Heriot-Watt University

Similar documents
Interfacial Properties at Elevated Pressures in Processes of Enhanced Oil and Gas Recovery

Interfacial Tension of Reservoir Fluids: an Integrated Experimental and Modelling Investigation

Heriot-Watt University

Comparison of the GERG-2008 and Peng-Robinson Equations of State for Natural Gas Mixtures

Heriot-Watt University

Preliminary Evaluation of the SPUNG Equation of State for Modelling the Thermodynamic Properties of CO 2 Water Mixtures

Experimental Vapor-Liquid Equilibria for the Carbon Dioxide + Octane, and Carbon Dioxide + Decane Systems from 313 to 373 K

CALCULATION OF THE COMPRESSIBILITY FACTOR AND FUGACITY IN OIL-GAS SYSTEMS USING CUBIC EQUATIONS OF STATE

Prediction of methanol content in natural gas with the GC-PR-CPA model Hajiw, Martha; Chapoy, Antonin; Coquelet, Christophe; Lauermann, Gerhard

Modeling of Pressure Dependence of Interfacial Tension Behaviors of Supercritical CO 2. + Crude Oil Systems Using a Basic Parachor Expression

Correlation of High Pressure Density Behaviors for Fluid Mixtures made of Carbon Dioxide with Solvent at K

Density and phase equilibrium of the binary system methane + n-decane under high temperatures and pressures

Thermodynamics I. Properties of Pure Substances

A modification of Wong-Sandler mixing rule for the prediction of vapor-liquid equilibria in binary asymmetric systems

Chemical Potential. Combining the First and Second Laws for a closed system, Considering (extensive properties)

Modeling Vapor Liquid Equilibrium of Binary and Ternary Systems of CO 2 + Hydrocarbons at High-Pressure Conditions

"Energy Applications: Impact of Data and Models"

Measurement and Prediction of Speed of Sound in Petroleum Fluids

EOS Higher Oil School 2017/5/26

PVTpetro: A COMPUTATIONAL TOOL FOR ISOTHERM TWO- PHASE PT-FLASH CALCULATION IN OIL-GAS SYSTEMS

Prediction of surface tension of binary mixtures with the parachor method

EXPERIMENTAL SETUP AND PROCEDURE

PETE 310 Lectures # 36 to 37

Petroleum Thermodynamic Research Group

SOLUBILITY OF CO 2 IN BRANCHED ALKANES IN ORDER TO EXTEND THE PPR78 MODEL TO SUCH SYSTEMS

Fluid Mechanics Introduction

Properties of Gases. The perfect gas. States of gases Gas laws Kinetic model of gases (Ch th ed, th ed.) Real gases

High-Pressure Volumetric Analyzer

LIQUID-LIQUID EQUILIBRIUM IN BINARY MIXTURES OF 1-ETHYL-3-METHYLIMIDAZOLIUM ETHYLSULFATE AND HYDROCARBONS

Available online at ScienceDirect. Energy Procedia 63 (2014 ) GHGT-12

(Refer Slide Time: 00:00:43 min) Welcome back in the last few lectures we discussed compression refrigeration systems.

Chapter 2: The Physical Properties of Pure Compounds

Thermodynamic Properties of Low-GWP Refrigerant for Centrifugal Chiller

PETROLEUM ENGINEERING 310 FIRST EXAM. September 19, 2001

AN EXPERIMENTAL INVESTIGATION OF BOILING HEAT CONVECTION WITH RADIAL FLOW IN A FRACTURE

CHAPTER 3 EXPERIMENTAL SET UP AND PROCEDURE

Hydrocarbon Reservoirs and Production: Thermodynamics and Rheology

Published in: Journal of Chemical & Engineering Data. DOI: /je Document Version Peer reviewed version

INHOMOGENEOUS DENSITY THEORIES COUPLED INTO A MOLECULAR EQUATION OF STATE FOR THE DESCRIPTION OF THE FLUID-FLUID INTERFACE

INVESTIGATION OF THE EFFECT OF TEMPERATURE AND PRESSURE ON INTERFACIAL TENSION AND WETTABILITY

Hydrate Formation: Considering the Effects of Pressure, Temperature, Composition and Water

INTRODUCTION AND BASIC CONCEPTS. Chapter 1. Mehmet Kanoglu. Thermodynamics: An Engineering Approach, 6 th Edition. Yunus A. Cengel, Michael A.

Phase Equilibria of binary mixtures by Molecular Simulation and PR-EOS: Methane + Xenon and Xenon + Ethane

A- Determination Of Boiling point B- Distillation

Chapter 1. The Properties of Gases Fall Semester Physical Chemistry 1 (CHM2201)

Carbon dioxide removal processes by alkanolamines in aqueous organic solvents Hamborg, Espen Steinseth

Thermodynamic Properties of Refrigerant R116 from Cubic Equations of State

ScienceDirect. Modelling CO 2 Water Thermodynamics Using SPUNG Equation of State (EoS) concept with Various Reference Fluids

Thermophysical Properties of Ethane from Cubic Equations of State

A 3 Vapor pressure of volatile liquids

This paper was prepared for presentation at the Unconventional Resources Technology Conference held in Denver, Colorado, USA, August 2014.

Rigorous calculation of LNG flow reliefs using the GERG-2004 equation of state

Viscosity and Liquid Density of Asymmetric n-alkane Mixtures: Measurement and Modeling 1

Chapter 10. Vapor/Liquid Equilibrium: Introduction

Surface Tension Prediction for Liquid Mixtures

Pool Boiling Heat Transfer to Pure Liquids

P1: IML/FFX P2: IML/FFX QC: IML/FFX T1: IML AT029-FM AT029-Manual AT029-Manual-v8.cls December 11, :59. Contents

12. Heat of melting and evaporation of water

Chemistry: The Central Science

Measurement and Calculation of Physico-Chemical Properties of Binary Mixtures Containing Xylene and 1- Alkanol

EPSRC Centre for Doctoral Training in Industrially Focused Mathematical Modelling

Chromatography. Gas Chromatography

Chapter 1 INTRODUCTION AND BASIC CONCEPTS

High-pressure qnmr spectroscopy in condensed- and gas-phase towards determination of impurities and compositions of gas mixtures

Vapor-hydrate phases equilibrium of (CH 4 +C 2 H 6 ) and (CH 4 +C 2 H 4 ) systems

States of matter Part 2

Detection and measurements of high pressure VLE and LLE - PvT data of binary mixtures through the vibrating tube densitometer technique.

Effect of Jatropha Bio-Surfactant on Residual Oil during Enhanced Oil Recovery Process

Vapor liquid equilibrium of carbon dioxide with ethyl caproate, ethyl caprylate and ethyl caprate at elevated pressures

18 a 21 de novembro de 2014, Caldas Novas - Goiás THERMODYNAMIC MODELING OF VAPOR-LIQUID EQUILIBRIUM FOR PETROLEUM FLUIDS

Enhanced Oil Recovery with CO2 Injection

Solubility of CO2 in the low-viscous ionic liquid 1-ethyl-3- methylimidazolium tris(pentafluoroethyl)trifluorophosphate

MODELING THE SURFACE TENSION OF PURE ALKANES AND PERFLUOROALKANES USING THE

A Multi-Continuum Multi-Component Model for Simultaneous Enhanced Gas Recovery and CO 2 Storage in Stimulated Fractured Shale Gas Reservoirs Jiamin

SCAL, Inc. Services & Capabilities

INTERNATIONAL STANDARD

SOFTWARE INTELIGENT PACKAGE FOR PHASE EQULIBRIA (PHEQ) IN SYSTEMS APPLIED IN CHEMISTRY AND CHEMICAL ENGINEERING

Dynamic Effects of Diabatization in Distillation Columns

CHAPTER SIX THERMODYNAMICS Vapor-Liquid Equilibrium in a Binary System 6.2. Investigation of the Thermodynamic Properties of Pure Water

Applied Fluid Mechanics

Changes of polymer material wettability by surface discharge

Modelling of methane gas hydrate incipient conditions via translated Trebble-Bishnoi-Salim equation of state

Modelling the Solubility of Solid Aromatic Compounds in Supercritical Fluids

Study of adsorption and desorption of asphaltene sediments inhibitor in the bottomhole formation zone

Experimental Methods and Analysis

Adam G. Hawley Darin L. George. Southwest Research Institute 6220 Culebra Road San Antonio, TX 78238

Upstream LNG Technology Prof. Pavitra Sandilya Department of Cryogenic Engineering Centre Indian Institute of Technology, Kharagpur

High-Pressure and High-Temperature Density Measurements of n-pentane, n-octane, 2,2,4-Trimethylpentane, Cyclooctane, n-decane, and Toluene

Diffusion and Adsorption in porous media. Ali Ahmadpour Chemical Eng. Dept. Ferdowsi University of Mashhad

THE DETERMINATION OF CRITICAL FLOW FACTORS FOR NATURAL GAS MIXTURES

Effect of positive rate sensitivity and inertia on gas condensate relative permeability at high velocity

Thermodynamics of Liquid (Xenon + Methane) Mixtures

Contribution to the study of neon-nitrogen mixtures at low temperatures

PREDICTION OF SATURATED LIQUID VOLUMES FROM A MODIFIED VAN DER WAALS EQUATION. By Charles R. Koppany

Isobaric Vapour-Liquid Equilibrium of Binary Mixture of 1, 2-Di-chloroethane with 1-Heptanol at Kpa

Peng-Robinson Equation of State Predictions for Gas Condensate Before and After Lumping

Theoretical Design and Analysis of Gravity Assisted Heat Pipes

Biochemistry. Biochemical Techniques HPLC

Chapter Four. Experimental

4.1. Physics Module Form 4 Chapter 4 - Heat GCKL UNDERSTANDING THERMAL EQUILIBRIUM. What is thermal equilibrium?

Transcription:

Heriot-Watt University Heriot-Watt University Research Gateway Measurement and Modelling of High Pressure Density and Interfacial Tension of (Gas + n- Alkane) Binary Mixtures Cravo Pereira, Luís Manuel; Chapoy, Antonin; Burgass, Rhoderick William; Tohidi Kalorazi, Bahman Published in: Journal of Chemical Thermodynamics DOI: 10.1016/j.jct.2015.12.036 Publication date: 2016 Document Version Peer reviewed version Link to publication in Heriot-Watt University Research Portal Citation for published version (APA): Cravo Pereira, L. M., Chapoy, A., Burgass, R. W., & Tohidi Kalorazi, B. (2016). Measurement and Modelling of High Pressure Density and Interfacial Tension of (Gas + n-alkane) Binary Mixtures. Journal of Chemical Thermodynamics, 97, 55-69. DOI: 10.1016/j.jct.2015.12.036 General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Accepted Manuscript Measurement and Modelling of High Pressure Density and Interfacial Tension of (Gas + n-alkane) Binary Mixtures Luís M.C. Pereira, Antonin Chapoy, Rod Burgass, Bahman Tohidi PII: S0021-9614(16)00026-4 DOI: http://dx.doi.org/10.1016/j.jct.2015.12.036 Reference: YJCHT 4524 To appear in: J. Chem. Thermodynamics Received Date: 24 September 2015 Revised Date: 9 November 2015 Accepted Date: 16 December 2015 Please cite this article as: L.M.C. Pereira, A. Chapoy, R. Burgass, B. Tohidi, Measurement and Modelling of High Pressure Density and Interfacial Tension of (Gas + n-alkane) Binary Mixtures, J. Chem. Thermodynamics (2016), doi: http://dx.doi.org/10.1016/j.jct.2015.12.036 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Measurement and Modelling of High Pressure Density and Interfacial Tension of (Gas + n-alkane) Binary Mixtures Luís M.C. Pereira, Antonin Chapoy, Rod Burgass, Bahman Tohidi * Institute of Petroleum Engineering, Heriot-Watt University, EH14 4AS Edinburgh, UK * To whom correspondence should be addressed: e-mail: bahman.tohidi@pet.hw.ac.uk Phone: +44(0) 1314513672 1

Abstract The deployment of more efficient and economical extraction methods and processing facilities of Oil and Gas requires the accurate knowledge of the interfacial tension (IFT) of fluid phases in contact. In this work, the capillary constant a 2 of binary mixtures containing n-decane and common gases such as carbon dioxide, methane and nitrogen was measured. Experimental measurements were carried at four temperatures (313, 343, 393 and 442 K) and pressures up 69 MPa, or near the complete vaporization of the organic phase into the gas-rich phase. To determine accurate IFT values, the capillary constants were combined with saturated phase density data measured with an Anton Paar densitometer and correlated with a model based on the Peng-Robinson 1978 equation of state (PR78 EoS). Correlated density showed an overall percentage absolute deviation (%AAD) to the measured data of (0.2 to 0.5) % for the liquid phase and (1.5 to 2.5) % for the vapour phase of the studied systems and P-T conditions. The predictive capability of models to accurately describe both the temperature and pressure dependence of the saturated phase density and IFT of 16 (gas + n-alkane) binary mixtures was assessed in this work by comparison with data gathered from the literature and measured in this work. The IFT models considered include the Parachor, the Linear Gradient Theory (LGT) and the Density Gradient Theory (DGT) approaches combined with the Volume-Translated Predictive Peng-Robinson 1978 EoS (VT-PPR78 EoS). With no adjustable parameters, the VT-PPR78 EoS allowed a good description of both solubility and volumetric properties of the mixtures measured in this work, with deviations to measured density data only slightly higher than those of correlated data. The best IFT predictions were obtained with the DGT method with an overall %AAD between (4.9 and 8.3) % for all systems considered and IFT data no lower than 1.5 mn.m -1. Furthermore, the impact of the relative adsorption of gas molecules in the interfacial region on the IFT was further investigated with the DGT. Keywords: interfacial tension, saturated density, VT-PPR78 EoS, Parachor method, linear gradient theory; density gradient theory. 2

1. Introduction Accurate knowledge of the interfacial tension (IFT) between adjacent phases is paramount in multiphase systems. It is used to identify and describe the behaviour and characteristics of interfacial phenomena and surface chemistry involved in numerous engineering applications throughout the chemical and petroleum industry [1 3]. This property is defined as the adhesive force (tension) exerted at the interface between two phases. This force keeps the fluids together and it is generally expressed as the force at the interface per unit length in units of dyne.cm -1 or mn.m -1. In many chemical engineering processes, such as distillation, adsorption and extraction, to name just a few, the interfacial tension is decisive in the control of mass and heat transfer between fluids and hence, it greatly influences the design of processes equipment [4,5]. This property also affects the quality of products such as glues, coatings, paints, agrochemicals, drugs and detergents [2]. With respect to the petroleum industry, interface phenomena influences most, if not all, processes involved in the extraction and refining of crude oil, from the optimization of reservoir engineering schemes to the design of petrochemical facilities. For example, it is well established that several rock properties such as wettability, capillary pressure and relative permeabilities strongly depend on IFT. As a result, the IFT is a key parameter which determines the distribution of hydrocarbons in the pore spaces of a reservoir rock, and in turn, the amount of oil produced [1,3]. Several studies have shown that lowering the IFT between the displacing and in-place fluids in enhanced oil recovery (EOR) processes reduces the capillary forces. This reduction improves the mobility of hydrocarbon reservoir fluids resulting in higher recovery efficiencies [6 12]. Surfactant flooding schemes and miscible gas injection processes are among the most commonly used EOR techniques. The latter is of particular interest for the present work. In a miscible gas injection process, the reduction of the oil-gas IFT is achieved as a result of enhanced mutual miscibility between the phases at high pressures. In general, the injected gas is rich in carbon dioxide and may contain other common gases such as methane and nitrogen. Wagner and Leach s [13] measurements confirmed that the residual oil saturation is significantly influenced by the variation of the IFT in methane-n-pentane systems. Other studies [14 16] revealed that liquid loss to the formation in gas condensate 3

reservoirs can be considerably reduced by maintaining flow rates at low IFT values (<0.1 mn.m -1 ). Knowledge of the interfacial tension is also needed to obtain more realistic representations of the slip between phases in dynamic multiphase flow simulations [17]. All aspects being considered, accurate estimation of the effect of pressure, temperature and composition on the IFT is necessary for the design and/or optimization of production and recovery processes of petroleum technology. The IFT of reservoir fluids is commonly measured by the Capillary Rise (CR) and Pendant Drop (PD) techniques. The latter was used in our previous work [18] for measuring the carbon dioxide/water IFT over a broad range of conditions, with measured IFT values ranging from (68.52 to 12.65) mn.m -1. In the context of the PD technique, a water droplet was formed at the tip of a capillary tube in a high pressure cell and kept in equilibrium in a CO 2 atmosphere. The Asymetric Drop Shape Analysis (ADSA) method of pendant drops was then used together with saturated density data to calculate the interfacial tension values. The diameter of the capillary tube can be used to control the drop size. Accurate measurement of IFT with the PD method requires a sufficiently large density difference between the phases to elongate the drop, reducing the errors that arise from the determination of the droplet characteristic dimensions [19,20]. For this reason, the CR technique is in general preferred for systems near critical conditions, where the density difference and the IFT values are considerably lower. The CR method relies on the fundamentals of capillarity and on the rise of fluids in capillary tubes, which in turn, depends upon the interfacial tension. In other words, the IFT is calculated by relating the column of liquid above the flat liquid surface and the pressure gradients that induce the fluids to move inside the capillary tube. This method has been widely used in the literature to measure the surface tension (ST) of pure substances and interfacial tension of multicomponent mixtures of interest to the petroleum industry. Of particular interest for this work is the study of Weinaug and Katz [21] who investigated the interfacial tension of methane-n-propane mixtures with this technique for temperatures up to 363.2 K and pressures ranging from 10.34 MPa down to ambient, with measured IFT values ranging from (12.13 to 0.50) mn.m -1. Swartz [22] performed interfacial tension measurements of gas saturated crude oil at 304 K and values ranged from (28 to 4) mn.m -1. More recently, Baidakov and co-workers [23] (and references within) measured the interfacial tension in cryogenic and light hydrocarbon systems via the capillary constant (a 2 ) 4

determined from capillary tubes with different internal diameters. Overall, this method is simple and accurate for application in systems at/or near critical conditions, where the IFT is low, and it can be easily implemented at high pressure and temperatures conditions. Another important aspect to consider is the density of the saturated phases used in the calculation of the IFT values. For systems with relatively low mutual solubility such as (hydrocarbon + water) and (nitrogen + water) mixtures, the density of the saturated phases has been commonly approximated to that of pure substances [24 27]. However, for more miscible systems such as (carbon dioxide + water) [18,28] and (carbon dioxide + n-alkanes) [29], studies have shown that these approximations can lead to inaccurate IFT values in particular at high pressures. Hence for these cases, a more rigorous approach consists on assessing the dissolution of the phases into each other and its effect on the density through integrated experimental measurements or estimated with models and correlations. A brief literature review carried by Nourozieh et al. [30] showed that experimental saturated density data of the carbon dioxide (CO 2 ) and n-decane system are still scarce, with most studies limited to the saturated liquid phase and a temperature of 344 K. The most wideranging study was reported by Reamer and Sage [31] who measured the volumetric properties of this system for temperatures over the range (278 to 511) K and at pressures up to 69 MPa. Nagarajan and Robinson [32], Shaver et al. [33] and Mejía et al. [34] have also reported saturated phase densities of this system for temperatures up to 377 K and pressures up to near the critical point. Saturated density data of methane (CH 4 ) and nitrogen (N 2 ) with n-decane are also, to the best of our knowledge, still limited to the works of Sage and coworkers [35,36], Amin et al. [37] and Jianhua et al. [38]. As a result, the experimental investigation of the IFT in these systems is also, in most cases, limited to the studies listed above. Only recently, Georgiadis et al. [39] extended the IFT measurements between CO 2 and n-alkanes, including n-decane, to a broader range of temperatures. Using the PD method, the authors measured the IFT of (CO 2 + n-decane) at temperatures ranging from (298 to 443) K and pressures up to 15 MPa. However, reported IFT values were obtained using pure compound densities for the bulk phases, which can lead to an under/over estimation of the IFT, as mentioned above and indicated also by the authors in their original work. 5

This communication aims at filling the experimental gap and extending the IFT and saturated density measurements of CO 2, CH 4 and N 2 with n-decane. Measurements were carried out at temperatures over the range (313-442) K and pressures up to 69 MPa, and both the equipment and methodology were validated through comparison with selected experimental data available in the literature [31 40]. The densities of the saturated phases were measured with a high pressure densitometer and the data correlated with a thermodynamic phase behaviour model based on the Peng Robinson 1978 equation of state (PR78 EoS) [41]. The differential capillary rise technique was used for measuring the capillary constant a 2 and combined with modelled density data to calculate the IFT values. Furthermore, the predictive capabilities of theoretical approaches such as the Parachor [42,43], the Density Gradient Theory (DGT) [44,45] and the Linear Gradient Theory (LGT) [46,47] were assessed against IFT data measured in this work and available in the literature [21,29,38,48 54]. Vapour-liquid equilibria (VLE) of the mixtures were predicted with the PPR78 model (Predictive 1978, Peng-Robinson EoS) developed by Jaubert and co-workers [55 59]. Density predictions using this EoS were improved by using the concept of volume translation for classical cubic EoS [60,61] and the volume corrections calculated with the correlation developed by Miqueu et al. [62]. As it will be shown, the overall lowest deviations to measured IFT data were provided with the DGT approach. 6

2. Experimental Section 2.1. Materials The specification and sources of the chemicals used in this work are summarized in Table 1. Toluene and n-heptane were used in this work for cleaning purpose only. All chemicals were used without further purification. Table 1. Suppliers and specification as stated by the supplier of the materials used in this study. 2.2. Apparatus Chemical Name Supplier Mass fraction purity Carbon dioxide BOC 0.99995 Methane BOC 0.99995 Nitrogen Air Products 0.9999 n-decane Acros Organic >0.99 n-heptane RathBurn Chemicals >0.99 Toluene Fischer Scientific >0.995 The IFT and equilibrium densities of (gas + n-decane) mixtures were measured with the experimental facility illustrated schematically in Figure 1. It consists of a 500 cm 3 highpressure see-through windowed equilibrium cell (A) with two fluid ports (A 1 and A 2 ), a U- shape vibrating-tube densitometer (B) Model DMA HPM and evaluation unit (C) Model mpds 5, both manufactured by Anton Paar, a 100 cm 3 small movable piston (D), two check valves (E 1 and E 2 ) Model 720.4631, manufactured by Sitec with an opening pressure between (0.01 and 0.02) MPa, two 300 cm 3 sample pistons (F 1 and F 2 ) and an automatic high pressure positive displacement DBR pump system (G) with a maximum capacity of 500 cm 3. 7

Figure 1. Illustrative scheme of the experimental facility based on the CR method. In the scheme the following are annotated: high-pressure see-through windowed equilibrium cell (A); fluid ports (A 1 and A 2 ); U-shape vibrating-tube densitometer (B) and evaluation unit (C); 100 cm 3 movable piston (D); check valves (E 1 and E 2 ); 300 cm 3 sample pistons (F 1 and F 2 ) and automatic high pressure DBR pump system (G). The high-pressure equilibrium cell (A) is designed to hold pressures up to 103.4 MPa and temperatures up to 473 K. A magnetic stirrer positioned at the bottom of the cell helps to minimize potential temperature gradients within the sample in the cell and provides a good homogenization of the system. Two capillary glass tubes, with inner diameter (i.d.) of 0.396 and 0.986 mm, are positioned inside the cell by means of a custom-build stainless steel circular holder. To determine the i.d., several pictures of the capillary tubes were taken with an USB camera fitted with magnifying lenses and the images analysed in a computer screen. The magnification ratio and a reference length were used to calculate the i.d. of the tubes from each picture and the values averaged accordingly. Taking into account all uncertainties and from dispersion of the calculated values, we estimated a standard uncertainty of ± 0.003 mm for the i.d. of the tubes. The fluid port A 1 of the cell is connected to the densitometer (B) by means of 0.29 mm i.d. high-pressure tubing with measured length of 3 m which ensures a minimum volume 8

of approximately 0.2 cm 3. Two check valves (E 1 and E 2 ) and a movable piston (D) controlled by an automatic DBR pump (G) enables a closed loop flow of fluids between the equilibrium cell and the densitometer. Turning the equilibrium cell upside down allows the density measurement of either the saturated liquid or vapour phase. All parts described before are housed inside an oven, manufactured by Cincinnati Sub-Zero Model Z-16, with an overall temperature control stability of ± 0.1 K. The pressure was monitored by means of a pressure transducer (Quartzdyne Series I) connected to valve V 1 and previously calibrated against a dead weight pressure balance. This calibration procedures ensures a standard uncertainty of u(p) = 0.04 MPa. The temperature was measured by a high precision built-in thermometer in the densitometer with a specified standard uncertainty of u(t) = 0.1 K. The temperature readings from the densitometer were checked against a Prema 3040 high precision thermometer and deviations were observed to be within the densitometer uncertainties. Methane, nitrogen, carbon dioxide and degassed n- decane are stored in 300 cm 3 sample pistons and loaded into equilibrium cell through valves V 1, V 2 and V 3 and fluid port A 1. The back pressure in both the sample pistons (F 1 and F 2 ) and movable piston (D) are controlled by means of a DBR pump system (G) and valves V 4 and V 5. An image capturing system (not shown in Figure 1) was used to measure the height of the liquid rise in the capillaries. It consists of a camera (Cohu Model DSP 3600) mounted in a structure capable of vertical displacement (cathetometer). High magnification lenses are fitted into the camera and images with a magnification factor higher than 15x are displayed in a computer screen. As an example, a picture taken with this setup is depicted in Figure 2. Figure 2. Picture obtained with the image capturing system of the liquid rise in the capillary tubes for the (CO 2 + n-decane) mixture. 9

Once the difference of the liquid rise heights in the capillaries h=h 1 -h 2 is known, the capillary constant a 2 and the IFT are calculated by: a 2 h = (1) 1 b 1 b 1 2 2 L V a ( ρ ρ )g IFT = (2) 2 where b 1 and b 2 correspond to the radii of curvature of the two capillary tubes determined in this work by solving the Lane equation [63], ρ L and ρ V correspond to the density of the coexisting saturated liquid and vapour phases, respectively, and g is the gravitational acceleration (9.81589 m.s -2 ). 2.3. Experimental Procedure The entire capillary rise apparatus was thoroughly cleaned with toluene and n-heptane and then dried with compressed air. Once the entire apparatus had been tested for leaks, it was kept under vacuum at 333 K. This cleaning procedure was repeated before loading a new mixture into the system and helped to remove surface active impurities. The desired temperature was set and let to stabilize overnight. For each binary mixture, sufficient n-decane was transferred from the sample piston to the equilibrium cell through port A 1 of the equilibrium cell as depicted in Figure 1. Sufficient liquid hydrocarbon was introduced into the system until (1/4 to 1/3) of the length of the capillary glass tubes was covered. The desired pressure inside the cell was set by injecting pure gas (i.e., methane, carbon dioxide or nitrogen) into the equilibrium cell and controlling the back pressure of the sample pistons. For each pressure/temperature state, valves V 1, V 4 and V 5 were kept closed and the content of the cell was stirred until pressure and temperature readings were stabilized. Once equilibrium was achieved, normally after 10 min, stirring was stopped and the liquid rise in each capillary tube was measured. The saturated liquid density was measured by displacing the phase in the equilibrium cell into the densitometer whilst stirring. Opening valve V 5 and controlling the movement of piston D with the aid of the DBR pump, allowed the saturated liquid phase to be displaced from the equilibrium cell into the densitometer, as depicted in Figure 1. The injection and 10

withdrawal of hydraulic fluid behind piston D were performed at a constant rate of 100 cm 3.h - 1 with the control software of the pump. The back-flow of fluids from the densitometer to the equilibrium cell was prevented by means of check valves E 1 and E 2. Generally 4-5 cycles (~5 cm 3 /each) allowed a complete renewal of the fluids inside the densitometer with a total pressure change in the system of not more than 0.1 MPa at each stroke, ensuring that the equilibrium conditions were maintained. Following this, the cycling was stopped, valve V 5 closed and the resonance period of the vibrating tube (τ) recorded after a 5 minutes period of stabilization. As followed in a previous study [18], the density (ρ) of the liquid phase was computed by: ρ = Dτ D (3) 2 1 2 where D 1 and D 2 are the temperature and pressure dependent densitometer parameters. Similarly, the density of the saturated vapour phase was measured by turning the cell upside down and repeating the procedure described above. The parameters D 1 and D 2 were calibrated against the density of carbon dioxide, methane, nitrogen and n-decane calculated from correlations given in REFPROP [64] for each P-T of interest. To check our calibration procedure, the density of the reference fluids was calculated at pertinent conditions with Eq. 3 and the percentage average absolute deviation (%AAD) to data [64] was found to be (0.26, 0.57, 0.49 and 0.05) %, for carbon dioxide, methane, nitrogen and n-decane respectively. Since gases are in general much more compressible and more sensitive to temperature and pressure changes, uncertainties in pressure (u(p) = 0.04 MPa) and temperature (u(t) = 0.1 K) can introduce significant variations on the density measurements. Furthermore, at low pressures it was difficult to obtain a highly reliable and stable reading of τ when the vapour phase in the binary mixture was circulated through the densitometer. This might be due to the retention of micro-droplets of the liquid phase in the densitometer which could not be completely displaced by the flow of the vapour phase alone. Hence, fewer points were obtained for the density of the vapour phase of the studied systems when compared to the liquid phase. Similar difficulties were highlighted by Chiquet et al. [28] when measuring the saturated density of the (CO 2 + H 2 O) system. Overall, taking into account the uncertainties in pressure and temperature measurements and from observation of dispersion of the density measurements, we estimate a 11

combined expanded uncertainty of U c (ρ) = 1.5 kg.m -3 and U c (ρ) = 7.0 kg.m -3 for the density measurements of the liquid and vapour phases respectively, with a level of confidence of 0.95. The generated densities were correlated with the PR78 EoS described in Appendix A. The adjustment of binary interaction parameters and the volume-translation technique within cubic EoSs [60,61] were employed for a correct description of the generated data. Originally proposed by Martin [60] and later developed by Peneloux et al. [61], the volume-translation concept aimed at improving the density predictions of classical two-parameter EoS, in particular the density of the saturated liquid phase. The method involves the introduction of a translation factor (volume correction) for each component within a mixture without any impact on the calculation of the equilibrium phase compositions. Optimized binary interaction parameters and volume corrections used for each system and temperature are listed in Table S1 in supporting information together with the calculated %AAD to experimental data. Finally, IFT values were calculated with Eq. 2 using the measured capillary constants a 2 and correlated density data at pertinent P-T conditions. 12

3. Model 3.1. Parachor The Parachor model [42,43] and its derivatives are indeed the most successful and widely used approaches in the petroleum industry due to its simplicity and good description of the IFT of both simple and complex reservoir mixtures at high pressure and temperature conditions. In the Parachor approach, the interfacial tension can be correlated to the bulk density difference of the phases in equilibrium by the equation proposed by Macleod [42] and Sugden [43] for pure substances: IFT = P ( ρ ρ ) (4) 1/ E L V ch where P ch is the Parachor value, ρ L and ρ V are the liquid and vapour molar density of the coexisting saturated phases and E is the scaling exponent. Using the simple molar averaging technique, Weinaug and Katz [21] extended this model to mixtures by: N comp 1/E L IFT = Pch, i i ρ yi i= 1 V (x ρ ) (5) where x i and y i are the equilibrium mole fractions of component i in the liquid and vapour phases, respectively, and P ch,i is the Parachor of component i. A considerable volume of research and review works show in detail modifications for the scaling exponent, with values ranging from 3.45 to 4, and correlations for expressing the Parachor values as function of properties such as molecular weight, specific gravity or critical properties. Reviewed modifications can be found in the work of Ali [65]. In this work, we set E equal to 4 and the Parachor values of pure substance were calculated with the correlation proposed by Fanchi [66]. 3.2. Density Gradient Theory The DGT is a more rigorous theoretical approach which has gradually been shown to be a good tool for predicting interfacial properties of a wide class of systems and interfaces. The reader is referred to previous studies [18,67] and references within for more details. In summary, the DGT approach is based on the use of phase equilibria properties of bulk phases 13

separated by a well-defined interface to compute the density distribution of each component across the interface and interfacial tension values. This theory has its origins in the work of van der Waals for inhomogeneous fluids [68], but only after the reformulation made by Cahn and Hilliard [44], it found its widespread use in the modelling of interfacial tension. Accordingly, the IFT between two equilibrated phases within the DGT framework is computed by: + + d ρ d ρ i j IFT = cij dz = 2 Ω( ρ) dz dz dz (6) i 0 j Ω ( ρ) = f ( ρ) ρ µ + Eq. i i p (7) i where Ω is the variation of the grand thermodynamic potential, f 0 is the Helmholtz free energy density of the homogeneous fluid at local density ρ, Eq. µ i are the pure component chemical potentials evaluated at the phase equilibrium conditions of the bulk phases, c ij is the cross influence parameter and p is the equilibrium pressure. The cross influence parameters are related to pure component influence parameters c ii and c jj by the mixing rule: c = ( 1 β ) c c (8) ij ij ii jj where β ij is the binary interaction coefficient. When β ij is set to zero, Eq. 8 is reduced to the simple geometric mixing rule and this makes the calculation of the interfacial tension fully predictive. Even though the influence parameters of pure substances c ii and c jj can be derived from theoretical expressions, its application to practical systems can be rather complicated as it requires the prior calculation of the radial density distribution function of a pure fluid in the homogeneous state [69 71]. To overcome this constraint, the parameters are commonly correlated from surface tension data. For practical purposes, several semi-empirical approaches have been developed and various EoS-dependent correlations have been reported in the literature [62,72 74] for the estimation of the influence parameter. This parameter increases slowly with increasing temperature but rapidly diverges near the critical point [62]. For this reason most approaches have considered this temperature dependency when combining the DGT with several EoSs. Among all, Miqueu et al. [62] derived a simple linear 14

correlation for the influence parameter of hydrocarbons, gases and refrigerants as a function of reduced temperatures (T r =T/T c ) when using a volume-translated version of the original PR EoS. Due to its simplicity and applicability to the systems investigated, this expression was adopted for the estimation of the influence parameter of the substances in the present work. The expression has the following form [62]: A ii 16 10 = 1.2326 + 1.3757ω ii (9) B ii 16 10 = 0.9051+ 1.5410ω ii (10) T c = A 1 + B a b 2/3 ii i i ii ii T c, ii (11) where T c,ii and ω ii correspond to the critical temperature and acentric factor of component ii, respectively, and a ii and b ii are the energy and co-volume parameters of the PR78 EoS. In order to compute interfacial tension values with Eq. 6, the density distribution of each component (dρ i /dz) needs to be determined. The profile of each component can be calculated by applying the minimization criterion of the Helmholtz free energy to planar interfaces and solving the set of N comp non-linear differential equations in a finite domain [0, L] defined as: 2 d ρ j Eq. cij = µ ( 2 i ρ1(z),..., ρ N (z)) µ i for i,j = 1 N comp (12) dz j with ρ i (z=0) = bulk vapour = ρ V L i and ρ i (z=l) =bulk liquid = ρ i In a recently published work [67], the system of non-linear equations was solved by casting the solution of the density profiles for a given length of the interface L and subsequent enlargement until the computed IFT variation was lower than 0.1 %. This procedure allows the computation of the density profiles in a mixture without any prior assumption on the distribution of the components across the interface. However, taking c ij to be equal to the geometric mean greatly simplifies Eq. 12 and the problem to be solved is reduced to a system of algebraic equations. For a binary mixture the problem is defined as: 15

c ( µ ( ρ, ρ ) µ ) = c ( µ ( ρ, ρ ) µ ) (13) Eq. Eq. 11 2 1 2 2 22 1 1 2 1 It is clear from the previous simplification that the density profile of each component can be computed as roots of the algebraic equation Eq. 13, i.e., the density of one component (dependent variable) can be described as function of the density of the other one, in which the latter is used as an independent (reference) variable in a Newton-Raphson root finding scheme. The selection of the reference component density needs to be done with care, as it needs to be a monotonic function of z over the entire interfacial region [75 79]. Since it is not known beforehand whether or not the density of a component has a monotonic behaviour, and in order to avoid problems in finding the roots of Eq. 13, several authors [75 77] suggested that the selection of the independent variable can be alternated at each calculation step and the root finding calculation continued over the domain in which the density of the reference component is monotonically changing with z. As an example, this means that if one chooses ρ 1 as reference, a turning point would be foreseen by the approach of the value of dρ 1 /dρ 2 to zero for each increment of ρ 1 and at this point one would merely change the reference variable to ρ 2. More recently, Miqueu et al. [79 81] concluded that for nonassociating compounds and in vapour-liquid interfaces, the criteria for the selection of the reference variable can be done on the basis that the components with the lowest surface tension would have a tendency to accumulate in the interfacial region, while the density of the less volatile components would pass monotonically from the vapour to the liquid phase and hence, this component can be selected as reference variable. This method has been validated in binary and multicomponent hydrocarbon mixtures containing methane, nitrogen and carbon dioxide [73,79 81] and therefore, it was also adopted in this work. A detailed description of the method used for casting the roots of Eq. 13 can be found elsewhere [77 79]. Once the density profiles are known, the interfacial tension with the DGT can be calculated by solving the following integral over the density of the reference component ρ 1 and given by: L ρ 2 1 dρ 2 dρ 2 IFT = 2 Ω ( ρ1, ρ2) c22 + 2c12 + c11 d ρ1 (14) V d ρ1 dρ ρ 1 1 16

Moreover, the coordinates in density space can be transformed to location space and the density profiles represented in the z domain by: 2 dρ dρ c + 2c + c 2 2 22 12 ρ 11 1 0 d ρ1 d ρ1 1 = 0 1 + 1 0 2 Ω( ρ1, ρ2) ρ1 z( ρ ) z ( ρ ) d ρ (15) In addition to computing IFT values, the distribution of species across the interface can also be used to compute the local adsorption of molecules in the interface. Accordingly, the adsorption of species i on species j (Γ ij ) can be calculated as defined by Gibbs from the following expression [82]: Γ 12 = 1 C( z) dz (16) α + where C(z) is the symmetrical interface segregation at symmetrical concentrations α i defined as [83,84]: L L ρ2( z) ρ2 ρ1( z) ρ1 C(z) = (17) α α 2 1 α ρ ρ ( ) ( ) L V i i i = L L V V ρ1 + ρ2 ρ1 + ρ2 (18) Complementarily, the adsorption isotherms can also be computed directly from experimental IFT data and bulk phase density by [85]: ρ ρ ρ ρ IFT V L L V 1 2 1 2 Γ 12 = L V ( ρ2 ρ2 ) P T (19) where the values Γ 12 can be directly compared with those predicted with Eq. 16 and the density profiles calculated with the DGT. In Eq. 16 and 19, Γ 12 corresponds to the adsorption of component 1 relative to component 2, and hence the adsorption of component 2 is assumed to be zero. 17

3.3. Linear Gradient Theory To overcome the need for solving computational-demanding density profile equations and to speed up IFT calculations, Zuo and Stanby [46,47] developed the LGT. In the LGT framework, the density of each component across the interface is assumed to be linearly distributed and thus, the density distribution of each species can be readily calculated by: dρi ( z) Di = (20) dz D i L V ( ρi ρi ) = (21) L where D i is a constant for each component i calculated from the density of the coexisting liquid ρ L and vapour ρ V phases separated by an interface of thickness L. With this assumption, the numerical effort inherent to the resolution of a set of equations for finding the density distribution of each component across the interface is eliminated. Based on the linear distribution of densities across the interface, Zuo and Stanby [46] directly derived the influence parameter of the mixture c from the original DGT: c = ρ N comp N comp ρi j cij i j ρ1 ρ (22) 1 where ρ i is the density difference of component i between the coexisting vapour and liquid phases and cij is the cross influence parameter calculated with Eq. 8. However, Zuo and Stanby [47] concluded that the mixing rule in Eq. 22 is not suitable for high pressures and hence, suggested the following expression based on the classical van der Waals mixing rule for the energy parameter in cubic EoS: Ncomp Ncomp c = c x x (23) i j ij i j where x i and x j correspond to the molar fraction of components i and j in the liquid phase and cij is the cross influence parameter calculated with Eq. 8. According to the LGT, the IFT can be computed similarly to the DGT by: 18

L ρ1 2 ( ρ1, ρ2) ρ1 (24) V ρ1 IFT = c Ω d The LGT has been successfully applied in the modelling of the IFT of both weak and strong associating fluids and complex mixtures [46,47,86 91]. However, the correct description of the IFT of several systems required the use of temperature dependent binary interaction coefficients β ij adjusted against experimental mixture data. This is particularly evident in systems such as (CO 2 + hydrocarbon) [47] and (gas + water) [89 92] where the adsorption of molecules in the interface plays a key role on the IFT and thus, the linear distribution of the density profiles, and in turn, the mixing rules for the influence parameter (Eq.22 and Eq.23), are incapable of accurately describing the adsorption phenomena in the interfacial region. A complementary investigation of these results is also carried in the present work. In summary, in this work we compare the predictive capabilities of the DGT and LGT models by setting β ij in Eq. 8 equal to zero and assessing the results against data measured in this work and from the literature, as well as with the predictions from the Parachor method (Eq. 5). 3.4. Phase equilibrium model The procedure for modelling interfacial tension with the IFT models described above requires first an accurate description of the equilibrium properties of the two phases separated by the interface being considered at a given temperature and pressure. These properties were calculated by applying the criterion of equality of the chemical potentials of each component in the coexisting equilibrium phases [93] in combination with the Predictive Peng-Robinson 1978 (PPR78) EoS developed by Jaubert and co-workers [55 59]. The PPR78 EoS [55 59] relies on the original PR EoS [41] as described in Appendix A and therefore, the simplicity and robustness of the model remains unchanged. However, unlike the classical PR EoS [41], where binary interaction parameters k ij are usually adjusted against VLE data for an accurate description of bulk phase compositions, in the PPR78 model these parameters are predicted for each temperature using a group contribution concept. A total of 21 different groups have been established and validated by the authors against an 19

extensive database of VLE measurements. In summary, the PPR78 model can be used for estimating the k ij of mixtures involving hydrocarbons, water, mercaptans and common gases (CO 2, N 2, H 2 S and H 2 ) and accurately predicting the VLE over a wide range of experimental conditions. In order to obtain the best prediction possible for the density of the saturated phases, the volume-translation method was also used. Unlike the procedure followed in the Experimental Section of this work, volume corrections were estimated from methods available in the literature. Presently, several volume correction methods have been presented for cubic EoS, with approaches including the use of a constant correction term or temperature and density dependent correlations. For further discussion of these methods the reader is referred to the works of Abudour et al. [94,95] Considering all approaches, in this work we chose to use the expression derived by Miqueu et al. [62] for calculating the volume corrections of all systems investigated. This correction was chosen because of its simplicity and applicability to a wide range of substances, including the ones in this work. To summarize, the bulk equilibrium properties (VLE and densities) were computed in a fully predictive manner with the volume-translated PPR78 (VT-PPR78) EoS and used in the IFT models. The model was also used to evaluate Ω within the DGT (Eq. 14) and LGT (Eq. 24) framework from fundamental thermodynamic relationships [73,89]. 20

4. Results and Discussion 4.1. Experimental Measurements 4.1.1. Equilibrium densities The saturated densities of methane, nitrogen and carbon dioxide with n-decane were measured at temperatures of (313, 344, 393 and 443) K and pressures up to 68 MPa or near the complete vaporization of the liquid hydrocarbon and the results are presented in Tables S2 through S4 in supporting information. Measured densities were also compared to selected literature data [31 36,38] and to correlated density and the results are plotted in Figure 3. As can be seen in Figure 3, our density measurements are in quite good agreement with the data of Sage and co-workers [31,35,36], Nagarajan and Robinson [32], Shaver et al. [33], Mejía et al. [34] and Jianhua et al. [38] within the range studied. The largest deviation were observed between our measurements and those reported by Sage and co-workers [31] for the density of the saturated vapour phase at high pressures and low temperatures. Indeed, despite the excellent agreement with the density of the saturated liquid phase, results from Sage and co-workers [31] at 313 K suggest a lower density for the vapour phase of (CO 2 + n- decane) with increasing pressures. It is worthy of note that no comparison was performed with the data measured by Amin et al. [37] for the saturated density of (CH 4 + n-decane) as quantitative results were not published in their original work nor made available at the time of this study. The PR78 EoS with binary interaction parameters and volume corrections listed in Table S1 in supporting information was capable of correlating the saturated density of the studied systems with relatively low deviations to the measured data. As listed in Table S1 and depicted in Figure 3, the highest deviations were obtained for the density of the vapour phases, with a maximum %AAD of 3.6% obtained for the (CH 4 + n-decane) system at the highest temperature. This approach also allowed correct description of the density of the vapour phase at low pressures, where a limited number of points were measured, as illustrated in Figure 3. Altogether, the PR78 EoS with adjusted parameters allowed a reproduction of the saturated density with an overall %AAD to measured data between (0.2 to 0.5) % in the liquid phase and between (1.5 to 2.5) % in the vapour phase. These results 21

validate the adequacy of our method for estimating the density difference of the systems studied herein. 50 (a) 40 P / MPa 30 20 10 20 0 0 200 400 600 800 ρ/ kg.m -3 (b) 15 P / MPa 10 5 P / MPa 0 0 200 400 600 800 ρ/ kg.m -3 70 (c) 60 50 40 30 20 10 0 0 200 400 600 800 ρ/ kg.m -3 Figure 3. Saturated density-pressure diagrams of CH 4 (a), CO 2 (b) and N 2 (c) with n-decane. Full symbols represent experimental data measured in this work: T = 313 K ( ), T = 343 K ( ), T = 393 K ( ) and T = 442 K ( ). Empty symbols represent literature data: (a) Reamer and Sage [36], T = 311 K ( ), T = 344 K ( ) and T = 444 K ( ); Sage et al. [35], T = 394 K ( ); (b) Reamer and Sage [31], T = 311 K ( ), T = 344 K ( ) and T = 444 K ( ); Nagarajan and Robinson [32], T = 344 K ( ); Shaver et al. [33], T = 344 K ( ); Mejía et al. [34], T =344 K ( ); (c) Jianhua et al. [38], T = 313 K ( ). Solid lines represent correlated densities obtained with the PR78 EoS and parameters listed in Table S1 in supporting information at pertinent temperatures. 22

4.1.2. Interfacial tension The capillary constant a 2 of three binary mixtures (CH 4 + n-decane), (CO 2 + n- decane) and (N 2 + n-decane) were measured in the temperature range (313-442) K and pressures up 69 MPa and the results combined with correlated density difference to estimate the IFT. Capillary constants and IFT values are listed in Tables 2 through 4 for each P-T condition studied. A comparison between our measurements and data in the literature [32 34,37 40] is shown in Figure 4. As depicted in Figure 4, our measurements are in quite good agreement with the measurements of Amin et al. [37] and Stegemeier et al. [40] for (CH 4 + n-decane), Nagarajan and Robinson [32], Shaver et al. [33], Georgiadis et al. [39] and Mejía et al. [34] for (CO 2 + n-decane) and Jianhua et al. [38] for (N 2 + n-decane). In general, our measurements show that the IFT of the studied systems are slightly lower than data reported by others in the literature. The largest deviations were obtained between our measurements and those from Amin et al. [37] at low pressures for (CH 4 + n-decane). However, data points from Amin et al.[37] at low pressures have been brought to question as the slope of their results suggest an IFT for this system at ambient pressure significantly higher than surface tension of n-decane at pertinent temperatures. As an example, we have plotted in Figure 4 (b) their measurements at 311 K (n-decane ST = 22.18 mn.m -1 at 311 K [64]). Moreover, it is worth mentioning that data measured by Georgiadis et al. [39] for (CO 2 + n-decane) at 344 K and 443 K were recalculated using the approach adopted in this work for the density difference between the equilibrated phases and plotted in Figure 4 (c). It was found that the absolute difference between the recalculated and the original IFT data [39] had an average of 0.18 mn.m -1 (%AAD = 3.7 %) and reached a maximum of 0.48 mn.m -1 (%AAD = 17.7%) at the highest temperature and pressure, where the density difference between the equilibrated phases reached the lowest value within the range studied. The reader is referred to our previous publication [18] for more details of the impact of the approximation to pure substance density on the IFT values. From Figure 4 it can be seen that the pressure increase leads to a reduction of the interfacial tension values in all systems investigated, more pronounced for the (CO 2 + n- decane) system. On the other hand, the temperature increase leads to a decrease of the IFT in 23

the (CH 4 + n-decane) and (N 2 + n-decane) systems over the entire pressure range, whereas a crossover was observed in the (CO 2 + n-decane) system for pressures ranging from (5 to 10) MPa, where the impact of temperature on the IFT is less marked. In general, the dependence of the IFT on temperature is more pronounced at low pressures and gradually decreases with increasing pressures, as depicted in Figure 4. The combined expanded uncertainties in the interfacial tension measurements, U c (IFT), are assessed by considering the effect of uncertainties on the measured capillary constants a 2, U 1 (IFT), and uncertainties on the estimation of the density of the equilibrated phases, U 2 (IFT). U 1 (IFT) can be estimated by considering the impact of uncertainties of the cathetometer and dispersion of the measured difference of the liquid rise height in the capillary tubes h. We estimated a combined expanded uncertainty for h of U c ( h) = 0.010 cm for measurements ranging from (1.876 to 0.129) cm, with a level of confidence of 0.95. The calculation of the expanded uncertainty of a 2 depends not only of the i.d of the capillary tubes, but also of the magnitude of h, and in turn, of the total pressure of the system. In our measurements, the combined expanded uncertainty of a 2 was found to be U c (a 2 ) = 0.000331 cm 2, with a level of confidence of 0.95, yielding expanded relative uncertainties for this parameter between (0.5 to 0.7) % at low pressures and up to 8 % for pressures close to the complete evaporation of n-decane into the gas-rich phase, where h reached the lowest values. Using the law of propagation of errors [96], U 1 (IFT) can be estimated by: 2 1/2 IFT 2 2 1( ) = ( ) 2 c U IFT U a a 2 2 2 a 2 Uc ( a ) = Uc ( h) h 1/2 (25) (26) The second contribution, U 2 (IFT), can be estimated by considering the deviations between correlated and measured density data for each system, phase and temperature of interest (Table S1 in supporting information). The impact of uncertainties in temperature (u(t) = 0.1 K) and pressure (u(p) = 0.04 MPa) on the estimation of the density of the saturated phases are somewhat suppressed by the deviations of the model to measured data 24

and hence, were not considered. The expression for estimating U 2 (IFT) has the following form [96]: 2 2 ( ) IFT ( L ) 2 IFT ( V ) 2 U2 IFT = U L c ρ + U V c ρ ρ ρ 1/2 (27) The combined expanded uncertainties in the IFT measurements estimated with a confidence level of 0.95, are shown in Tables 2 through 4 for each system and experimental condition, with values averaging (0.13, 0.12 and 0.14) mn.m -1 in the (CO 2 + n-decane), (CH 4 + n-decane) and (N 2 + n-decane) systems, respectively. Overall, the expanded uncertainty of the IFT measurements increased with pressure, with estimated expanded relative uncertainties lower than 1.5% at low pressures and up to 11.3 % near the critical point, where IFT values are very low (<1.5 mn.m -1 ). 25

25 20 (a) 25 20 15 IFT / mn.m -1 15 10 10 5 0 0 5 10 15 20 25 30 35 5 IFT / mn.m -1 0 20 15 10 5 0 5 10 15 20 25 30 35 P / MPa (b) 20 15 10 5 0 0 5 10 15 20 IFT / mn.m -1 0 25 20 15 10 0 5 10 15 20 P / MPa (c) 25 20 15 10 5 0 0 20 40 60 5 0 0 10 20 30 40 50 60 70 80 P / MPa Figure 4. IFT-pressure diagrams of CH 4 (a), CO 2 (b) and N 2 (c) with n-decane. Full symbols represent experimental data measured in this work: T = 313 K ( ), T = 343 K ( ), T = 393 K ( ) and T = 442 K ( ). Empty symbols represent literature data: (a) Stegemeier et al. [40], smoothed data T = 311 K ( ); Amin et al. [37], T = 311 K ( ); (b) Nagarajan and Robinson [32], T = 344 K ( ); Shaver et al. [33], T = 344 K ( ); Georgiadis et al. [39], T = 344 K ( ) and T = 443 K ( ); Mejía et al. [34], T =344 K ( ); (c) Jianhua et al. [38], T = 313 K ( ). Data from Georgiadis et al. [39] were recalculated using the approach adopted in this work for the density difference between the equilibrated phases. 26

Table 2. Measured interfacial tension data of the (CH 4 + n-decane) system. The density difference between equilibrated phases corresponds to correlated data using the PR78 EoS and parameters listed in Table S1 in supporting information. P /MPa a 2 /cm 2 ρ / (kg.m -3 ) IFT / (mn.m -1 ) Experimental Error / (mn.m -1 ) U 1 U 2 U c =U 1 +U 2 T = 313.3 K 0.45 0.055604 716.8 19.56 0.12 0.04 0.16 1.08 0.053354 709.6 18.58 0.12 0.04 0.16 2.97 0.047465 687.4 16.01 0.11 0.03 0.14 5.83 0.040651 652.2 13.01 0.11 0.03 0.14 10.43 0.030826 592.2 8.96 0.10 0.03 0.13 12.46 0.027056 565.0 7.50 0.09 0.03 0.12 15.28 0.022493 526.8 5.81 0.09 0.03 0.12 18.20 0.017667 487.0 4.22 0.08 0.02 0.10 22.04 0.013141 434.3 2.80 0.07 0.02 0.09 28.02 0.006550 349.2 1.12 0.06 0.01 0.07 29.64 0.005008 324.8 0.80 0.05 0.01 0.06 30.50 0.004485 311.5 0.69 0.05 0.01 0.06 T = 343.2 K 0.86 0.049980 687.8 16.87 0.11 0.09 0.20 1.87 0.047895 677.1 15.92 0.11 0.09 0.20 3.62 0.042900 658.1 13.86 0.11 0.08 0.19 7.11 0.035523 618.5 10.78 0.10 0.07 0.17 10.69 0.028974 575.7 8.19 0.09 0.06 0.15 13.97 0.023981 534.9 6.30 0.09 0.05 0.14 17.34 0.019055 491.4 4.60 0.08 0.04 0.12 20.83 0.014462 444.4 3.15 0.07 0.03 0.10 24.43 0.010336 393.0 1.99 0.06 0.02 0.08 29.16 0.005532 318.3 0.86 0.05 0.01 0.06 T = 392.6 K 0.94 0.040320 654.9 12.96 0.11 0.09 0.20 3.85 0.034366 624.8 10.54 0.10 0.07 0.17 7.45 0.028610 585.7 8.22 0.10 0.06 0.16 10.50 0.024874 550.9 6.73 0.09 0.05 0.14 14.08 0.019385 507.9 4.83 0.08 0.04 0.12 17.86 0.015420 459.4 3.48 0.07 0.03 0.10 21.14 0.012118 413.9 2.46 0.07 0.03 0.10 23.99 0.009248 371.1 1.68 0.06 0.02 0.08 28.49 0.005795 293.4 0.83 0.05 0.02 0.07 T = 442.3 K 4.18 0.027188 573.7 7.66 0.09 0.05 0.14 7.18 0.023121 540.9 6.14 0.09 0.05 0.14 11.11 0.018592 494.9 4.52 0.08 0.04 0.12 14.25 0.014825 454.9 3.31 0.07 0.03 0.10 17.43 0.011425 410.9 2.30 0.07 0.02 0.10 20.70 0.007503 360.7 1.33 0.06 0.02 0.08 22.53 0.005729 329.6 0.93 0.05 0.02 0.07 24.63 0.004289 290.2 0.61 0.05 0.01 0.06 u(p) = 0.04 MPa, u(t) = 0.1 K, U c (IFT) calculated with a level of confidence of 0.95. 27

Table 3. Measured interfacial tension data of the (CO 2 + n-decane) system. The density difference between equilibrated phases corresponds to correlated data using the PR78 EoS and parameters listed in Table S1 in supporting information. P / MPa a 2 /cm 2 ρ / (kg.m -3 ) IFT / (mn.m -1 ) Experimental error / (mn.m -1 ) U 1 U 2 U c =U 1 +U 2 T = 313.5 K 3.49 0.040452 669.2 13.29 0.11 0.03 0.14 5.51 0.025601 621.2 7.81 0.10 0.02 0.12 6.71 0.017105 574.1 4.82 0.09 0.02 0.11 6.79 0.015717 569.9 4.40 0.09 0.02 0.11 7.09 0.013273 551.9 3.60 0.09 0.02 0.11 T = 343.2 K 0.99 0.04816 682.4 16.13 0.11 0.17 0.28 3.59 0.03873 640.3 12.17 0.10 0.14 0.24 5.69 0.03007 596.0 8.79 0.10 0.11 0.21 7.02 0.02481 560.6 6.83 0.09 0.10 0.19 8.39 0.01882 514.8 4.76 0.08 0.07 0.15 10.43 0.00991 415.7 2.02 0.07 0.04 0.11 10.84 0.00842 387.9 1.60 0.06 0.04 0.10 11.52 0.00517 331.9 0.84 0.05 0.02 0.07 11.80 0.00403 302.9 0.60 0.05 0.02 0.07 T = 392.7 K 1.21 0.03959 639.3 12.42 0.10 0.04 0.14 4.84 0.03142 584.0 9.01 0.09 0.04 0.13 8.73 0.02190 506.7 5.45 0.08 0.05 0.13 10.81 0.01664 454.2 3.71 0.07 0.04 0.11 12.65 0.01222 398.5 2.39 0.06 0.04 0.10 13.98 0.00938 351.1 1.62 0.06 0.03 0.09 15.47 0.00413 288.1 0.58 0.05 0.02 0.07 T = 442.5 K 3.60 0.02752 564.8 7.63 0.09 0.04 0.13 6.02 0.02342 524.5 6.03 0.09 0.04 0.12 7.86 0.01985 490.2 4.78 0.08 0.03 0.11 10.26 0.01572 439.6 3.39 0.07 0.03 0.10 12.64 0.01166 380.9 2.18 0.06 0.03 0.09 14.44 0.00806 328.7 1.30 0.05 0.02 0.07 u(p) = 0.04 MPa, u(t) = 0.1 K, U c (IFT) calculated with a level of confidence of 0.95. 28

Table 4. Measured interfacial tension data of the (N 2 + n-decane) system. The density difference between equilibrated phases corresponds to correlated data using the PR78 EoS and parameters listed in Table S1 in supporting information. P / MPa a 2 /cm 2 ρ / (kg.m -3 ) IFT / (mn.m -1 ) Experimental Error / (mn.m -1 ) U 1 U 2 U c =U 1 +U 2 T = 313.4 K 1.19 0.061758 712.0 21.58 0.12 0.06 0.18 3.74 0.058648 684.0 19.69 0.11 0.06 0.17 10.54 0.052924 611.7 15.89 0.10 0.08 0.18 14.75 0.050178 569.9 14.04 0.09 0.10 0.19 22.00 0.047366 504.6 11.73 0.08 0.13 0.21 27.98 0.045745 457.1 10.26 0.07 0.15 0.22 34.68 0.043297 410.6 8.72 0.07 0.16 0.23 42.36 0.041875 364.4 7.49 0.06 0.18 0.24 47.33 0.041180 338.1 6.83 0.05 0.19 0.24 51.02 0.040452 320.2 6.36 0.05 0.19 0.24 55.78 0.040055 298.8 5.87 0.05 0.20 0.25 62.17 0.039129 273.0 5.24 0.04 0.21 0.25 T = 343.2 K 0.71 0.056266 689.3 19.03 0.11 0.04 0.15 3.72 0.053454 659.2 17.29 0.11 0.04 0.15 7.32 0.050145 624.1 15.36 0.10 0.05 0.15 14.50 0.045547 558.0 12.47 0.09 0.07 0.16 21.10 0.042437 502.9 10.47 0.08 0.09 0.17 28.29 0.039294 449.3 8.66 0.07 0.10 0.17 35.01 0.036946 404.9 7.34 0.07 0.11 0.18 41.75 0.034663 365.5 6.22 0.06 0.12 0.18 48.85 0.033274 328.7 5.37 0.05 0.13 0.18 57.07 0.031356 291.4 4.48 0.05 0.13 0.18 64.47 0.029603 261.9 3.81 0.04 0.13 0.17 69.00 0.028445 245.6 3.43 0.04 0.13 0.17 T = 392.6 K 0.41 0.046374 653.1 14.86 0.11 0.03 0.14 3.80 0.043231 622.6 13.21 0.10 0.03 0.13 7.48 0.041081 590.7 11.91 0.10 0.03 0.13 14.67 0.036582 531.7 9.55 0.09 0.03 0.12 20.90 0.032976 484.7 7.84 0.08 0.03 0.11 28.27 0.02871 434.1 6.12 0.07 0.03 0.10 34.92 0.025899 392.8 4.99 0.06 0.04 0.10 41.99 0.023948 353.0 4.15 0.06 0.04 0.10 49.27 0.020675 316.1 3.21 0.05 0.03 0.08 55.60 0.018956 287.0 2.67 0.05 0.03 0.08 62.59 0.015949 257.5 2.02 0.04 0.03 0.07 64.93 0.015255 248.3 1.86 0.04 0.03 0.07 29

Table 4. (continued) P / MPa a 2 /cm 2 ρ / (kg.m -3 ) IFT / (mn.m -1 ) Experimental Error / (mn.m -1 ) U 1 U 2 U c =U 1 +U 2 T = 442.2 K 0.90 0.036251 606.3 10.79 0.10 0.01 0.11 4.38 0.033340 576.8 9.44 0.09 0.01 0.10 7.63 0.032050 549.9 8.65 0.09 0.01 0.10 16.35 0.026196 481.3 6.19 0.08 0.01 0.09 21.24 0.023749 445.2 5.19 0.07 0.01 0.08 28.27 0.019947 396.3 3.88 0.06 0.01 0.07 35.48 0.016675 349.2 2.86 0.06 0.01 0.07 45.91 0.011062 285.7 1.55 0.05 0.01 0.06 48.67 0.009874 269.8 1.31 0.04 0.01 0.05 55.67 0.007964 230.4 0.90 0.04 0.01 0.05 u(p) = 0.04 MPa, u(t) = 0.1 K, U c (IFT) calculated with a level of confidence of 0.95. 30

4.2. Modelling The development of predictive and accurate models for calculating equilibrium phase properties and interfacial tension is necessary for the design of more efficient and reliable oil recovery processes and multiphase flow simulations. Hence, in this work we evaluate the predictive capabilities of the VT-PPR78 EoS and IFT models to represent the experimental data of the studied systems. This analysis is also extended to binary mixtures containing CH 4, CO 2 and N 2 and other n-alkanes by comparing the IFT predictions against selected literature data [21,29,38,48 54]. Classical cubic equations of state have been extensively applied in the literature in combination with IFT methods to model the interfacial tension of (gas + n-alkane) systems [46,47,72,76,78 81,86,88,97 99]. Unlike most of these studies, where volume corrections and/or binary interaction parameters within the EoS framework were adjusted to bulk phase property data in order to obtain a good representation of the phase behaviour of the mixtures, here the most reliable bulk phase properties were obtained in a fully predictive manner with the VT-PPR78 model described in Section 3. Discrepancies between measured and predicted bulk phase properties of the studied systems in terms of %AAD are given in Table S1 in supporting information and, as an example, a comparison at 343 K is plotted in Figure 5. The deviations between measured and predicted IFT with the Parachor, LGT and DGT approaches for the studied systems as well as for other (gas + n-alkanes) mixtures are listed in Table 5. A graphical representation of the results for selected systems and conditions is depicted in Figures 6 through 8. As it can be seen in Figure 5, the VT-PPR78 predictions are in quite good agreement with experimental solubility and density data. Moreover, as listed in Table S1 in supporting information, the VT-PPR78 EoS allowed a prediction of the saturated density data measured in this work with an overall %AAD between (1 to 2.9) % for the liquid phase and (3.2 to 4.8)% for the vapour phase. A complementary graphical representation of the predictive capabilities in systems containing other n-alkanes is plotted in Figure S1 in supporting information. Overall, deviations from predicted densities to measured data are only slightly higher than those obtained with correlated density data (Table S1 in supporting information), 31

endorsing the adequacy of the model to predict the effect of pressure and temperature on the phase behaviour of the studied systems. 70 (a) 70 15 (b) 60 60 10 50 50 5 P / MPa 40 30 P / MPa 40 30 0 500 600 700 800 20 20 10 10 0 0 0.2 0.4 0.6 0.8 1 x,y 0 0 200 400 600 800 ρ/ kg.m -3 Figure 5. Pressure-composition (a) and pressure-density (b) diagrams of CH 4 (red), CO 2 (green) and N 2 (black) with n-decane. Symbols in (a) represent solubility data taken from the literature: Reamer et al. [36], T = 344 K ( ), Reamer and Sage [31], T = 344 K ( ) and Azarnoosh and McKetta [100], smoothed data T = 344 K ( ). Symbols in (b) represent density data measured in this work at 343 K. Solid lines represent predictions from VT-PPR78 EoS. Table 5 and Figure 6 show that the DGT approach provided the best predictions of the IFT of n-alkanes systems with methane, carbon dioxide and nitrogen, with an overall %AAD of (8.3, 5.5 and 4.9) %, respectively. In general the highest deviations were observed near the critical point, where IFT values are rather low (IFT < 1.5 mn.m -1 ) and experimental uncertainties considerable. In this region, the overall deviations of all models ranged between (15.9 to 54.9)% and they are somewhat related with the inappropriate description of the phase behaviour by cubic equations of state near the critical point, as depicted in Figure 5. This is characteristic of equations of state without the cross-over approach. Nonetheless, the VT- PPR78 + DGT results are in good agreement with other DGT studies [29,34,51,52,101], in which more complex and molecular-based equations of state, such as the Statistical Association Fluid Theory (SAFT) and its derivatives, are used for the description the bulk saturated properties and the Helmholtz free energy density of the fluid across the interfacial region. It can be seen in Table 5 that the LGT model outperforms in general the Parachor method for systems in methane and nitrogen atmospheres. Overall, the LGT allowed a prediction of the IFT of (CH 4 + n-alkanes) and (N 2 + n-alkanes) systems far from the critical 32

point (i.e., IFT > 1.5 mn.m -1 ) with a %AAD of (10.1 and 7.7) %, respectively, whereas deviations to measured data with the Parachor method were of (14.5 and 13.9) %, respectively. On the other hand, a significant under estimation of the IFT was obtained for the systems containing carbon dioxide with the simplified version of the DGT at moderate pressures, as depicted in Figure 6 (b), resulting in an overall %AAD of 20.3 %, whereas overall deviations with the Parachor method yield only 9.3%. These findings are agreement with the modelling results of Zuo and Stenby [47], in which a large and negative binary interaction parameter for the cross influence parameter was necessary for correct description of the IFT of systems containing CO 2 with the LGT model. 33

20 15 (a) 3.0 2.5 2.0 1.5 IFT / mn.m -1 10 5 1.0 0.5 0.0 2 4 6 8 IFT / mn.m -1 0 25 20 15 10 0 10 20 30 40 P / MPa (b) 25 20 15 10 5 0 0 5 10 15 5 0 0 5 10 15 P / MPa 20 (c) 20 IFT / mn.m -1 15 10 15 10 5 0 0 10 20 30 5 0 0 20 40 60 80 P / MPa Figure 6. IFT-pressure diagrams of CH 4 (a), CO 2 (b) and N 2 (c) and n-alkane systems. Symbols in (a) represent experimental IFT data of CH 4 and: n-propane taken from Weinaug and Katz [21], T = 338 K ( ); n-hexane taken from Nino-Amezquita et al. [52], T = 350 K ( ); n-heptane taken from Jaeger et al. [53], T = 323 K ( ) and n-decane from this work, T = 343 K ( ). Symbols in (b) represent experimental IFT data of CO 2 and: n- butane taken from Hsu et al. [49], T = 344 K ( ); n-heptane taken from Jaeger et al. [50], T = 353 K ( ); n- decane from this work, T = 343 K ( ) and n-tetradecane taken from Cumicheo et al. [29], T = 344 K ( ). Symbols in (c) represent experimental IFT data of N 2 and: n-hexane taken from Garrido et al. [54], T = 333 K ( ); n-heptane taken from Nino-Amezquita et al. [51], T = 323 K ( ) and n-decane from this work, T = 343 K ( ) and T = 442 K ( ). Dashed, dotted and solid lines represent predictions from the Parachor, LGT and DGT approaches, respectively, in combination with the VT-PPR78 EoS. 34

As an example, the density profiles and adsorption isotherms obtained with the DGT are plotted in Figure 7 and Figure 8 for a temperature of 344 K and systems with n-decane. The profiles showed an enrichment of the interface in gas molecules, as reflected by the observation of a peak in the density profile of methane, carbon dioxide and nitrogen. This is in agreement with other DGT studies [29,34,51,52,72,79,98,101 104] and Molecular Dynamic simulation and Monte Carlo approaches [54,104,105]. Additionally, this relative enrichment seems to decrease with increasing pressure as the relative height of the peaks of all three light components decreases and moves towards the liquid surface and the interface thickness is enlarged. These results are supported by the adsorption isotherms plotted in Figure 8 and calculated with Eq. 16 and using experimental IFT data from the literature [33] and measured in this work. Accordingly, the adsorption of gas molecules at the interface increases with pressure and reaches a maximum for all three gases, suggesting a saturation limit for the relative adsorption of gas molecules, as also concluded by others [29,34,85,104]. Furthermore, the adsorption rate (dγ 12 /dp) is more pronounced for CO 2 when compared to that of CH 4 and N 2 in n-decane. This behaviour can help explaining the high pressure dependence of the IFT of hydrocarbon systems containing CO 2, in which the large adsorption of CO 2 leads to a higher reduction of the IFT. These results also demonstrate the inadequacy of the approximation of the mixture influence parameter (Eq. 23) and linear density assumption under the LGT framework and the poor IFT predictions obtained with this model for systems containing carbon dioxide. In summary, the good agreement between both theoretical and experimental approaches for the calculation of Γ 12 at low and moderate pressures validates to some extent the density profiles calculated with the DGT approach. The results also endorse the capability of the model to provide a perspective of the distribution of species across the interfacial region and its impact on the IFT. However, for pressures near the critical point, the theoretical approach yields values significantly different from those obtained with Eq. 19 as depicted in Figure 8. Deviations between the two approaches are related in part to large uncertainties in the calculation of the derivative ( IFT/ P) T in Eq. 19 and in limitations of both DGT and cubic EoS near the critical point. 35

ρ i /(kmol.m -3 ) 12 (a) 10 18 (b) (c) 9 16 10 8 14 8 7 12 6 10 6 5 8 4 4 6 3 2 2 4 1 2 0 0 0-3 -2-1 0 1 2 3-3 -2-1 0 1 2 3-3 -2-1 0 1 2 3 z / nm z / nm z / nm ρ i /(kmol.m -3 ) ρ i /(kmol.m -3 ) Figure 7. Vapour and liquid density profiles as function of the distance from the interface at 343 K for the systems containing n-decane and methane (a), carbon dioxide (b) or nitrogen (c) computed at different pressures with the DGT in combination with the VT-PPR78 EoS. Solid lines represent the density profile for n-decane and the dashed lines represent the density profiles of gases accordingly. Pressures in (a) equal to 1.87 MPa (Black), 10.69 MPa (Green) and 24.43 MPa (Red). Pressures in (b) equal to 0.99 MPa (Black), 5.69 MPa (Green) and 10.43 MPa (Red). Pressures in (c) equal to 14.50 MPa (Black), 28.29 MPa (Green) and 64.47 MPa (Red). 10 6 Γ 12 /(mol.m -2 ) 10 9 8 7 6 5 4 3 2 1 0 10 8 6 4 2 0 0 5 10 15 0 20 40 60 P / MPa Figure 8. Gas adsorption (Γ 12 )-pressure diagram for the systems containing methane (Green), carbon dioxide (Black) or nitrogen (Red) with n-decane at 343 K. Symbols represent the gas adsorption calculated with Eq. 19 and experimental IFT data measured in this work (Full Symbols) and from the literature [33] (Empty Symbols). Lines represent the gas adsorption predicted with Eq. 16 and density profiles obtained with the DGT approach. 36

Table 5. Summary of the calculated %AAD between measured and predicted IFT of several (gas + n-alkanes) binary systems using the VT-PPR78 EoS in combination with Parachor, LGT and DGT models. %AAD a) %AAD a) System Data Source T /K NP IFT > 1.5 mn.m -1 NP IFT 1.5 mn.m -1 Parachor LGT DGT Parachor LGT DGT Methane + n-propane Weinaug and Katz [21] 258-338 26 26.0 10.4 4.8 15 29.7 13.4 11.3 n-butane Pennington and Hough [48] 311-344 4 25.2 19.3 4.0 11 29.5 19.4 4.3 n-hexane Nino et al. [52] 300-350 9 5.4 6.4 8.2 n-heptane Jaeger et al. [50,53] 298-323 18 6.1 6.8 10.1 n-decane This work 313-442 31 10.0 7.4 14.4 8 30.5 14.9 59.8 Overall 88 14.5 10.1 8.3 34 29.9 15.9 25.1 Carbon dioxide + n-butane Hsu et al. [49] 319-377 14 18.2 24.6 10.2 24 74.1 66.4 40.9 n-heptane Jaeger et al. [50] 323-353 10 7.9 18.0 5.7 4 40.8 54.7 11.5 n-decane This work 313-442 23 10.0 23.7 3.7 4 10.5 19.6 30.9 n-dodecane Cumicheo et al. [29] 344 12 7.8 14.6 5.9 n-tridecane Cumicheo et al. [29] 344 13 3.6 20.3 3.7 1 10.0 4.1 67.7 n-tetradecane Cumicheo et al. [29] 344 14 6.9 20.5 3.9 1 4.8 1.0 63.1 Overall 86 9.1 20.3 5.5 34 28.0 29.2 42.8 Nitrogen + n-pentane Jianhua et al. [38] 313 6 16.6 6.3 5.0 n-hexane Garrido et al. [54] 303-333 28 7.2 5.3 2.9 n-heptane Nino et al. [51] 295-373 11 9.8 8.9 7.1 n-octane Jianhua et al. [38] 313 8 14.0 5.2 1.1 n-decane This work 313-442 44 21.9 12.7 8.5 2 23.8 22.4 54.9 a) NP Exp Exp % 1/ ( Model AAD = NP IFT IFT ) IFT 100 i i i i Overall 97 13.9 7.7 4.9 2 23.8 22.4 54.9 37

5. Conclusion Multiphase flow frequently occurs in many stages of chemical and petroleum engineering processes. The IFT between phases plays a decisive role in the flow characteristics and behaviour not only at the reservoir level, but also throughout the well and pipeline, and production facilities. In this work, the IFT between oil and gas were measured for representative binary systems containing n-decane and common gases such as carbon dioxide, methane and nitrogen for temperatures in the range (313-442) K and pressures up to 69 MPa. Saturated phase density data were also measured and correlated with a model based on the Peng-Robinson 1978 EoS. The adjustment of binary parameters and volume corrections allowed a reproduction of the data with relative low deviations to the density of both liquid and vapour phases. The capillary constant a 2 measured from the differential capillary rise method was combined with modelled density to estimate the IFT of the studied systems and the results were found to be in good agreement with selected literature data, validating both the methodology and the experimental procedure followed. Moreover, both density and IFT measurements complement the experimental gap found in the literature and extend these measurements to reservoir conditions. The coupling of the VT-PPR78 EoS with the Parachor, the LGT and the DGT approaches was applied to evaluate their predictive capabilities by comparison with density and IFT data measured in this work and from the literature. The IFT and phase behaviour properties of a total of 16 (gas + n-alkane) binary systems were modelled with no adjustable parameters to mixtures (i.e. in a fully predictive manner). The VT-PPR78 EoS allowed a good prediction of the effect of pressure and temperature on both VLE and volumetric properties of the studied systems. The best IFT predictions were obtained with the DGT model with an overall %AAD between (4.9 and 8.3) % to data far from the critical point. The poorest predictions were obtained with all models for IFT values lower 1.5 mn.m -1, where experimental uncertainties are considerable and classical cubic EoSs fail to describe the critical region. The DGT, through the calculation of the distribution of species in the interfacial region, showed a local enrichment and relative adsorption of gas molecules in the interfacial region, which increased with pressure and reached a saturation limit. This 38

behaviour was more pronounced for CO 2 which seems to explain the high pressure dependence of the IFT values of (CO 2 + n-alkane) systems when compared to that of (N 2 + n- alkane) and (CH 4 + n-alkane). These results also explain the large negative deviations obtained with the LGT for systems with CO 2, in which the assumption of a linear density distribution of species in the interfacial region proved inadequate. Nonetheless, the LGT stands as a good alternative of the Parachor method for n-alkane systems with methane and nitrogen. Altogether, the good agreement found between the relative Gibbs adsorption isotherms calculated using experimental IFT data and density profiles calculated with the DGT endorses the adequacy of the model for describing the complex microstructures that arise from fluid interfaces and their impact on the IFT values. 39

6. Acknowledgements This research work is part of an ongoing Joint Industrial Project (JIP) conducted jointly at the Institute of Petroleum Engineering, Heriot-Watt University and the CTP laboratory of MINES ParisTech. The JIPs is supported by Chevron, GALP Energia, Linde AG Engineering Division, OMV, Petroleum Expert, Statoil, TOTAL and National Grid Carbon Ltd, which is gratefully acknowledged. The participation of National Grid Carbon in the JIP was funded by the European Commission s European Energy Programme for Recovery. The authors would also like to thank the members of the steering committee for their fruitful comments and discussions. Luís M. C. Pereira acknowledges the financial support from Galp Energia through his PhD grant. 40

7. References [1] D.W. Green, G.P. Willhite, Enhanced Oil Recovery, SPE Textbook Series, 1998. [2] K. Holmberg, D.O. Shah, M.J. Schwuger, eds., Handbook of Applied Surface and Colloid Chemistry, John Wiley & Sons, Ltd, 2002. [3] A.Y. Dandekar, Petroleum Reservoir Rock and Fluid Properties, Taylor & Francis, 2006. [4] J. Straub, Exp. Therm. Fluid Sci. 9 (1994) 253 273. [5] J.W. Rose, Chem. Eng. Res. Des. 82 (2004) 419 429. [6] W.R. Foster, J. Pet. Technol. 25 (1973) 205 210. [7] E.J. Lefebvre du Prey, Soc. Pet. Eng. J. 13 (1973) 39 47. [8] R. Ehrlich, H.H. Hasiba, P. Raimondi, J. Pet. Technol. 26 (1974) 1335 1343. [9] J.C. Melrose, J. Can. Pet. Technol. 13 (1974). [10] A. Abrams, Soc. Pet. Eng. J. 15 (1975) 437 447. [11] S.G. Oh, J.C. Slattery, Soc. Pet. Eng. J. 19 (1979) 83 96. [12] S. Zhang, G.-C. Jiang, L. Wang, H.-T. Guo, X. Tang, D.-G. Bai, J. Dispers. Sci. Technol. 35 (2014) 403 410. [13] O.R. Wagner, R.O. Leach, Soc. Pet. Eng. J. 6 (1966). [14] C. Bardon, D.G. Longeron, Soc. Pet. Eng. J. (1980) 391 401. [15] M.S. Haniff, J.K. Ali, Relative Permeability and Low Tension Fluid Flow in Gas Condensate Systems, in: Proc. Eur. Pet. Conf., Society of Petroleum Engineers, 1990: pp. 1 8. [16] J.K. Ali, S. Butler, L. Allen, P. Wardle, The Influence of Interfacial Tension on Liquid Mobility in Gas Condensate Systems, in: Offshore Eur., Society of Petroleum Engineers, 1993. [17] A. Prosperetti, G. Tryggvason, eds., Computational Methods for Multiphase Flow, Cambridge University Press, Cambridge, 2007. [18] L.M.C. Pereira, A. Chapoy, R. Burgass, M.B. Oliveira, J.A.P. Coutinho, B. Tohidi, J. Chem. Thermodyn. (2015). http://dx.doi.org/10.1016/j.jct.2015.05.005 41

[19] P. Cheng, D. Li, L. Boruvka, Y. Rotenberg, A.W. Neumann, Colloids and Surfaces. 43 (1990) 151 167. [20] B. Song, J. Springer, J. Colloid Interface Sci. 184 (1996) 64 76. [21] C.F. Weinaug, D.L. Katz, Ind. Eng. Chem. 35 (1943) 239 246. [22] C.A. Swartz, Physics. 1 (1931) 245. [23] V.G. Baidakov, A.M. Kaverin, M.N. Khotienkova, Fluid Phase Equilib. 356 (2013) 90 95. [24] G. Wiegand, E.U. Franck, Berichte Der Bunsengesellschaft Für Phys. Chemie. 98 (1994) 809 817. [25] B.-Y. Cai, J.-T. Yang, T.-M. Guo, J. Chem. Eng. Data. 41 (1996) 493 496. [26] Q.-Y. Ren, G.-J. Chen, W. Yan, T.-M. Guo, J. Chem. Eng. Data. 45 (2000) 610 612. [27] G.Y. Zhao, W. Yan, G.J. Chen, X.Q. Guo, Shiyou Daxue Xuebao/Journal Univ. Pet. China. 26 (2002) 75 78+82. [28] P. Chiquet, J.-L. Daridon, D. Broseta, S. Thibeau, Energy Convers. Manag. 48 (2007) 736 744. [29] C. Cumicheo, M. Cartes, H. Segura, E.A. Müller, A. Mejía, Fluid Phase Equilib. 380 (2014) 82 92. [30] H. Nourozieh, M. Kariznovi, J. Abedi, J. Chem. Thermodyn. 58 (2013) 377 384. [31] H.H. Reamer, B.H. Sage, J. Chem. Eng. Data. 8 (1963) 508 513. [32] N. Nagarajan, R.L. Robinson, J. Chem. Eng. Data. 31 (1986) 168 171. [33] R. Shaver, R.. Robinson, K. Gasem, Fluid Phase Equilib. 179 (2001) 43 66. [34] A. Mejía, M. Cartes, H. Segura, E.A. Müller, J. Chem. Eng. Data. 59 (2014) 2928 2941. [35] B.H. Sage, H.M. Lavender, W.N. Lacey, Ind. Eng. Chem. 32 (1940) 743 747. [36] H.H. Reamer, R.H. Olds, B.H. Sage, W.N. Lacey, Ind. Eng. Chem. 34 (1942) 1526 1531. [37] R. Amin, T.N. Smith, Fluid Phase Equilib. 142 (1998) 231 241. [38] T. Jianhua, J. Satherley, D.J. Schiffrin, Chin.J.Chem.Eng. 1 (1993) 223 231. 42

[39] A. Georgiadis, F. Llovell, A. Bismarck, F.J. Blas, A. Galindo, G.C. Maitland, J.P.M. Trusler, G. Jackson, J. Supercrit. Fluids. 55 (2010) 743 754. [40] G. Stegemeier, B. Pennington, Old SPE J. 2 (1962) 257 260. [41] D.B. Robinson, D.-Y. Peng, The characterization of the heptanes and heavier fractions for the GPA Peng-Robinson programs, Gas Processors Association, Research Report RR-28, 1978. [42] D. Macleod, Trans. Faraday Soc. 2 (1923) 2 5. [43] S. Sugden, J. Chem. Soc. Trans. (1924) 32 41. [44] J.W. Cahn, J.E. Hilliard, J. Chem. Phys. 28 (1958) 258. [45] J.S. Rowlinson, J. Stat. Phys. 20 (1979) 197 200. [46] Y.-X. Zuo, E.H. Stenby, J. Chem. Eng. Japan. 29 (1996) 159 165. [47] Y.-X. Zuo, E.H. Stenby, J. Colloid Interface Sci. 182 (1996) 126 132. [48] B.F. Pennington, E.W. Hough, Prod. Mon. July (1965) 4. [49] J.J.C. Hsu, N. Nagarajan, R.L. Robinson, J. Chem. Eng. Data. 30 (1985) 485 491. [50] P.T. Jaeger, M.B. Alotaibi, H.A. Nasr-El-Din, J. Chem. Eng. Data. 55 (2010) 5246 5251. [51] O.G. Niño Amézquita, S. Enders, P.T. Jaeger, R. Eggers, J. Supercrit. Fluids. 55 (2010) 724 734. [52] O.G. Niño-Amezquita, S. Enders, P.T. Jaeger, R. Eggers, Ind. Eng. Chem. Res. 49 (2010) 592 601. [53] P.T. Jaeger, R. Eggers, J. Supercrit. Fluids. 66 (2012) 80 85. [54] J.M. Garrido, L. Cifuentes, M. Cartes, H. Segura, A. Mejía, J. Supercrit. Fluids. (2014). [55] J.-N. Jaubert, F. Mutelet, Fluid Phase Equilib. 224 (2004) 285 304. [56] J.-N. Jaubert, S. Vitu, F. Mutelet, J.-P. Corriou, Fluid Phase Equilib. 237 (2005) 193 211. [57] F. Mutelet, S. Vitu, R. Privat, J.-N. Jaubert, Fluid Phase Equilib. 238 (2005) 157 168. [58] J.-N. Jaubert, R. Privat, F. Mutelet, AIChE J. 56 (2010) 3225 3235. 43

[59] J.-W. Qian, J.-N. Jaubert, R. Privat, Fluid Phase Equilib. 354 (2013) 212 235. [60] J.J. Martin, Ind. Eng. Chem. Fundam. 18 (1979) 81 97. [61] A. Peneloux, E. Rauzy, R. Freze, Fluid Phase Equilib. 8 (1982) 7 23. [62] C. Miqueu, B. Mendiboure, A. Graciaa, J. Lachaise, Fluid Phase Equilib. 207 (2003) 225 246. [63] J. Lane, J. Colloid Interface Sci. 42 (1973) 145 149. [64] E.W. Lemmon, M.O. McLinden, M.L. Huber, NIST Reference Fluid Thermodynamic and Transport Properties REFPROP, Version 8.0, NIST, Gaithersburg, MD, 2007. [65] J. Ali, Fluid Phase Equilib. 95 (1994) 383 398. [66] J.R. Fanchi, SPE Reserv. Eng. (1990) 433 436. [67] K. Kashefi, L.M.C. Pereira, A. Chapoy, R. Burgass, B. Tohidi, Fluid Phase Equilib. 409 (2016) 301 311. [68] Z. J.D. Van der Waals, Phys Chem. 13. (1894) 657 672. [69] A.J.M. Yang, P.D. Fleming III, J.H. Gibbs, J. Chem. Phys. 64 (1976) 3732. [70] V. Bongiorno, L.E. Scriven, H.T. Davis, J. Colloid Interface Sci. 57 (1976) 462 475. [71] B. Breure, C.J. Peters, Fluid Phase Equilib. 334 (2012) 189 196. [72] H. Lin, Y.-Y. Duan, Q. Min, Fluid Phase Equilib. 254 (2007) 75 90. [73] M.B. Oliveira, I.M. Marrucho, J.A.P. Coutinho, A.J. Queimada, Fluid Phase Equilib. 267 (2008) 83 91. [74] M.B. Oliveira, J.A.P. Coutinho, A.J. Queimada, Fluid Phase Equilib. 303 (2011) 56 61. [75] B.S. Carey, L.E. Scriven, H.T. Davis, J. Chem. Phys. 69 (1978) 5040. [76] M. Sahimi, H.T. Davis, L.E. Scriven, Soc. Pet. Eng. J. 25 (1985) 235 254. [77] H.T. Davis, Statistical Mechanics of Phases, Interfaces and Thin Films, Wiley, 1995. [78] P.M.W. Cornelisse, The Square Gradient Theory Applied: Simultaneous Modelling of Interfacial Tension and Phase Behaviour (PhD Thesis), Delft University of Technology, 1997. 44

[79] C. Miqueu, B. Mendiboure, C. Graciaa, J. Lachaise, Fluid Phase Equilib. 218 (2004) 189 203. [80] C. Miqueu, B. Mendiboure, A. Graciaa, J. Lachaise, Ind. Eng. Chem. Res. 44 (2005) 3321 3329. [81] C. Miqueu, B. Mendiboure, A. Graciaa, J. Lachaise, Fuel. 87 (2008) 612 621. [82] T. Wadewitz, J. Winkelmann, Berichte Der Bunsengesellschaft Für Phys. Chemie. 100 (1996) 1825 1832. [83] M.M. Telo da Gama, R. Evans, Mol. Phys. 41 (1980) 1091 1112. [84] M.M. Telo da Gama, R. Evans, Mol. Phys. 48 (1983) 229 250. [85] C. Miqueu, Modélisation à température et pression élevées de la tension superficielle de composants des fluides pétroliers et de leurs mélanges synthétiques ou réels (PhD Thesis), Université de Pau et des Pays de l Adour, 2001. [86] Y.-X. Zuo, E.H. Stenby, SPE J. 3 (1998) 134 145. [87] S. Khosharay, M.S. Mazraeno, F. Varaminian, Int. J. Refrig. 36 (2013) 2223 2232. [88] Y.-X. Zuo, E.H. Stenby, Fluid Phase Equilib. 132 (1997) 139 158. [89] K.A.G. Schmidt, G.K. Folas, B. Kvamme, Fluid Phase Equilib. 261 (2007) 230 237. [90] W. Yan, G.-Y. Zhao, G.-J. Chen, T.-M. Guo, J. Chem. Eng. Data. 46 (2001) 1544 1548. [91] S. Khosharay, F. Varaminian, Korean J. Chem. Eng. 30 (2013) 724 732. [92] Y.-X. Zuo, E.H. Stenby, Situ. 22 (1998) 157 180. [93] M.L. Michelsen, J.M. Mollerup, Thermodynamic Models: Fundamentals & Computational Aspects, second ed., Tie-Line Publications, 2007. [94] A.M. Abudour, S.A. Mohammad, R.L. Robinson, K.A.M. Gasem, Fluid Phase Equilib. 335 (2012) 74 87. [95] A.M. Abudour, S.A. Mohammad, R.L. Robinson, K.A.M. Gasem, Fluid Phase Equilib. 349 (2013) 37 55. [96] B.N. Taylor, C.E. Kuyatt, Guidelines for Evaluating and Expressing the Uncertainty of NIST Measurement Results; NIST: Gaithersburg, MD, 1994. [97] M.K. Gupta, R.L. Robinson Jr., SPE Reserv. Eng. (1987) 528 530. 45

[98] M. Sahimi, B.N. Taylor, J. Chem. Phys. 95 (1991) 6749. [99] Y.-X. Zuo, E.H. Stenby, Can. J. Chem. Eng. 75 (1997) 1130 1137. [100] A. Azarnoosh, J.J. McKetta, J. Chem. Eng. Data. 8 (1963) 494 496. [101] D. Fu, Y. Wei, Ind. Eng. Chem. Res. 47 (2008) 4490 4495. [102] P.M.W. Cornelisse, C.J. Peters, J.D. Arons, Fluid Phase Equilib. 82 (1993) 119 129. [103] P.M.W. Cornelisse, C.J. Peters, J. De Swaan Arons, Mol. Phys. 80 (1993) 941 955. [104] E.A. Müller, A. Mejía, Fluid Phase Equilib. 282 (2009) 68 81. [105] J. Neyt, A. Wender, V. Lachet, P. Malfreyt, J. Phys. Chem. C. 116 (2012) 10563 10572. 46

8. Appendix A The thermodynamic phase behaviour model used in this work to model the VLE and densities of the studied mixtures was based on the PR78 EoS [41] with the volume-translation technique. Accordingly, this EoS is expressed for a pure substance as: RT a( T ) p = v b v( v + b) + b( v b) (A.1) 2 2 0.45724R Tc a( T) = 1 + m(1 Tr ) P c 2 (A.2) 2 = 0.37464 + 1.54226 0.26992 for ω<0.49 (A.3) m ω ω 2 3 = 0.3796 + 1.485 0.1644 + 0.01667 for ω 0.49 (A.4) m ω ω ω RTc b = 0.07780 (A.5) P PR c c v = v + v (A.6) where a and b are the EoS energy and co-volume pure substance parameters, respectively, obtained from the pure critical data, T c and P c, and the acentric factor, ω. When dealing with the volume-translated version of this EoS, v PR and v c are the untranslated molar volume and the volume correction, respectively. For mixtures, the energy and co-volume parameters are calculated employing the conventional van der Waals one-fluid mixing rules, with one binary interaction parameter, k ij, for the energy parameter and the linear mixing rule for the mixture volume correction. They have the following form: a = x x (1 k ) a a i j i j ij i j (A.7) b = xibi (A.8) i v = x v c i c i i (A.9) 47

48