets resulting from collisions are bigger than a critical size (5000 microns) they

are assumed to coalesce with the oil layer. Oil is added to the oil layer, and water (due

to the WC in the packed layer) to the loose dispersion layer.

13- After coalescence, layer level and velocity are recalculated (Eqs. 4.19 and 4.20,

respectively) to determine the layer readjustment due to collisions.

14- The process is repeated from step 5 for all other HPS© segments until reaching the

HPS© discharge.

15- The procedure is repeated from step 1, dividing the HPS© in a larger amount of

segments, until discretization convergence is attained.

110

Figure 4.8 Calculation Procedure Flowchart

Legend: In gray boxes: Initial data and discretization; in thick boxes: sub-models and closure relationships; in yellow boxes: models; in thin boxes: results from models.

Initial data: - Fluids properties - - QO - QM - QW - Pipe Geometry - Overall Droplet Size Distribution

Geometry Discretization

Bin Discretization

Layer Level Estimation:

3-layer Taitel et. al.,1995 Model

- Oil Layer WC - Packed Layer WC- Mixture Models:

a) Density b) Viscosity

- Interface Shear Stress Model

Settling Model

- Settling velocity model

hO hW UO UM UW

QM QW

Coalescence

Model (Packed Layer)

- Local Velocity Model- Local WC Model - Local Droplet Size Distribution Model - Local concentration

QO, QM, QW Drop Size Distribution in Packed Layer

Is last segment calculated?

No

Yes

Stop

Next Segment

Info to first segment

111

CHAPTER 5

RESULTS AND DISCUSSION

This chapter presents comparisons between the developed model predictions and

the experimental data. Comparisons are presented on: (1) layer height evolution along

the HPS©, (2) Velocity and Water Cut Profiles, and (3) Droplet size distribution changes

in the packed layer along the separator.

5.1 Comparison of Layer Height Evolution

The boundaries between the three layers in the HPS© are measured in the

experimental program and reported in Tables 3.5 and 3.6. The two measured boundaries

are the oil layer/packed dispersion layer boundary and the packed dispersion layer/loose

dispersion layer boundary. The developed model is capable of predicting the evolution of

the two boundary heights along the HPS©. Figure 5.1 and Figure 5.2 present comparisons

between model predictions and experimental data for the layer heights evolution along

the HPS© for mixture velocities of 0.44 ft/s and 0.58 ft/s, respectively. In each figure the

sub-figures (a), (b), and (c) corresponds to water cuts of 30%, 50% and 70%,

respectively.

112

As can be seen in Figure 5.1 and Figure 5.2, the dimensionless height of the oil

layer/packed dispersion layer boundary is almost 1. This corresponds to a very thin oil

layer at the top of the pipe. This thin layer is the result of a very small amount of clean

oil, which is given as an input to the developed model and is required for numerical

convergence. This thin layer was also observed in the experimental runs. As can be

seen, the calculated thickness of the oil layer is constant, demonstrating that for these

flow conditions there is no oil mass transfer between the packed dispersion layer and the

oil layer.

The predictions of the evolution of the packed dispersion layer/loose dispersion

layer boundaries exhibit a sharp change at the HPS© inlet, decreasing from dimensionless

height of almost 1 to a steady-state value, which depends on the operating conditions.

This indicates fast settling of the loose dispersion phase (oil) to the packed dispersion

layer. As a result, the height of the packed dispersion layer/loose dispersion layer

decreases, as the thickness of the packed layer increases and the height of the dispersed

layer decreases. The comparison between the model predictions and the experimental

data shows a good agreement at lower water cuts of 30% (sub-figure (a)) and a fair

agreement for middle water cuts of 50% (sub-figure (b)). Finally, for high water cuts of

70%, the model overpredicts the layer height by more than 70%.

113

a)

Layer Heights Comparison30%WC v M =0.44 ft/s, w/mixer, θ =0º

0

0.2

0.4

0.6

0.8

1

0 10 20Length [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/LooseDispersion Boundary

Experimental Oil/PackedDispersion

Experimental PackedDispersion/LooseDispersion

b)

Layer Heights Comparison50%WC v M =0.44 ft/s, w/mixer, θ =0º

0

0.2

0.4

0.6

0.8

1

0 10 20Length [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/LooseDispersion Boundary

Experimental Oil/PackedDispersion

Experimental PackedDispersion/LooseDispersion

c)

Layer Heights Comparison70%WC v M =0.44 ft/s, w/mixer, θ =0º

0

0.2

0.4

0.6

0.8

1

0 10 20Length [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/LooseDispersion Boundary

Experimental Oil/PackedDispersion

Experimental PackedDispersion/LooseDispersion

Figure 5.1. Comparison of Model Predictions and Experimental Data for Layer Heights Evolution (vM=0.44 ft/s)

113

114

a)

Layer Heights Comparison30%WC v M=0.58 ft/s, w/mixer, θ =0º

0

0.2

0.4

0.6

0.8

1

0 10 20Length [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/LooseDispersion Boundary

Experimental Oil/PackedDispersion

Experimental PackedDispersion/LooseDispersion

b)

Layer Heights Comparison50%WC v M =0.58 ft/s, w/mixer, θ =0º

0

0.2

0.4

0.6

0.8

1

0 10 20Length [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/LooseDispersion Boundary

Experimental Oil/PackedDispersion

Experimental PackedDispersion/Loose Dispersion

c)

Layer Heights Comparison70%WC v M =0.58 ft/s, w/mixer, θ =0º

0

0.2

0.4

0.6

0.8

1

0 10 20Length [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/LooseDispersion Boundary

Experimental Oil-PackedDispersion

Experimental PackedDispersion-Loose Dispersion

Figure 5.2. Comparison of Model Predictions and Experimental Data for Layer Heights Evolution (vM=0.58 ft/s)

114

Length [ft]

115

5.2 Comparison of Velocity and Water Cut Profiles

The velocity and water cut profiles were measured in this study at the vertical plane

in the two metering stations located at 7.5 and 13.5 ft from the inlet. The results are

presented in Figures 3.9 through 3.16 in sections 3.2.1 and 3.2.2.

The comparison between model predictions and experimental data for both velocity

and water cut profiles, for water cuts of 30%, 50% and 70% are given, respectively, in

Figure 5.3, Figure 5.4 and Figure 5.5. It can be noted that the developed model assumes

linear velocity profile in the packed dispersion layer, and bulk (average velocity) for the

oil and loose dispersion layers, as presented in section 4.3.2. The model gives near

identical results for the two metering stations, as shown by the solid line in the figures.

The following observations can be seen from these figures:

• The velocity and water cut profiles predicted by the model follow the

experimental trend.

• The velocity profile in the packed layer predicted by the model

underpredicts the experimental data. This could be due to the assumption of

a two-dimensional velocity profile in the packed dispersion layer.

• The velocity in the loose dispersion layer is also underpredicted. The results

shift away from the experimental data at higher water cuts. This could be

due to the overprediction of the height of the packed-dispersion-layer/loose-

dispersion-layer boundary (see Section 5.1)

116

a.1)

30%WC, v M=0.44 fts/s, w/mix, θ=0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght

[in]

7.5 ft

13.5 ft

Model (BothLocations)

a.2)

30%WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in] 7.5 ft

13.5 ftModel (both locations)

b.1)

30%WC v M=0.58 ft/s w/mix, θ=0 deg

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in] 7.5 ft

13.5 ft

Model (Bothlocations)

b.2)

30%WC, vM=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in] 7.5 ft

13.5 ft

Model (bothlocations)

Figure 5.3 Comparison of Model Predictions and Experimental Data for 30% WC. Mixture Velocities vM=0.44 and 0.58 ft/s

116

117

a.1)

50%WC, v M=0.44 ft/s, w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

13.5 ft

Model (bothlocations)

a.2)

50%WC, v M=0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in] 7.5 ft

13.5 ftModel (both locations)

b.1)

50%WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in] 7.5 ft

13.5 ft

Model (bothlocations)

b.2)

50%WC, vM =0.58 ft/s w/mix θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft

13.5 ft

Model (bothlocations)

Figure 5.4 Comparison of Model Predictions and Experimental Data for 50% WC. Mixture Velocities vM=0.44 and 0.58 ft/s

117

118

a.1)

70%WC, v M=0.44 ft/s, w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in] 7.5 ft

13.5 ft

Model (bothlocations)

a.2)

70%WC, vM=0.44 ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft

13.5 ft

Model (bothlocations)

b.1)

70%WC v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

13.5 ft

Model (bothlocations)

b.2)

70%WC, vM=0.58 ft/s w/mix, θ=0°

0.00

0.75

1.50

2.25

3.00

3.75

0 20 40 60 80 100Water Cut

Hei

ght [

in]

7.5 ft

13.5 ft

Model (bothHeights)

Figure 5.5 Comparison of Model Predictions and Experimental Data for 70% WC. Mix. Velocities vM=0.44 (a) and 0.58 (b) ft/s

118

119

5.3 Comparison of Droplet Size Distribution Evolution in Packed Layer

The overall droplet size distribution in the packed layer was measured in the

metering stations located 7.5 and 13.5 ft from the inlet. The results are plotted in Figures

3.17 and 3.18. The developed model is capable of tracking the evolution of the droplet

size distribution along the HPS©, considering settling and coalescence processes. For

comparison purposes, the droplet size distribution data measured at the 7.5 ft metering

station was used as an input to the model. The model was then run along the HPS© and

the resulting droplet size distribution at the 13.5 ft metering station was predicted. Thus,

comparison between the predicted and measured droplet size distributions at the 13.5 ft

metering station location could be carried out.

Figure 5.6 and Figure 5.7 present comparisons between the model predictions and

the measured data for the droplet size distribution at the 13.5 ft metering station for

mixture velocities of 0.44 ft/s and 0.58 ft/s, respectively. Examining the comparisons in

Figure 5.6 and Figure 5.7, it can be observed that the model predictions follow the trend

of the data fairly well. At high flow rates and water cuts, no significant coalescence

occurs due to small residence time, and the droplet size distributions in the two metering

stations are similar. For low flowrates and low water cuts, there is an increase in the size

of the droplets due to coalescence, and the droplet size distribution in the 13.5 ft metering

station shifts toward larger droplet diameters. Again, the model predictions follow this

trend and predict the experimental results fairly accurately. From Figure 5.6 (a) it is

possible to observe that the model predicts the formation of a smaller amount of large

diameter droplets than the measured for these flow conditions.

120

a)

Packed Layer Droplet Size Dist. Development30%WC, v M=0.44 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

F

13.5 ft (model)

Exp. Freq. at 7.5 ft

Exp. Freq. at 13.5 ft

b)

Packed Layer Droplet Size Dist. Development 50%WC v M =0.44 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

F

13.5 ft (model)

Exp. Freq at 7.5 ft

Exp Freq at 13.5 ft

c)

Packed Layer Droplet Size Dist. Development70%WC, v M =0.44 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

F

13.5 ft (model)

Exp. Freq. at 7.5 ft

Exp. Freq. at 13.5 ft

Figure 5.6. Comparison of Model Predictions and Experimental Data for Droplet Size

Distribution Change between Metering Stations (vM=0.44 ft/s)

121

a)

Packed Layer Droplet Size Dist. Development 30%WC, v M=0.58 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

F

13.5 ft (model)

Exp. Freq. at 7.5 ft

Exp. Freq. at 13.5 ft

b)

Packed Layer Droplet Size Dist. Development50%WC v M=0.58 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

F

13.5 ft (model)

Exp. Freq. at 7.5 ft

Exp. Freq. at 13.5 ft

c)

Packed Layer Droplet Size Dist. Development70%WC v M =0.58 ft/s, w/mixer

0

0.2

0.4

0.6

0.8

1

0 1000 2000 3000 4000Diameter (μm)

F

13.5 ft (model)

Exp. Freq. at 7.5 ft

Exp. Freq. at 13.5 ft

Figure 5.7 Comparison of Model Predictions and Experimental Data for Droplet Size

Distribution Change between Metering Stations (vM=0.58 ft/s)

122

The coalescence time used in the proposed model is the semi-mobile interface

coalescence time, jiSMDt ,, − , proposed by Chesters (1991). This coalescence time was

increased by a factor of 40 in order to better match the data, as an indication of presence

of salts, surfactants or other contaminants in the fluids. The fact that the coalescing time

is higher than the one predicted by the semi-mobile interface coalescence model suggest

that the interface might behave more like a immobile interface, so a comparison between

the results obtained by both models under typical experimental conditions follows.

Chesters (1991) proposed expressions for both the semi-mobile interface

coalescence time ( jiSMDt ,, − ) and the immobile deformable coalescence time ( jiIMDt ,, − ).

When the ratio between these expressions is calculated, the following expression is

obtained:

D

C

D

c

jiSMD

jiIMD

hd

hd

dh

d

tt

μμ

πσσππμ

πσπσμ

121

42

)(

1)16(

)12(

5.1

5.0

22

3

,,

,, =

⎟⎠⎞

⎜⎝⎛

=−

− ..........................................................................(5.1)

; being h the height of the film between coalescing droplets (Refer to section 4.2.3).

Assuming 810−=h m (minimum distance before coalescence), an average droplet size of

1000 microns, and a viscosity ratio of 26=CD μμ (the corresponding to the

experimental fluids) the resulting value is 4

,,

,, 105×=−

jiSMD

jiIMD

tt

.

123

Again, in the present model the semi-mobile interface coalescence time jiSMDt ,, −

was increased by a factor of 40. From the ratio between the semi-mobile and immobile

coalescing times, estimated in the previous paragraph, is evident that this correction

factor is still small enough for considering the interface as semi-mobile, and not

immobile. The reason for this correction factor is be the probability of salts present in the

continuous-phase, surfactants or contamination in the fluids. As previously published

(Prince and Blanch, 1990), the presence of salts in the water-phase can delay the thinning

of the interface due to concentration gradients.

5.4 Comparison of Droplet Size Distribution as a Function of the Height in Packed

Layer

Figure 5.8 (a), (b) and (c) and Fig. 5.9 (a), (b) and (c) present a comparison of the

model predictions and the experimental values of the three parameters of the droplet size

distribution shown in Figs. 3.17 (a), (b) and (c) and 3.18 (a), (b), (c) for water continuous

flows. Figures 5.10 (a), (b) and (c) and 5.11 (a), (b) and (c) show the same variables at

13.5 ft from inlet. Also, Table 5.1 shows the value of the parameter b used to adjust the

distributions at 7.5 ft from the inlet.

As can be seen, the model follows the trend of the experimental data, but shows

overprediction of the parameters, indicating less mixing of the droplets with smaller

diameters, as compared to the experimental measured ones. The prediction improves as

the number of bins of the discretization is increased, but higher number of bins also

124

increases considerably the computational time for solving the flow evolution along the

HPS©. The results presented corresponds to a number of bins of nb=750. Considering that

the model takes into account the diffusion only as a function of the droplet volumes and

the flowrates, the results are encouraging.

Table 5.1

Values of b used to adjust the local droplet size distributions in model.

30% 50% 70%

vM=0.44 ft/s 0.035 0.035 0.040

vM=0.58 ft/s 0.035 0.040 0.040

125

a)

30% WC, v M =0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 7.5 ftd50, 7.5 ftdmax, 7.5 ftdmin Modeld50 Modeldmax Model

b)

50% WC, v M =0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 6.5 ftd50, 6.5 ftdmax, 6.5 ftdmin Modeld50 Modeldmax Model

c)

70% WC, v M =0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000 6000Droplet Diameter [μm]

Hei

ght [

in] dmin, 7.5 ft

d50, 7.5 ftdmax, 7.5 ftdmin Modeld50 Modeldmax Model

Figure 5.8 Comparison of Adjusted Model Predictions and Experimental Droplet Size Distributions at 7.5 ft from Inlet

vM=0.44 ft/s

125

126

a)

30%WC, v M =0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 7.5 ftd50, 7.5 ftdmax, 7.5 ftdmin Modeld50 Modeldmax Model

b)

50% WC, v M =0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 7.5 ftd50, 7.5 ftdmax, 7.5 ftdmin Modeld50 Modeldmax Model

c)

70% WC, v M =0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 7.5 ftd50, 7.5 ftdmax, 7.5 ftdmin Modeld50 Modeldmax Model

Figure 5.9 Comparison of Adjusted Model Predictions and Experimental Droplet Size Distributions at 7.5 ft from Inlet

vM=0.58 ft/s

126

127

a)

30% WC, v M =0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in] dmin, 13.5 ft

d50, 13.5 ftdmax, 13.5 ftdmin Modeld50 Modeldmax Model

b)

50% WC, v M =0.44 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in] dmin, 13.5 ft

d50, 13.5 ftdmax, 13.5 ftdmin Modeld50 Modeldmax Model

c)

70% WC, v M =0.44 ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 2000 4000 6000Droplet Diameter [μm]

Hei

ght [

in] dmin, 13.5 ft

d50, 13.5 ftdmax, 13.5 ftdmin Modeld50 Modeldmax Model

Figure 5.10 Comparison of Model and Experimental Droplet Size Distributions at 13.5 ft from Inlet, vM=0.44 ft/s

127

128

a)

30% WC, v M =0.58 ft/s w/mix, θ =0°

0

0.75

1.5

2.25

3

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in]

dmin, 13.5 ftd50, 13.5 ftdmax, 13.5 ftdmin Modeld50 Modeldmax Model

b)

50% WC, v M =0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [mm]

Hei

ght [

in] dmin, 13.5 ft

d50, 13.5 ftdmax, 13.5 ftdmin Modeld50 Modedmax Model

c)

70% WC, v M =0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0 1000 2000 3000 4000 5000Droplet Diameter [μm]

Hei

ght [

in] dmin, 13.5 ft

d50, 13.5 ftdmax, 13.5 ftdmin Modeld50 Modeldmax Mode

Figure 5.11 Comparison of Model and Experimental Droplet Size Distributions at 13.5 ft from Inlet, vM=0.58 ft/s

128

129

CHAPTER 6

CONCLUSIONS AND RECOMMENDATIONS

The following conclusions can be reached from this study:

1. An experimental HPS© facility has been designed and constructed to enable

measurements of the following local parameters in the developing region of oil-

water flow, at two separated cross sections: local velocity profiles (measured

using an in-house developed, continuous flushed pitot tube); water cut profiles

(measured using a isokinetic sampling system); and local droplet size distribution

(measured using a borescope/video processing technique).

2. Experimental data have been acquired for mixture velocities of 0.44 and 0.58 ft/s,

and water cuts of 10, 30, 50 and 70%. All the data were acquired at two metering

stations, located at 7.5 ft and 13.5 ft from the inlet, respectively. For each run, the

local velocity profiles were measured along the horizontal and vertical diameters,

as well as in two additional positions (30º and 60º from the vertical). The water

cut profile for each run was measured only along the vertical diameter. For each

run the droplet size distributions were also measured along the vertical diameter.

3. The data were acquired for a concentric inlet with and without a mixer. Also,

three different outlet configurations were used: straight tee, vessel and fishbone.

For each outlet configuration, the water cut in both the oil and water outlet was

130

measured as a function of the split ratio. The separation results of the different

outlet designs show similar efficiency, and to the one occurring if the flow

upstream of the outlets is split (at the given split ratio) with a horizontal plane (or

a wedge). This indicates that the HPS© overall separation efficiency is a strong

function of the hydrodynamic flow behavior, and a weak function of the tested

outlet configurations.

4. The flow did not reach fully developed flow conditions at any experimental

conditions.

5. The measured velocity profiles varied from nearly parabolic (for 10% WC) to a

shear-type profile, as the water cut increased. The higher velocities are found at

the high water concentration zones, while the oil tends to settle and flow at low

velocity at the top. The difference in the velocity profiles between the two

metering stations is small, and within the experimental error, indicating that the

flow is momentum-developed for the experiments, except for 10% WC, where the

difference between the profiles is caused by slow settling of water droplets.

6. For the 30%, 50% and 70% water cut experiments; water tends to quickly flow

towards the bottom of the pipe, resulting in a small difference in the measured

water cut between both metering stations. This also indicates small diffusion-

developing flow length conditions. However, for 10% water cut, the changes of

the concentration profile between the metering stations indicates slow settling of

water droplets, and longer diffusion-developing length.

7. Smaller average droplet size distributions are measured at the first metering port,

at the higher mixture velocity of vM=0.58 ft/s. This shows that the average mixture

131

interfacial area at the pipe inlet is a function of the flowrate through the static

mixer.

8. The measured droplet diameter profiles as a function of the height are steep,

indicating gravitational segregation of the droplets in the HPS© vertical plane.

Also, the change of the average droplet diameters as a function of the height

demonstrate that the flow is oil continuous for water cut of 10%, and water

continuous for all other water cuts.

9. The droplet size distributions changed between the metering locations, especially

for vM=0.44 ft/s. This indicates that the diffusion-developing flow length is

smaller than the interfacial area concentration developing length. Smaller changes

in the droplet size distribution between the two metering stations were observed at

higher flowrates, as the residence time decreases.

10. A model is developed for prediction of the flow evolution in the developing

region of the HPS©. The model assumes one-dimensional flow of three layers

(hydrodynamic sub-model), and population balance coalescence theory

(coalescence sub-model). The three layers are, from top to bottom (for water-

continuous flow at the inlet): pure oil, packed dispersion of oil in water and loose

dispersion of oil in water. For oil-continuous flow, the three layers are loose

dispersion of water in oil, packed dispersion of water in oil and clear water. This

approach is justified by the fact that the diffusion-developing flow length is much

smaller than the interfacial area concentration development length. The results of

the model match fairly the experimental data, with respect to layer height

development, velocity and water cut profiles and overall droplet size distributions.

132

The following recommendations are made for future studies:

1. Alternative instrumentation needs to be developed for the measurement of local

parameters, especially for droplet size distribution.

2. The investigation should be extended to longer pipe lengths to better study the

development towards steady-state flow. Lower oil viscosities are also

recommended for dealing with shorter development lengths.

3. The flushed pitot tube measurement methods shall be developed further due to its

ruggedness, low cost, and applicability for velocity measurement in non-optical

transparent media, with any droplet size distribution and concentration. This

method needs to be applied together with water cut measurement method (i.e.

isokinetic sampling).

4. The development of a closure relationship for the water cut in the packed layer as

a function of the flow conditions is required.

5. A further study of the rheology of the packed layer and its impact on one-

dimensional fluid flow models in segregated two-phase flows is required for

improving the level estimations of the model.

6. Further research on the segregation of the different droplet diameters as function

of the height in the packed dispersion layer as function of the flowing parameters

is needed.

133

7. Study the oil-water entry region phenomena using two or three-dimensional

approaches with simplified area-concentration models for the dispersed phase

(Kataoka, Ishii and Serizawa, 1986) is recommended.

8. Analyze the distribution of oil-water flows in parallel pipe sections through

manifolds, to study the characteristics of the natural split of the overall flow

between the different tubes.

9. Data acquisition for model validation on more real fluid conditions (i.e. crude and

salty water)

10. Extend the model for accounting for break-up, as this phenomenon might be more

important as the pipe diameter increases, and the flows become more turbulent.

134

NOMENCLATURE

A = area, L2

a = coefficient in Rosin-Rammler cumulative frequency distribution (-)

b = degree of dispersion coefficient (-)

BR = systematic uncertainty (-)

C = coefficient (-) (friction factor, drag)

Cn = Concentration of droplets per unit volume, 1/L3

c = normalizing coefficient (-)

D = pipe diameter, L

d = droplet diameter, L

dx = length of HPS segment, L

dP = differential pressure, M/Lt2

F = Cumulative frequency (-); Force ML/t2

F1, F2, F3 = Functions for calculating shear rate (Eq.4.41 to 4.44)

G1, G2, G3= Functions for calculating Water Cut Slope (Eq. 4.45 to 4.48)

f = probability density function (-); friction factor (-)

g = acceleration due to gravity, L/t2

h = height, L; film thickness, L

l = length, L

M = mass, M

N = Number of collisions per unit time, 1/t

NC = Number of coalescences per unit time, 1/t

n = number of droplets per unit time , 1/t; coefficient (-) (Friction

Factor)

nb = number of bins of distribution discretization

p = pressure, M/Lt2

135

Q = flow rate, L3/t

Re = Reynolds number (-)

S = perimeter, L;

RXS

, = Uncertainty from statistical analysis

T = temperature, T

t = time, t

t95 = Student’s t

U = Uncertainty Value of the variable in brackets

V = Droplet Volume, L3; Sampling Vessel Volume, L3

v = velocity, L/t

x = flow direction coordinate, L

PΔ = Differential pressure reading from pitot tube, M/Lt2

Greek Letters

β = HPS© inclination angle, degrees

δ = coefficient, Rosin-Rammler fitting parameter (-)

ε = turbulent dissipation energy, L2/t3

φ = diameter, L

λ = coalescence efficiency (-)

μ = viscosity, M/Lt, lbm/ft×s; average for Normal Distribution (-)

0μ = average for Log Normal Distribution (-)

π = 3.1415926…

ρ = density, M/L3

σ = surface tension, M/t2, standard déviation, normal distribution (-)

0σ = standard déviation, Log-Normal distribution (-)

τ = shear stress, M/Lt2

θ = inclination angle from vertical, positive counterclockwise, degrees

136

Subscripts

B = Buoyant

C = Contact time; continuous phase

D = Drainage time; dispersed phase

DR = drag

d = droplet

F = final

FL = flushing

I = initial

IM = immobile interface

i = index; interfacial

j = index

k = index

M = mixture

MAX = maximum

MED = median

MIN = minimum

O = oil

PDL = Packed Dispersion Layer

S = Stokes

SM = semi-mobile interface

OS = superficial oil

TOT = total

W = water

WS = superficial water

x = generic location along the pipe

x+dx = location along the pipe separated dx from x

32 = Sauter Mean Diameter

50 = Median Droplet Diameter

137

Abbreviations

ID = Internal Pipe Diameter

OD = External Pipe Diameter

WC = Water Cut

Symbols

~ = Non dimensional (height)

= Average

' = Turbulent (Velocity)

138

REFERENCES

1. Angeli, P. “Liquid Liquid Dispersed Flows in Horizontal Pipes”. PhD

Dissertation, The University of London (1996)

2. Angeli, P.; Hewitt, G. F. “Pressure Gradient in Horizontal Liquid-Liquid Flows”.

Int. J. Multiphase Flow, 24, pp. 1183-1203 (1998)

3. Angeli, P.; Hewitt, G. F. “Drop Size Distributions in Horizontal Oil-Water

Dispersed Flows”. Chem. Eng. Science, 55. pp. 3133-3143 (2000 a)

4. Angeli, P.; Hewitt, G. F. “Flow Structure in Horizontal Oil-Water Flow”. Int. J.

Multiphase Flow, 26, pp. 1117-1140 (2000 b)

5. Arirachakaran, S.; Oglesby, K.; Shoham, O.; Brill, J. P. “An Analysis of Oil-

Water Flow Phenomena in Horizontal Pipes”. SPE 18836, presented at the SPE

Production Operations Symposium in Oklahoma City, OK (1989)

6. Baker, O. "Designing for Simultaneous Flow of Oil and Gas", Oil and Gas J., 53,

pp. 185 (1954)

7. Brauner, N.; Moalen Maron, D. “Two Phase Liquid-Liquid Stratified Flow”.

PhysicoChem. Hydrodinam., 11, pp. 487-506 (1989)

8. Brauner, N.; Moalem Maron, D. “Stability Analysis of Stratified Liquid-Liquid

Flow”. Int. J. Multiphase Flow, 18, pp. 103-121 (1992 a)

139

9. Brauner, N.; Moalem Maron, D. “Flow Pattern Transitions in Two Phase Liquid-

Liquid Flow in Horizontal Tubes”. Int. J. Multiphase Flow, 18, pp. 123-140 (1992

b)

10. Brinkman, H. C. "The Viscosity of Concentrated Suspensions and Solutions". J.

Chem. Phys., 20, pp. 571 (1952)

11. Clark, A. F.; Shapiro, A. U.S. Patent 2533878 (1949)

12. Charles, M.E., Govier, G.W., Hodgson, G.W. "The Horizontal Pipeline Flow of

Equal Density Oil-Water Mixtures". Can. J. Chem. Eng., 39, pp. 27-36 (1961)

13. Charles, M. E.; Redberger, P. J. “The Reduction of Pressure Gradient in Oil

Pipelines by the Addition of Water: Numerical Analysis of Stratified Flow”. ".

Can. J. Chem. Eng., 40, pp. 70-75 (1962)

14. Chesters, A. K.; Hofman. G. “Bubble Coalescence in Pure Liquids”. Appl. Sci.

Res., 38, pp. 353-361 (1982)

15. Chesters, A. K. “The Modelling of Coalescence Processes in Fluid-Liquid

Dispersions. A Review of Current Understanding”. Trans. IChemE, 69, Part A,

pp. 259-270 (1991)

16. Crowe, C.; Sommerfeld, M.; Tsuji, Y. “Multiphase Flows with Droplets and

Particles”. First Edition, CRC Press LLC, ISBN 0-8493-9469-4 (1998)

17. Coulaloglou, C. A., and Tavlarides, L. L., “Description of Interactions Processes

in Agitated Liquid-Liquid Dispersions”. Chem. Eng. Sci., 32, pp. 1289-1297

(1977)

140

18. Elseth, Geir. “ An Experimental Study of Oil/Water Flow in Horizontal Pipes”.

PhD Dissertation, Norwegian University of Science and Technology (NTNU)

(2001)

19. Farrar. B.; Samways, A. L.; Bruun, H. H. “A Computer-based Hot Film

Technique for Two-Phase Flow Measurements”. Meas. Sci. Technol. 6, pp 1528-

1537 (1995)

20. Farrar, B.; Bruun, H. H. “A Computer Based Hot-Film Technique Used For Flow

Measurements in a Vertical Kerosene-Water Pipe Flow”. Int. J. Multiphase Flow,

22, pp. 733-751 (1996)

21. Flanigan, D. A.; Stolhand, J.E.; Scribner, M.E. Shimoda, E. “Droplet Size

Analisys: A New Tool for Improving Oilfield Separation”. SPE 18204, presented

at the 63rd annual Technical Conference and Exhibition of the Society of

Petroleum Engineers held in Houston TX (1988)

22. Friedlander, Sheldon K. “Smoke, Dust and Haze: Fundamentals of Aerosol

Dynamics”. Second Edition, Oxford University Press, ISBN 0-19-512999-7

(2000)

23. Guzhov, A. I., Grishiti, A. D., Medvedev, V, F., Medvedeva, O. P. "Emulsion

Formation During the Flow of Two Liquids” (in Russian). NeftKhoz, 8, pp. 58-61

(1973)

24. Guzhov, A. L, Medvedev, V. F. & Savelev, V. A. "Movement of Gas-Water-Oil

Mixtures Through Pipelines". International Chemical Engineering, 14, pp. 713-

714 (1974)

141

25. Haheim, S. A. “Oil/Water Slip in Inclined Pipes and Application to Downhole

Separation”. Presented at MULTIPHASE 01, 10th Int. Conference, Cannes (2001)

26. Hamidullin, F. F., Hamidullin, M. F., Hamidullin, R. F. “Methods for Phase Pre-

Separation of High Viscosity, Gas-Water-Oil Emulsions”. Ex-Soviet Union Patent

SU 1809911 A3 (1993, in Russian)

27. Hewitt, G. F., Pan, L. & Khor, S. H. "Three-Phase Gas-Liquid-Liquid Flow: Flow

Patterns, Holdups and Pressure Drop", presented at the International Symposium

On Multiphase Flow (ISMF'97), at Beijing, China (1997)

28. Hinze, J. O., “Fundamentals of the Hydrodynamic Mechanism of Splitting in

Dispersion Processes”. AIChE J.,1, pp. 289-295 (1955)

29. Ishii, M. “Thermo-Fluid Dynamic Theory of Two-Phase Flow”, Eyrolles (1975)

30. Karabelas, A. J., “Vertical Distribution of Dilute Suspensions in Turbulent Pipe

Flow”. AlChE J., 23, pp. 426-434 (1977)

31. Kataoka, I.; Ishii, M.; Serizawa, A. “Local Formulation and Measurements of

Interfacial Area Concentration in Two-Phase Flow”. Int. J. Multiphase Flow, 12,

pp. 505-529 (1986)

32. Khor, S. H.; Mendes-Tatsis, M. A.; Hewitt, G. F. “Application of Isokinetic

Sampling Technique in Stratified Multiphase Flows”. Proceedings of the ASME

Heat Transfer Division, HTD-Vol 334 (3), pp 101-107 (1996)

33. Khor, S. H. “Three-Phase Liquid-Liquid-Gas Stratified Flow in Pipelines”. PhD

Dissertation, The University of London (1998)

142

34. Kurban, A. P. A.; Angeli, P. A.; Mendes-Tatsis, M. A.; Hewitt, G. F. “Stratified

and Dispersed Oil Water Flow in Horizontal Pipes”. MULTIPHASE 95, 7th Int.

Conference, Cannes (1995)

35. Kurban, A. P. A. “Stratified Liquid-Liquid Flow”. PhD Dissertation, The

University of London (1997)

36. Lahey Jr., R. T. “Phase Distribution in a Triangular Duct”. Multiphase Science

and Technology, Vol 3, Hemisphere Press, pp 184-231 (1987)

37. Lovick, J. “Horizontal, Oil-Water Flows in the Dual Continuous Flow Regime”

PhD Dissertation, The University of London (2004)

38. Nadler, M; Mewes, D. “Flow Induced Emulsification in the Flow of Two

Inmiscible Liquids in Horizontal Pipes”. Int. J. Multiphase Flow, 23, No. 1, pp

55-68 (1997)

39. Oolman, T. O.; Blanch, H. W. “Bubble Coalescence in Stagnant Liquids”. Chem.

Eng. Commun., 43, pp. 237-261 (1986)

40. Oglesby. K.D. "An Experimental Study on the Effects of Oil Viscosity, Mixture

Velocity and Water Fraction on Horizontal Oil-Water Flow". M.Sc. Thesis, The

University of Tulsa, Tulsa, OK (1979)

41. Pal R. “Emulsions: Pipeline Flow Behaviour, Viscosity Equations and Flow

Measurements”. Ph.D. Dissertation, University of Waterloo, Ontario (1987)

42. Pal R., Rhodes E., “Emulsion Flow in Pipelines”. Int. J. of Multiphase Flow, 15,

No. 6, pp. 1011-1017 (1989)

43. Pal, R. (1993) " Pipeline Flow of Unstable and Surfactant-Stabilized Emulsions".

AIChE. J., 39, 1754-1764.

143

44. Prince, M. J.; Blanch, H. W. “Bubble Coalescence and Break-Up in Air-Sparged

Bubble Columns”. AIChE. J., 36, pp. 1485-1499 (1990)

45. Rotta, J. C. “Turbulent Stromungen”. Teubner, Stuttgart, pp 96 (1972)

46. Russell, T. W. F.; Charles, M. E. “The Effect of the Less Viscous Liquid in the

Laminar Flow of Two Immiscible Liquids”. Can. J. Chem. Eng. 37, pp. 18-24

(1959)

47. Russell, T. W. F.; Hodgson, G. W.; Govier, G. W. “Horizontal Pipeline Flow

Mixtures of Oil and Water”. Can. J. Chem. Eng. 37, pp. 9-17 (1959)

48. Shi, H.; Cai, J. Y.; Jepson, W. P. “The Effect of Surfactants on Flow

Characteristics in Oil/Water Flows in Large Diameter Horizontal Pipelines”.

MULTIPHASE 99, 9th Int. Conference, Cannes (1999)

49. Shi, H.; Cai, J. Y.; Gopal, M.; Jepson, W. P. “Oil Water Two Phase Flows in

Large Diameter Pipelines”. Proceedings of the ETCE/OMAE2000 Joint

Conference, New Orleans, LA (2000)

50. Shi, H. “A Study of Oil-Water Flows in Large Diameter Horizontal Pipelines”.

PhD Dissertation, Ohio University (2001)

51. Shi, H.; Jepson, W. P.; Rhyne. L. “Segregated Modeling of Oil-Water Flows”.

SPE 84232, presented at the SPE Annual Technical Conference and Exhibition in

Denver, CO (2003)

52. Shoham, O. “Mechanistic Modelling of Gas-Liqud Two-Phase Flow in Pipes”.

SPE Publications (to be published in 2006)

53. Simmons, M. J. H., Azzopardi, B. J. “Drop Size Distribution in Dispersed Liquid-

Liquid Pipe Flow”. Int. J. Multiphase Flow, 27, pp 843-859 (2001)

144

54. Smoluchowski, M. “Versuch einer Mathematischen Theorie der

Koagulationskinetik Kollider Losungen”. Z. Physik Chem., 92, pp. 129 (1917)

55. Soleimani, A.; Lawrence, C. J.; Hewitt, G. F. “Effect of Mixers on Flow Pattern

and Pressure Drop in Horizontal Oil-Water Pipe Flow”. MULTIPHASE 99, 9th

Int. Conference, Cannes (1999)

56. Soleimani, A. “Phase Distribution and Associated Phenomena in Oil-Water Flows

in Horizontal Tubes”. PhD Dissertation, University of London (1999)

57. Sontvedt, T.; Gramme P. “A Method and Device for the Separation of a Fluid in a

Well”. WO 98/41304 World Patent, September (1998)

58. Taitel, Y.; Barnea, D.; Brill, J. P. “Stratified Three Phase Flow in Pipes”. Int. J.

Multiphase Flow, 21, pp. 53-60 (1995)

59. Taitel, Y.; Dukler, A. E. “A Model for Predicting Flow Regime Transitions in

Horizontal and Near Horizontal Gas-Liquid Flow”. AIChE. J., 22, 47-55 (1976)

60. Trallero, J. L. “Oil Water Flow in Horizontal Pipes”. PhD Dissertation, The

University of Tulsa, Tulsa OK (1995)

61. Valle, A., Kvandal, H. K. “Pressure drop and dispersion characteristics of

separated oil/water flow”. Proceedings of the International Symposium on Two-

Phase Flow Modeling and Experimentation, Oct 9-11 Rome, Italy. Edizione ETS,

Vol 1, pp. 583-591 (1995)

62. Vedapuri, D; Bessette, D.; Jepson, W. P. “A Segregated Flow Model to Predict

Water Layer Thickness in Oil-Water Flows in Horizontal and Slightly Inclined

Pipelines”. MULTIPHASE 97, 8th Int. Conference, Cannes, pp. 75-105 (1997)

145

APPENDIX I

LOCAL MEASUREMENT SYSTEMS

The velocity and water cut measurements were obtained using a combination of

two-instruments: a continuously flushed pitot tube and a sampling tube. Also, a

borescope was used for droplet size measurement. The descriptions of each of these

measurement systems as well as the operating procedures followed are given below.

A1.1 Continuously Flushed Pitot Tube

A1.1.1 System Description

A continuously flushed pitot tube measurement system is similar to a standard

pitot tube metering system, but it also allows the continuous injection of a flushing fluid

that discharges through the pitot openings, preventing the flowing fluids to invade and

contaminate the internal chambers of the pitot. A proper design and operation can ensure

minimal disturbances and errors in the measured parameters (i.e., difference between the

overall and the static pressure of a flowing fluid). The most important physical

characteristics of the continuously flushed pitot tube that was built for this investigation

are:

146

- Materials and Diameters: brass built, 3/16” ID, 5/16” OD, nominal.

- Meeasurement equipment:

o 1 Differential pressure transducer:

Rosemount 3051CD1A22A1AQ4.

Calibration: 0-2 inches of water, accuracy ±0.0013 inch of water,

resolution 0.0001 inch of water.

o 2 Rotameters:

Total Pressure: Gilmont Accucal EW-32121-22.

• Calibration: 0-150 ml/min. Accuracy: 2% of reading, or 1

division (the largest).

Static Pressure: Gilmont Accucal EW-32121-18.

• Calibration: 0-17 ml/min. Accuracy: 2% of reading, or 1

division (the largest).

- Traverse mechanisms: liners, ¼” nominal thickness (average liner measured

thickness is 0.238± 0.008 inches).

Figure AI.1 shows a photo of the actual pitot tube system, Figure AI.2 presents a

schematic of the pitot tube with its dimensions, and Figure AI.3 is a schematic of the

experimental arrangement.

147

Figure AI.1 Photo of Actual Pitot Figure AI.2 Pitot Geometry

Tube Assembly Used (Lateral View) (not to scale)

Figure AI.3 Schematic of Continuously Flushed Pitot Tube Arrangement

HPS Flow direction

dP

Static press flush flowmeter

Total press flush flowmeter

Pressure source a) Tap Water b) Recirc. Pump

Pressure Meter-Pitot tube Assembly

Flushing Fluid

Liners

1/8”

2.5”

1/2”, φ 1/4”

3.0”

φ 1/32” φ 3/16”

Brass, φ 5/16” nominal

Brass, φ 3/16” nominal

1/4”

148

The pressure sources for the flushing fluid are:

- Tap water: 40 psig , ±1 psi oscillations

- Centrifugal recirculation pump: 14 to 30 psig.

Tests with both pressure sources were conducted, and better results were obtained

when the higher pressure source was used.

A1.1.2 Velocity Calculation

To calculate the velocity at a given point, the following procedure was used:

a) The pressure differential across the pitot tube is read.

b) The local velocity is calculated using the pitot tube equation:

( )ρ

212 PPv

Δ−Δ= .....................................................................................................(AI.1)

Where =Δ 1P Difference between total pressure and static pressure, and

=Δ 2P Zero velocity calibration constant, due to flushing flow through the pitot.

In this equation, density is estimated from a local no-slip flow assumption,

namely:

149

( )WCWC OWM −+= 1ρρρ ........................................................................................(AI.2)

If no WC value is known, 100% water is assumed.

c) After the velocity is calculated, water cut is measured with an iso-

kinetic sampling, using the calculated velocity as a reference.

Steps a) to c) are repeated until convergence. Due to the small density difference

between the oil and water, convergence is attained in two or three iterations.

The standard method for estimating 2PΔ is to measure its value while flushing when

there is no flow through the HPS©. The HPS should be filled with the expected

continuous phase fluid for the scheduled experiment.

A1.1.3 Operating Procedure

The following operating procedure was used for all pitot tube measurements:

1- Flushing fluid tank is filled with flushing fluid.

2- Flushing fluid tank is pressurized, and pressurization line and flushing fluid tank are

degassed.

3- Water circulation is established in the HPS© for separator degassing.

150

4- After the HPS© is degassed, the water flow is stopped. A small amount of oil is

initiated into the HPS©, until the water-oil interface level in the separator reaches

three quarters of the total height. Then, all flow is stopped.

5- Pitot tube is moved to the top of the pipe. The pitot must be in oil-continuous phase

before going to the next step.

6- Flushing fluid hoses are connected to pitot tube. Valves controlling access of

flushing fluid to pitot tube are opened.

7- Valves at rotameters are fully opened, one at a time, for cleaning and degassing the

tubing, pitot tube and differential pressure transducer with the flushing fluid. After

cleaning, flushing flowrates are adjusted to the desired values.

8- Flow to the HPS© is started again at the experimental run conditions. Pitot tube is

located at the centerline of the HPS©, and readings are monitored until steady-state

flow is attained.

9- When steady-state is attained, pitot tube is moved along the HPS© chord for local

velocity measurement.

10- Flow is stopped, and pitot tube is placed at the top of the experimental section. Steps

7 to 9 are repeated at different flowrates, until experimental matrix is completed.

11- After finishing the experiments, valves at rotameters are closed, then valves at pitot

tube assembly are also closed.

12- Flushing Fluid Vessel is depressurized. Hoses to pitot tube assembly may be

removed, if required.

151

A1.1.4 Calibration Results

The system was initially tested as a standard pitot tube in single phase-flow (fluid

flowing through the HPS© and inside the pitot are the same, no flushing is allowed).

Later, these measurements were repeated with flushing. The results with no flushing

conditions are shown in Figure AI.4.

Different flushing flowrate combinations were tested, for the expected HPS©

testing flowrates. From Figure AI.5 and Figure AI.6 show the results obtained with the

best flushing conditions found in this investigation:

- Flushing Fluid: Oil

- Flushing Flowrates:

o Static Pressure tap: between 0.022 and 0.026 gal/h

o Total Pressure tap: between 0.5 and 0.6 gal/h

For measurements while flushing, the standard procedure for estimating 2PΔ was

not valid, and an overall mass balance approach was used. This will be further explained

in a following section. From Figure AI.5, it can be seen that the flushing has almost no

effect on the results when pure oil is flowing. However, Figure AI.6 shows an increase of

the uncertainty of the velocity values. Note that the amplitude of these oscillations does

not change too much between experimental conditions. A discussion on this phenomenon

follows.

152

0%WC, v M=0.41ft/s w/mixer, θ=0º

Re=388

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

ght [

in]

PitotTheoretical

0%WC, v M=0.51ft/sw/mixer θ =0º

Re=482

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

ght [

in]

PitotTheoretical

(a) Average Velocity 0.35 ft/s , 0% WC (b) Average Velocity 0.52 ft/s, 0% WC

100%WC, v M=0.35 ft/s w/o mixer θ =0º

No flushing, Re=8500

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

ght [

in]

ExperimentalTheoretical

100%WC, v M=0.52 ft/s w/mixer, θ =0º

No flushingRe=15000

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

gth

[in]

PitotTheoretical

(c) Average Velocity 0.35 ft/s ,100% WC (d) Average Velocity 0.52 ft/s,100% WC

Figure AI.4 Pure Oil and Water Flow Velocity Profiles at 13.5 ft From Inlet. Pitot Filled With Same Fluid as HPS©, With no Flushing

152

153

0%WC, v M=0.41ft/sw/mixer, θ =0º

Flushing: Static 0.024 gal/hTotal: 0.56 gal/h

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

ght [

in]

PitotTheoretical

0%WC, v M=0.51ft/sw/mixer, θ =0º

Flushing:Static: 0.027 gal/hTotal=0.56 gal/h

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

PitotTheoretical

Hei

ght [

in]

Figure AI.5: Pure Oil Flow Velocity Profiles Measurement Inside the HPS at 13.5 ft From the Inlet. Pitot Tube Flushed With Oil

(a) Average Velocity: 0.41 ft/s

(b) Average Velocity: 0.51 ft/s

153

154

100%WC, v M=0.52 ft/sw/mixer, θ =0º

Flushing: Static 0.02 gal/hTotal 0.5 gal/h

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

ght [

in]

PitotTheoretical

100%WC,v M=0.58 ft/s w/mixer, θ =0º

Flushing: Static 0.02 gal/hTotal 0.55 gal/h

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

ght [

in]

PitotTheoretical

a) Average Velocity 0.52 ft/s b) Average Velocity 0.58 ft/s

100%WC, v M=0.87 ft/s w/mixer, θ =0º

Flushing: Static:0.042 gal/h Total: 0.56 gal/h

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

ght [

in]

PitotTheoretical

100%WC, v M=1.23 ft/s w/mixer θ =0º

Flushing: Static: 0.02 gal/hTotal: 0.56 gal/h

0

0.75

1.5

2.25

3

3.75

0 0.5 1 1.5 2Velocity [ft/s]

Hei

ght [

in]

PitotTheoretical

c) Average Velocity .87 ft/s d) Average Velocity 1.23 ft/s

Figure AI.6 Pure Water Flow Local Velocity Profile Measurements Inside the HPS© at 13.5 ft from Inlet. Pitot Tube Flushed with Oil.

154

155

A1.1.5 Continuously Flushing Pitot Tube Uncertainty Analysis

An explanation of the uncertainty analysis used for the uncertainty on the velocity

measurements follows. Subsequently, a discussion of uncertainty sources is given.

Uncertainty Analysis

The calculation of the velocity from a pitot tube is given by the equation AI.1.

When uncertainty propagation is applied to the expression, the following expression is

obtained:

( ) ( )

( )

( )

5.0

22

2

21

21

2

21

2

2

2/321

))((2

1

))((2

1

)(2

21

)(

⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜⎜

Δ⎟⎟⎠

⎞⎜⎜⎝

Δ−Δ−

+Δ⎟⎟⎠

⎞⎜⎜⎝

Δ−Δ

+⎟⎟⎠

⎞⎜⎜⎝

⎛ Δ−Δ−

=

PUPP

PUPP

UPP

vU

ρ

ρ

ρρ

..............................................................(AI.3)

The uncertainty of the velocity is a function of the uncertainties on the density of

the fluid impinging the pitot ρ, the uncertainty of the pressure reading at flowing

conditions ΔP1 and the uncertainty of the pressure at zero flowing conditions ΔP2.

In the present investigation, due to the high value of the density involved

(between 850 to 1000 kg/m3) and the small pressure drops measured (up to 2-in of water,

156

mainly due to the low velocities measured), the first term in the parenthesis at the RHS is

neglected, and the uncertainty is equal to:

( )( ) 5.02

22

121

))(())((2

1)( PUPUPP

vU Δ+Δ⎟⎟⎠

⎞⎜⎜⎝

Δ−Δ=

ρ...............................................(AI.4)

It is assumed that the uncertainty of the pressure drop at no-flow conditions is the

same as the one at flowing conditions, so the uncertainty is estimated as:

( ))(2

21)( 1

21

PUPP

vU Δ⎟⎟⎠

⎞⎜⎜⎝

Δ−Δ=

ρ.......................................................................(AI.5)

As can be inferred, the uncertainty will increase as the denominator goes to zero.

(very low velocities). The uncertainty of the pressure differential measured by the pitot

was calculated using standard procedures, through the equation:

( )5.0

2,

2

951 2)(

⎥⎥⎦

⎢⎢⎣

⎡+⎟

⎠⎞

⎜⎝⎛±=Δ RX

R SBtPU ..............................................................................(AI.6)

Where the systematic uncertainty (BR) is given by the instrument accuracy and

resolution (previously given), and the random uncertainty, RXS , , was estimated from the

data taken. Each velocity measurement at any given location comprised of a set of more

157

than 100 differential pressure readings, so the used Student’s t value was 2 (for 95%

confidence interval).

Uncertainty Sources

An increase in the uncertainty occurs when flushing is carried out and water is the

continuous-phase. This is caused both by the natural oscillations of the pressure readings

due to capillarity when oil droplets from the flushing oil flows out of the pitot through the

pitot holes; and also by HPS© pressure circuit oscillations, due to the liquid inflow and

outflow in the downstream 3-phase separator of the experimental facility.

The first cause of pressure oscillations in the readings can only be dampened by a

compromise in the design of the pitot tube: bigger the pitot tube and the pitot holes,

smaller the oscillations due to the flushing, but the measurement is less accurate . This

fact was taken in account on the pitot design used in this investigation, as different

diameters were tested, and the results were obtained with the best performance design.

An analysis of the second pressure oscillation source follows. A plot of the

pressure inside the downstream three-phase separator as a function of time is given in

Figure AI.7, from zero flow to steady-state flow conditions in the HPS©. The flowing

conditions through the HPS© are as follows: fluid: water, flowrate: 18.0 gal/min. The

flushing flowrates at the pitot were: 0.006 gal/h through the static pressure port, and 1.6

gal/h through the total pressure port.

158

Figure AI.7 Pressure at Downstream 3-phase Separator and Water Flowrate Through the

HPS© as Functions of Time During Test Start-up

Two oscillation cycles are seen in Figure AI.7: one with a period of approx 50

sec, with the same opening and closing frequency of the water drainage valve on the 3-

phase separator, and another not periodical, that dampens with time, with a local period

larger than 2 minutes, caused by the separator gas pressure control valve.

Note that the steady-state mass flow condition is reached quickly (in less than 200

sec). However, the pressure in the separator changes from zero flow conditions to steady-

state conditions in 1100 sec (approx 18 min). This delay is caused by internal operating

159

Figure AI.8 Pressure at 3-Phase Separator and Pitot dP Readings as Functions of Time

During Test Start-up

level rearrangements inside the 3-phase separator and in the pure oil and pure water

tanks.

A plot of the Pitot-dP reading and the downstream 3-phase separator internal

pressure as functions of time is given in Fig. AI.8. Both the Pitot-dP differential pressure

and the 3-phase separator absolute pressure have oscillations at similar frequencies.

The phenomenon occurs because the liquid used for flushing comes from a source

with a different pressure control loop than the HPS©. As valves at the rotameters control

160

the flushing flowrate, small changes on differential pressure between the flushing fluid

pressure source and the HPS© can cause small changes of flushing flowrates. These

flushing flowrate changes are hardly noticeable on the flushing flowmeter because of the

amplitude of the signal and the accuracy of the meters. Figure AI.9 shows a schematic of

the process described.

Figure AI.9 Effect of Oscillating Pressure in HPS© on the Flushing Calibration Constant.

This explanation also supports experimental evidence showing increasing pitot

differential pressure oscillation amplitude when gas is present on the flushing fluid tank

t

P

t

P

t

QFL α PΔ

HPS

QFL

161

or inside the pitot chambers. This gas gives extra elasticity to the system, amplifying

flushing flow oscillations.

Note that although these flow changes are small, the pitot differential pressure

reading is so small (due to the low velocities at the HPS©) that small changes can heavily

affect the final results. Finally, note that these oscillations are not seen on the pure-oil-

flow experiments with flushing. This is related to a smoother operation of the oil-draining

valve in the three-phase separator, due the oil viscosity, valve settings and three-phase

separator pressure set points.

To avoid these oscillations, the installation of a data filter is recommended, and/or

the recalibration of the 3-phase separator water discharge valve. The last recommendation

depends on the flexibility of the valve control system.

A1.1.6 Estimation of the Reference Differential Pressure 2PΔ

The standard procedure for determining the calibration differential pressure 2PΔ

fails for the following reason: the pressure differential between the flushing fluid source

and the HPS© is different when flow takes place in the HPS© as compared to when there

is no flow through the HPS©. This is due the hydraulic pressure losses due to the piping

between the HPS© and the pressure sink (three-phase separator). The higher the flowrate,

the higher will be the pressure inside the HPS©, as compared to the separator pressure.

This result is a smaller pressure differential between the HPS© and the flushing fluid

162

tank. Recalibration of the rotameter valves was attempted but this was not successful due

to the coarse valve response and the rotameter resolution.

Thus, the value of 2PΔ is obtained utilizing an alternative procedure, applying a

bulk volumetric flowrate balance, as the bulk flowrate of the phases is measured

upstream of the HPS© inlet. The following procedure was used:

a) The differential pressure profile was measured along the HPS©

diameter. If the flow is symmetric, the measurement was made only in

the vertical plane. For non-symmetric flow, measurements were made

at 0º, 30º, 60º and 90º from the vertical.

b) The velocity profile was constructed using a 2PΔ assumed value.

c) The overall flowrate is integrated from the experimental velocity profile

measurements, through discrete integration.

d) Steps b) and c) are repeated until the calculated overall flowrate

matches the measured flowrate.

All the results shown previously as well as the results in the dissertation are

obtained using this procedure.

163

A1.1.7 Conclusions

A method was devised for local velocity measurement for oil-water flow through

closed conduits when oil-water phases are present, from a similar method used by Lahey

(1987) for gas-liquid velocity measurement. The method was successfully employed to

measure the velocity profiles for pure oil and water flows, which matched the expected

theoretical results. Noise in the measurements was found, and corresponding to the

occurrence of pressure oscillations inside the HPS©.

164

A1.2 Isokinetic Sampling

Along with the velocity measurement, sampling was required for determination of

the average density of the flow impinging the pitot tube, and also the phase distribution

inside the HPS©. The sampling was carried out with the use of a sampling tube,

connected to a sampling vessel (with a volume of 300 cubic centimeters) where the

sample is collected and weighed, for determining the mixture water cut.

The sampling flowrate was estimated using the velocity profile measured through

the pitot tube, to ensure that the sampling occurred under iso-kinetic conditions.

Figure AI.10. Photo of the Sampling Tube Assembly

165

Figure AI.10. shows a photo of the actual sampling tube used in this study, and

Figure AI.11. shows a schematic of the experimental arrangement.

Figure AI.11 Schematic of Sampling Tube Operation

A1.2.1 Operating Procedure

The following procedure was used for the operation of the sampling tube system

(Refer to Figure AI.11):

a) Desired operating conditions are established in the HPS©.

b) Sampling tube is adjusted at the height of interest using the traverse liners.

HPS Flow direction

Sampling Vessel 1

Draining flowmeter

Collecting Cup

Liners

Sampling Vessel 2

Flushing Line

Drain

Sampling tube Assembly

166

c) With the valve at the draining flowmeter closed, the flushing line is opened.

Water flushes all the oil left in the sampling vessels and conduits into the HPS©.

d) After all the system is filled with water, flushing line valve is closed.

e) Sampling Vessel 2 (see Fig. AI.11) is disconnected and weighed. This weight will

be compared with the weight after sampling, for Water Cut estimations.

f) Sampling Vessel 1 suction and discharge valves are opened. Sampling Vessel 2

discharge valve is opened, while the suction valve is kept closed. Drain Valve is

opened.

g) Draining flowmeter valve is opened. Draining flowrate is adjusted to insure

sampling at isokinetic conditions.

h) A time window of 2 minutes is allowed for reaching the appropriate sampling

flowrate and reaching steady-state sampling flow conditions.

i) The following steps are applied simultaneously to start the sampling process:

a. Suction valve on Sampling Vessel 1 is closed, while suction valve on

Sampling Vessel 2 is opened.

b. Drain valve is closed, and collecting cup (see Figure AI.11) valve is

opened

j) Sampling is allowed until oil fills the sampling vessel, or oil-contaminated water

is obtained from the discharge of the Sampling Vessel 2.

k) When sampling is finished, the suction and discharge valves on Sampling Vessel

2 are closed.

l) Sampling Vessel 2 is disconnected and weighed, to estimate the amount of oil

sampled. WC is estimated.

167

m) Collecting cup fluid might be weighed also, for checking the sampling flowrates.

n) Steps b) to n) are repeated until the desired Water Cut measurements are made.

For water cut estimations, the following equation are used:

oilwater

finalvesselinitialvesseloil

MMV

ρρ −

−= ),(),( ..................................................................................(AI.7)

tQM

V Samplingwater

displacedwaterTotal ==

ρ) ( ................................................................................(AI.8)

%100*Total

oilTotal

VVVWC −

= ..............................................................................................(AI.9)

Uncertainty Estimation

The uncertainty of the water cut was estimated as follows: when Eq. AI.7 and

AI.8 are substituted in Eq. AI.9, the following expression is obtained:

%100*)()( ),(),(

tQMMtQ

WCSampling

oilwatfinalvesselinitialvesselSampling ρρ −−−= ..........................(AI.10)

Simplifying,

%100*)(

)(%100 ),(),(

oilwatSampling

finalvesselinitialvessel

tQMM

WCρρ −

−−= ....................................................(AI.11)

168

The uncertainty propagation equation is:

5.0

42

2

42

2

242

2

24

2

2),(),(

2

2),(

2),(

)()())((

)()())((

)()())((

))(()())((

)(

))(())(())((

%100)(

⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢

⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢

⎥⎥⎦

⎢⎢⎣

+⎥⎥⎦

⎢⎢⎣

+⎥⎥⎦

⎢⎢⎣

+⎥⎥⎦

⎢⎢⎣

+⎥⎥⎦

⎢⎢⎣

+

=

oilwatSampling

oil

oilwatSampling

wat

oilwatSampling

oilwatSampling

Sampling

finalvesselinitialvessel

oilwatSampling

finalvesselinitialvessel

tQU

tQU

tQtU

tQQU

MM

tQMUMU

WCU

ρρρ

ρρρ

ρρ

ρρ

ρρ

.(AI.12)

Regrouping

5.0

2

2

2

2

2

2

2

2

2

2),(),(

2

2),(

2),(

)())((

)())((

))((

)())((

))(()(

))(())(())((

%100)(

⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢

⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢

⎥⎦

⎤⎢⎣

⎡−

+⎥⎦

⎤⎢⎣

⎡−

+⎥⎦

⎤⎢⎣

+⎥⎥⎦

⎢⎢⎣

+⎥⎥⎦

⎢⎢⎣

+

=

oilwat

oil

oilwat

wat

Sampling

Sampling

oilwatSampling

finalvesselinitialvessel

oilwatSampling

finalvesselinitialvessel

U

U

ttU

QQU

tQMM

tQMUMU

WCU

ρρρ

ρρρρρ

ρρ

...................(AI.13)

The main source of uncertainty while sampling comes from the fact that the

sampling flowrate has higher uncertainties for smaller flowrates (from the rotameter

169

uncertainty). Also, smaller flowrates and short sampling times increase the effect of the

mass measurement uncertainty on the Water Cut uncertainty. Assuming that the most

important source of uncertainty is the sampling flowrate, the following expression is

obtained:

5.0

2

2

2

2),(),(

2

2),(

2),(

)())((

))(()(

))(())(())((

%100)(

⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢

+⎥⎥⎦

⎢⎢⎣

+

=

Sampling

Sampling

oilwatSampling

finalvesselinitialvessel

oilwatSampling

finalvesselinitialvessel

QQU

tQMM

tQMUMU

WCU

ρρ

ρρ.............................(AI.14)

The rotameter uncertainty in the lower 20% range was in the order of 10%. The

uncertainty on the mass measurement was in the order of ±1 gr, with a sample weight of

the order of 210 gr. The uncertainty in the densities is considered 0.1%, and in the time

was 0.1 sec. The uncertainty increases at lower velocity zones as sampling flowrates were

decreased for maintaining isokinetic conditions. Due to time constrains only one water

cut measurement per location was obtained.

170

A1.3 Borescope for Droplet Size Measurement

Droplet Size Measurement was carried out through video image processing,

utilizing an Olympus K27-18-00-62 borescope, equipped with a 90 degree mirror sleeve

for allowing lateral view. The borescope was installed inside a closed, protecting acrylic

sleeve, full of glycerin, to avoid contact between the borescope and pressurized oil-water

flow.

The borescope is equipped with a GENWAC Neptune N-100 black and white

camera. This camera is specially suited for low-light conditions, and has shutter speeds

up to1/100000 of a second, which is well suited for this application. The standard video

signal was digitized using a Hauppauge WINTV-USB2 video digitizer and transformed

in a MPEG format video, with a resolution of 800x600 pixels, at 30 frames per second. A

SONY VAIO PCG-V505DX laptop was used for video saving and post-proccessing.

After the test run, the videos were processed using Roxio Videowave 7 software,

where frames were chosen for droplet size analysis. The chosen frames were analyzed

using the SigmaScan Pro software.

Figure AI.12 shows the borescope installed in the experimental facility.

171

Figure AI.12 Borescope Experimental Arrangement

A1.3.1 Calibration

The calibration was performed taking images of an Olympus Series 5 Borescope

Test Chart, in contact with the observation window of the borescope sleeve. Figure AI.13

is a image of the 1 millimeter grid of the test chart. From the image, a correlation

between pixels and physical dimensions was obtained, and used to estimate the diameter

of the droplets flowing in front of the borescope sleeve observation window.

Borescope

Video Camera

172

Figure AI.13 Image of the 1mm x 1mm Squares Grid of a Olympus Series 5 Borescope

Test Chart as Observed Through the Borescope

A1.3.2. Measurement procedure

a) The borescope is set at the required level inside the HPS©.

b) Flow is initiated in the HPS©, for an experimental run.

c) When the system attains steady-state, a video of the flow in front of the

observation window is taken.

173

Figure AI.14. Frame from Video Taken at 3-in from Bottom, at 7.5 ft from Inlet, at the

Vertical Plane. Flow Conditions: 50% WC, vM=0.58 ft/s, with Mixer

d) From this frame, droplets are painted in black, assuming oval or circular shape.

This procedure is required to obtain reliable diameter results from the

SigmascanTM software. Multiple copies are required to process overlapping

droplets.

Figure AI.15 Frame Shown at Figure AI.14, After Processing

174

e) SigmascanTM measures the droplet area, perimeter and equivalent diameter for a

chosen droplet. The data are sent to ExcelTM to obtain histograms and relevant

information.

Figure AI.16 Example of SigmaScan Pro Output

175

Uncertainty Analysis

The uncertainty on the measurement comes from the deformation of the grid due

to the borescope optics, and to the grid of the bitmap when the picture is digitized. This

value was estimated as 75 microns from the averaging of the measured lengths of the

square sides of the Fig. AI.13.

176

APPENDIX II

LAYER HEIGHT COMPARISON BETWEEN LOCAL WATER CUT AND

PHOTOGRAPHIC METHODS MEASUREMENTS

In this appendix, a comparison between the layer heights measurements and

photographic measurements is presented. The layer height between the packed

dispersion/ loose dispersion layer was measured using as a criterion of local value of

WC=75% for the determination of the boundary between these layers. This measurement

was obtained at 7.5 and 13.5 ft from the inlet, while the photograph was taken at 14 ft

from the inlet, at the observation window. As can be seen in Figs. AII.1 through AII.6,

the WC method boundary heights results are slightly higher than the photographic

method ones.

177

AII.1 Comparison Between Photographic (left) and WC (right) Level Measurement

Methods for 30%WC, vM=0.44 ft/s, With Mixer, at θ=0º

AII.2 Comparison Between Photographic (left) and WC (right) Level Measurement

Methods for 50%WC, vM=0.44 ft/s, With Mixer, at θ=0º

0

0.2

0.4

0.6

0.8

1

0 10 20

Lenght [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/ LooseDispersion Boundary(75%WC local)

0

0.2

0.4

0.6

0.8

1

0 10 20Lenght [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/ LooseDispersion Boundary(75%WC local)

Length [ft]

Length [ft]

178

AII.3 Comparison Between Photographic (left) and WC (right) Level Measurement

Methods for 70%WC, vM=0.44 ft/s, With Mixer, at θ=0º

AII.4 Comparison Between Photographic (left) and WC (right) Level Measurement

Methods for 30%WC, vM=0.58 ft/s, With Mixer, at θ=0º

0

0.2

0.4

0.6

0.8

1

0 10 20

Lenght [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/ LooseDispersion Boundary(75%WC Local)

0

0.2

0.4

0.6

0.8

1

0 10 20Lenght [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/ LooseDispersion Boundary(75%WC Local)

Length [ft]

Length [ft]

179

AII.5 Comparison Between Photographic (left) and WC (right) Level Measurement

Methods for 50%WC, vM=0.58 ft/s, With Mixer, at θ=0º

AII.6 Comparison Between Photographic (left) and WC (right) Level Measurement

Methods for 70%WC, vM=0.58 ft/s, With Mixer, at θ=0º

0

0.2

0.4

0.6

0.8

1

0 10 20Lenght [ft]

h/D

Oil/Packed DispersionBoundary

Packed Dispersion/ LooseDispersion Boundary(75%WC Local)

0

0.2

0.4

0.6

0.8

1

0 10 20Lenght [ft]

h/D

Oil-Packed Dispersion

Packed Dispersion/ LooseDispersion Boundary(75%WC Local)

Length [ft]

Length [ft]

180

APPENDIX III

LOCAL VELOCITY MEASUREMENT

In Chapter 3, Figs. 3.9 and 3.10 show the measured velocity profile at the vertical

plane (θ=0º), and the velocity profile contours at the HPS© cross-sectional area are shown

in Figs. 3.11 and 3.12 at 7.5 and 13.5 ft from the inlet. The following figures show the

measurement chords (Fig AIII.1), and the velocity profiles obtained at 7.5 and 13.5 ft

from the inlet of the HPS© for the different operating conditions (Figs AIII.2 through

AIII.7). As can be seen, the results are given not only in the vertical plane (0º), but also at

three more chords (θ=0º, θ=30º, θ=60º, and θ=90º).

Fig AIII.1 Location of the Different Velocity Measurement Chords in the HPS©

Cross-Sectional Area

0º from vertical 30º from vertical

60º from vertical

90º from vertical

HPS© Cross Section (Flow into the page plane)

g

181

10%WC, v M=0.44ft/sw/mix, θ=0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

10%WC, v M=0.44ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

10%WC, v M=0.44ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

10%WC, v M=0.44ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

Fig. AIII.2 Velocity Measurements at Different Chords (10%WC, vM=0.44 ft/s, 7.5 ft From Inlet, With Mixer)

181

182

10%WC, v M=0.44ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

10%WC, v M=0.44ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

10%WC, v M=0.44ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

10%WC, v M=0.44ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

Fig. AIII.3 Velocity Measurements at Different Chords (10%WC, vM=0.44 ft/s, 13.5 ft From Inlet, With Mixer)

182

183

10%WC v M= 0.58 ft/s w/mix, θ=0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

10%WC, v M=0.58ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

10%WC, v M=0.58ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

10%WC, v M=0.58ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

Fig. AIII.4 Velocity Measurements at Different Chords (10%WC, vM=0.58 ft/s, 7.5 ft From Inlet, With Mixer)

183

184

10%WC, v M=0.58ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

10%WC, v M=0.58ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

10%WC, v M=0.58ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

10%WC, v M=0.58ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

Fig. AIII.5 Velocity Measurements at Different Chords (10%WC, vM=0.58 ft/s, 13.5 ft From Inlet, With Mixer)

184

185

30%WC, v M=0.44 fts/s, w/mix, θ=0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght

[in]

7.5 ft

30%WC, v M=0.44ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

30%WC, v M=0.44ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

30%WC, v M=0.44ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

Fig. AIII.6 Velocity Measurements at Different Chords (30%WC, vM=0.44 ft/s, 7.5 ft From Inlet, With Mixer)

185

186

30%WC, v M=0.44ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

30%WC, v M=0.44ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

30%WC, v M=0.44ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

30%WC, v M=0.44ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

Fig. AIII.7 Velocity Measurements at Different Chords (30%WC, vM=0.44 ft/s, 13.5 ft From Inlet, With Mixer)

186

187

30%WC v M=0.58 ft/s w/mix, θ=0º

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

30%WC, v M=0.58ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

30%WC, v M=0.58ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

30%WC, v M=0.58ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

Fig. AIII.8 Velocity Measurements at Different Chords (30%WC, vM=0.58 ft/s, 7.5 ft From Inlet, With Mixer)

187

188

30%WC, v M=0.58ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

30%WC, v M=0.58ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

30%WC, v M=0.58ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

30%WC, v M=0.58ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

Fig. AIII.9 Velocity Measurements at Different Chords (30%WC, vM=0.58 ft/s, 13.5 ft From Inlet, With Mixer)

188

189

50%WC, v M=0.44 ft/s, w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

50%WC, v M=0.44ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

50%WC, v M=0.44ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

50%WC, v M=0.44ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

Fig. AIII.10 Velocity Measurements at Different Chords (50%WC, vM=0.44 ft/s, 7.5 ft From Inlet, With Mixer)

189

190

50%WC, v M=0.44ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

50%WC, v M=0.44ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

50%WC, v M=0.44ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

50%WC, v M=0.44ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

Fig. AIII.11 Velocity Measurements at Different Chords (50%WC, vM=0.44 ft/s, 13.5 ft From Inlet, With Mixer)

190

191

50%WC, v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

50%WC, v M=0.58ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

50%WC, v M=0.58ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

50%WC, v M=0.58ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

Fig. AIII.12 Velocity Measurements at Different Chords (50%WC, vM=0.58 ft/s, 7.5 ft From Inlet, With Mixer)

191

192

50%WC, v M=0.58ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

50%WC, v M=0.58ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

50%WC, v M=0.58ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

50%WC, v M=0.58ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

Fig. AIII.13 Velocity Measurements at Different Chords (50%WC, vM=0.58 ft/s, 13.5 ft From Inlet, With Mixer)

192

193

70%WC, v M=0.44 ft/s, w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

70%WC, v M=0.44ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

70%WC, v M=0.44ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

70%WC, v M=0.44ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

Fig. AIII.14 Velocity Measurements at Different Chords (70%WC, vM=0.44 ft/s, 7.5 ft From Inlet, With Mixer)

193

194

70%WC, v M=0.44ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

70%WC, v M=0.44ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

70%WC, v M=0.44ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

70%WC, v M=0.44ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

Fig. AIII.15 Velocity Measurements at Different Chords (70%WC, vM=0.44 ft/s, 13.5 ft From Inlet, With Mixer)

194

195

70%WC v M=0.58 ft/s w/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

70%WC, v M=0.58ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

70%WC, v M=0.58ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

70%WC, v M=0.58ft/s w/mix, θ =90°0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

7.5 ft

Fig. AIII.16 Velocity Measurements at Different Chords (70%WC, vM=0.58 ft/s, 7.5 ft From Inlet, With Mixer)

195

196

70%WC, v M=0.58ft/sw/mix, θ =0°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

70%WC, v M=0.58ft/sw/mix, θ =30°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

70%WC, v M=0.58ft/sw/mix, θ =60°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

70%WC, v M=0.58ft/sw/mix, θ =90°

0.00

0.75

1.50

2.25

3.00

3.75

0.0 0.3 0.6 0.9 1.2 1.5 1.8Velocity [ft/s]

Hei

ght [

in]

13.5 ft

Fig. AIII.17 Velocity Measurements at Different Chords (70%WC, vM=0.58 ft/s, 13.5 ft From Inlet, With Mixer)

196

197

APPENDIX IV

CALCULATION OF THE VELOCITY AND WATER CUT SLOPES FOR

LINEAR VELOCITY PROFILE APPROXIMATION

The following equations present a step-by-step calculation of the expressions used

for the estimation of the slopes of the linear velocity and water cut profiles considered on

the packed layer section, as stated in section 3.4

Estimation of the velocity profile slope

Starting from Eq. 4.32, the velocity profile in the packed zone is assumed to be

linear:

)~~()~~)(~(

)~( ,OW

OMOi hh

hhhddvvhv

−+= − ..............................................................................(AIV.1)

The integration of this velocity profile between the upper and lower boundaries of

the packed layer will give the flowrate of the packed layer:

∫= vdAQ ..................................................................................................................(AIV.2)

198

hdhvhDQO

W

h

h

~)~()))1~2(1((~

~

22∫ −−= ...........................................................................(AIV.3)

hdhh

hhhddvvhDQ

O

W

h

h OW

OMOi

~))~~(

)~~)(~()())1~2(1((

~

~,

22∫ −

−+−−= − .......................................(AIV.4)

The RHS can be expressed in two integrals

hdhDhh

hhhddvhdhDvQ

O

W

O

W

h

h OW

Oh

hMOi

~)))1~2(1()~~(

)~~)(~((~)))1~2(1((

~

~

22

~

~

22, ∫∫ −−

−+−−= − .(AIV.5)

The indefinite solution of each of the the integrals is:

⎥⎦

⎤⎢⎣

⎡ −+−−

=−−

−∫

2)1~2arcsin()~~)(1~2(

2)(

~)))1~2(1((

5.022,

22,

hhhhDv

hdhDv

MOi

MOi

....................................................(AIV.6)

⎥⎥⎥⎥

⎢⎢⎢⎢

−−−

+−+−−

⎥⎥⎥⎥

⎢⎢⎢⎢

+−⎥⎦

⎤⎢⎣

⎡ −+−−

−=

−−−∫

5.02

5.12

2

5.12

5.02

2

22

)~~)(1~2)(5.0~(21

)~~(32)1~2arcsin()5.0~(

41

)~~()~(

)~~(31

)5.0~(8

)1~2arcsin()~~)(1~2(41

)~~()~(2

~))~~(

)~~)(~()())1~2(1((

hhhh

hhhh

hhhddvD

hh

hhhhh

hhhddvD

hdhh

hhhddvhD

O

O

WO

O

WO

WO

O

..............(AIV.7)

199

After simplification, Eq. AIV.5 transforms to:

O

W

h

hh

W

WOMOi

WO

hddvvhh

hhddvhhhh

hhvhhhh

hhDQ

~

~2

5.12

5.02

,5.02

2

)~()(31

)~5.0()~()12arcsin(81))(12(

41

)~~()12arcsin(81))(12(

41

~~2

⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢

⎟⎠⎞

⎜⎝⎛ −

−−⎟⎠⎞

⎜⎝⎛ −+−−

+−⎟⎠⎞

⎜⎝⎛ −+−−

−−=

..(AIV.8)

The LHS of the expression can be substituted with Eq. 4.37. The resultant

equation is solved for the slope of the velocity profile, from where the expressions for

Equations 4.41 to 4.44 are obtained.

Estimation of the Water Cut Profile slope

For the water cut profile, the total water flowrate in the packed layer can be

calculated from the following expression:

∫= WCvdAQW ..........................................................................................................(AIV.9)

hdhvhWChDQO

W

h

hW

~)~()~()))1~2(1((~

~

22∫ −−= ...........................................................(AIV.10)

200

hd

hhhhhddv

v

hhhhhddWC

WC

hDQO

W

h

hWO

O

MOi

WO

O

MOi

W~

)~~()~~)(~(

)~~()~~)(~()))1~2(1((

~

~

,,

22∫⎟⎟⎟⎟

⎜⎜⎜⎜

⎟⎟⎟⎟

⎜⎜⎜⎜

+

⎟⎟⎟⎟

⎜⎜⎜⎜

+

−−=−−

........(AIV.11)

Now, the RHS can be split into four expressions:

hdhh

hhhddvhh

hhhddWChD

hdhh

hhhddvvWChD

hdvhh

hhhddWCWChD

hdvWChD

Q

O

W

O

W

O

W

O

W

h

h WO

O

WO

O

h

h WO

OMOiMOi

h

hMOi

WO

OMOi

h

hMOiMOi

W

~))~~(

)~~)(~()(

)~~()~~)(~(

)())1~2(1((

~))~~(

)~~)(~()()())1~2(1((

~))()~~(

)~~)(~()())1~2(1((

~))()())1~2(1((

~

~

22

~

~,,

22

~

~,,

22

~

~,,

22

−−−

+−

−+−−

+−

−+−−

+−−

=

−−

−−

−−

.....................(AIV.12)

The indefinite integrals of the four expressions follow:

a)

⎥⎦

⎤⎢⎣

⎡ −+−−

=−−

−−

−−∫

2)1~2arcsin()~~)(1~2(

2))((

~))()())1~2(1((

5.022,,

,,22

hhhhDWCv

hdvWChD

MOiMOi

MOiMOi

.................................(AIV.13)

The second and third expressions are very similar:

201

b)

⎥⎥⎥⎥

⎢⎢⎢⎢

−−−

+−+−−

⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢

+−

⎥⎥⎥⎥

⎢⎢⎢⎢

+−−

−=

−−−

−∫

5.02

5.12

2,

5.12

5.02

2,

,22

)~~)(1~2)(5.0~(21

)~~(32)1~2arcsin()5.0~(

41

)~~()~()(

)~~(31

)5.0~(

8)1~2arcsin(

)~~)(1~2(41

)~~()~()(2

~))()~~(

)~~)(~()())1~2(1((

hhhh

hhhh

hhhddWCDv

hh

hh

hhh

hhhddWCDv

hdvhh

hhhddWChD

O

O

WOMOi

O

WOMOi

MOiWO

O

....................(AIV.14)

After regrouping:

⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢

+−−−

+−−

−=

−−−

−∫

5.12

5.022,

,22

)~~(32

)~~)(1~2)(5.0~(21

)1~2arcsin()5.0~(41

)~~()~()(

~))()~~(

)~~)(~()())1~2(1((

hh

hhhh

hh

hhhddWCDv

hdvhh

hhhddWChD

O

O

WOMOi

MOiWO

O

.................................(AIV.15)

202

c)

⎥⎥⎥⎥

⎢⎢⎢⎢

−−−

+−+−−

⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢

+−

⎥⎥⎥⎥

⎢⎢⎢⎢

+−−

−=

−−−

−∫

5.02

5.12

2,

5.12

5.02

2,

,22

)~~)(1~2)(5.0~(21

)~~(32)1~2arcsin()5.0~(

41

)~~()~()(

)~~(31

)5.0~(

8)1~2arcsin(

)~~)(1~2(41

)~~()~()(2

~))()~~(

)~~)(~()())1~2(1((

hhhh

hhhh

hhhddvDWC

hh

hh

hhh

hhhddvDWC

hdWChh

hhhddvhD

O

O

WOMOi

O

WOMOi

MOiWO

O

.................(AIV.16)

After regrouping:

⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢

+−−−

+−−

−=

−−−

−∫

5.12

5.022,

,22

)~~(32

)~~)(1~2)(5.0~(21

)1~2arcsin()5.0~(41

)~~()~()(

~))()~~(

)~~)(~()())1~2(1((

hh

hhhh

hh

hhhddvDWC

hdWChh

hhhddvhD

O

O

WOMOi

MOiWO

O

...............................(AIV.17)

And the fourth term:

203

d)

⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢

−−−

−−−−−

−−−−−

=−

−−−∫

)1~2arcsin()165~~(

41

)~~)(1~2)(165~~(

21

)~~(2

~)~~)(

165~(

34

)~~()~)(~(

~))~~(

)~~)(~()(

)~~()~~)(~(

)())1~2(1((

2

5.022

5.125.12

2

22

hhh

hhhhh

hhhhhh

hhhddWChddv

hdhh

hhhddvhh

hhhddWChD

OO

OO

O

WO

WO

O

WO

O

.....................(AIV.18)

When these expressions are added, the following expression is obtained for the

water flowrate through the packed layer:

204

O

W

h

hWO

WOO

MOiMOi

WO

WO

OO

WO

MOiMOiO

MOiMOi

WO

MOiMOiO

MOiMOi

WO

OO

W

hhhddWChddvhhh

hh

hhhddWChddvh

hddWCvhddvWC

hh

h

hhhddWChddvhh

hh

hddWCvhddvWCh

WCv

hhh

hhhddWCvhddvWC

h

WCv

hhhddWChddvhh

Q~

~2

5.12

5.12,,

2

2

,,

,,

5.02

,,

,,2

2

)~~()~)(~()~~(~

21

)~~(

)~~()~)(~()

85~2(

)~()~(

)~~(32

)1~2arcsin(

)~~()~)(~()

4

~

645

4

~(

)~~(

)~()~()5.0~(

41

4

)~~)(~21(

)~~()~()~(

)5.0~(21

2)~~()~)(~()

2

~

325

2

~(

⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢⎢

−−

−−⎥⎥⎥

⎢⎢⎢

−−

++

+−

⎥⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢

−−−

−⎥⎥⎦

⎢⎢⎣

+−

+−−

⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢

⎥⎥⎦

⎢⎢⎣

+−

−−−

−−

=

−−

−−

−−

−−

−−

(AIV.19)

As previously done, the LHS can be calculated from a given expression, in this

case the Eq. 4.39. Then it is possible to solve this final expression for the slope of the

water cut, as a function of the overall water flowrate, the heights of the oil and water

boundaries, and the slope of the velocity, from where Eq. 4.45 to 4.48 are obtained.

205

APPENDIX V

DROPLET SIZE MEASUREMENTS

The included CD have a compilation of tables with the droplet size distributions

measured at different heights in the two metering ports of the HPS, for all the

experimental conditions.

Each of the tables indicates the experimental conditions, the metering port, and

the height from the bottom of the pipe where the data was measured.


Recommended