Anda di halaman 1dari 34

1.

Composition and physical properties of hydrocarbons

1.1.1 Introduction
Petroleum is a naturally occurring liquid, gaseous or even solid mixture, composed principally of hydrocarbons, that accumulates in subterranean reservoirs. As a generic term that keeps varying with time, petroleum is difficult to define precisely. In this text, it is used interchangeably with hydrocarbons when these compounds represent the predominant portion in the mixture. However, our discussion will not be exclusive to hydrocarbons but will also include relevant non-hydrocarbons. The constituents of a natural reservoir fluid form an almost continuous hydrocarbon spectrum from the lightest one, methane, through intermediate molecular weights and up to very large molecules. The relative proportion of these different parts can vary in a large range, which results in petroleum fluids showing very different features. For instance, the simplest reservoir fluids are natural gas, while the most complex molecular mixtures are those of black oil and bitumen. As can be seen from this example, the physical properties of a petroleum fluid are determined by its composition and, moreover, many of these vary significantly as a function of temperature and pressure. Petroleum reservoirs have temperatures that vary from ambient to more than 200C and a pressure that can be as high as 150 MPa. The reservoir conditions depend on the depth of the reservoir and on the geological processes that the reservoir has experienced since it was filled by the reservoir fluid. The temperature and pressure change dramatically, relative to the reservoir conditions, during production where the pressure is lowered to fulfil transport and storage conditions. These changes in temperature

and pressure will often lead to a change in the state of the mixture from a single phase state to a two-phase state where the two phases will each have a different composition in comparison to the original petroleum fluid. Therefore, the properties of the new phases will change with the change in composition, temperature and pressure. Even within the reservoir the physical state of the mixture (gas, liquid or solid) can change due to varying composition, temperature and pressure.

1.1.2 Composition of petroleum fluids


Two important questions regarding the composition of petroleum fluids are what is in the mixture? (qualitative chemical analysis) and how can the amount of each component be quantitatively determined? (quantitative chemical analysis). The chemical composition of petroleum is largely speculative due to the difficulty of a complete identification caused by the enormous number of components. For a hydrocarbon with a given molecular formula CxHy, the number of possible isomers increases dramatically as the molecule becomes larger. To illustrate this point, Fig. 1 shows how drastically the number of possible isomers increases in the paraffin group alone. Even though not all of them exist in real oils, the presence of other hydrocarbon groups and heterocompounds will complicate the task. Even for the paraffins in the C5-C12 range, the number of possible isomers is greater than 600 (Altgelt and Boduszynski, 1994), around 200-400 of which have been experimentally observed (though not all identified). Fortunately, in most cases, we do not need an all-out effort to identify and quantify all of the

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

31

GEOSCIENCES

1040 5.92.1039

(Speight, 2001) as shown by the following values of percentage weight:


1.06.1031

number of isomers

1030

1020

2.21.1022 6.24.1013 .109 4.11 3.66.103 75 10 20 30 40 50 60 70 80 90 100 110

Carbon: Hydrogen: Nitrogen: Oxygen: Sulphur:

83.0-87.0% 10.0-14.0% 0.1-2.0% 0.05-1.5% 0.05-6.0%

1010 3 0

carbon number
Fig. 1. Increase in the number of isomers with carbon number in paraffin.

components. Hydrocarbons up to and including C5 are easily determined on a molar basis. The remaining identified components, which normally go up to 20, or in some cases 30, carbon atoms, are not individual molecular compounds but rather fractions defined by a normal boiling point range. These fractions are defined based on the normal boiling point of the normal alkanes and are called the C6 , C7, etc. fractions (or components) according to the carbon number of the normal alkane within the fraction. Finally, the part of the oil that cannot be analysed is called the residue or plus fraction: for example, the symbol C20 indicates the residue that includes the C20 fraction and all of the material in the oil less volatile than this. The above compositional analysis is generally considered to be sufficient in terms of oil characterization for physical property studies. However, some laboratories systematically identify all components up to C12 by the use of gas chromatography combined with mass spectroscopy. Chemical composition Gases produced from a petroleum reservoir mainly contain alkanes lighter than heptane, with methane and ethane being the predominant components. However, light non-hydrocarbons including nitrogen, carbon dioxide, and hydrogen sulphide are also common, their proportions being related to the reservoir where they originate from. Petroleum liquid, or crude oil, contains larger molecules and its appearance, composition, and other properties vary a lot with different petroleum reservoirs. Nevertheless, nearly all naturally occurring petroleum liquids have fairly narrow limits of elemental composition

The narrow range of the carbon to hydrogen ratio reflects the fact that CH2 group is the primary unit in various organic molecules in crude oils. Generally speaking, components in petroleum liquids can be classified into hydrocarbons and heterocompounds. The term hydrocarbon is used for molecules made up only of carbon and hydrogen atoms. On the other hand, heterocompounds are compounds which, in addition to carbon and hydrogen, also contain one or more heteroatoms such as sulphur, nitrogen, oxygen, vanadium, nickel or iron. The principal chemical and physical characteristics of three crudes originating from three different oilfields and the hydrocarbon fractions obtained from these using the true boiling point distillation method (see below) are reported in Table 1.
Hydrocarbon components

The hydrocarbon components of petroleum fall into three groups: paraffins, naphthenes and aromatics. Olefins (also called alkenes) are so scarce in naturally occurring petroleum that they may be neglected. Also, the presence of dienes (R CH CH R') and acetylenes (RC CR') is considered to be extremely unlikely. Paraffins, also known as acyclic alkanes, are saturated hydrocarbons with straight (normal paraffin) or branched (isoparaffin) chains, but without any ring structure. Both normal and isoparaffins have the same molecular formula CnH2n+2. Naphthenes, also known as cycloalkanes or alicyclic hydrocarbons, are saturated hydrocarbons containing one or more rings, each of which may have one or more paraffinic side chains. Naphthenes are present in all fractions in which the constituent molecules contain more than five carbon atoms. Aromatics are compounds containing at least one benzene ring. Many of the aromatic hydrocarbons in petroleum consist of aromatic and naphthenic rings and bear normal and/or branched alkane side chains. The proportions of the above three groups vary with the type of crude, but within any crude oil, the proportion of paraffinic hydrocarbons usually decreases with increasing molecular weight or boiling

32

ENCYCLOPAEDIA OF HYDROCARBONS

Table 1. Chemical and physical characteristics of three crudes originating in different fields

Crude A Yields and characteristics of products


Gas Naphthas Kerosines Gasoils V Dist. .
370-530 370+ 410+ 530+ 43.84 38.77 1.0883 C1 C2 4.31 5.25 15.80 25.69 32.12 6.22 8.66 8.99 4.85 50.73 3.35 1.8700 36 48.76 0.0500 0.3850 0.4506 1.7650 52 93 108 168 56.90 9.36 C3 i-C4 n-C4 i-C5 0.77 0.79 0.78 1.50 1.62 82.2 70.7 66.1 22.9 11.0 23.8 25.1 44.9 24 23 53 6 9 46.7 0.00003 0.00003 0.00003 0.00003 399 521 0.14 1.9 16.1 161.2 3.0 19.73 1.07 13.26 126.0 175.0 3.3 21.42 14.78 172.9 240.2 2.9 29.40 19.76 21 18 48.5 0.3854 0.4796 53 68 >100 56 21.1 8.2 37.5 13.8 4.0 21.7 n-C5 80 100 120 140 160 180 210 230 250 270 290 320 350 370 390 410 530 11.89 11.40 11.30 550 60.17 54.93 1.0541 65.31 60.28 1.0426 21.47 21.52 0.9603 80-180 160-230 7.18 8.54 0.8090 0.8149 0.8730 0.9248 7.20 26.71 5.35 6.10 24.23 5.14 3.12 3.98 0.7531 180-230 230-370 370-416

Crude Oil
Residues
C1-C4 80-160 2.04 2.65 0.7408 0.29 0.97 1.37 0.6800 0.50 0.5491 C5-80

TBP range

TBP Distillation A
Cut point %Wt Cum

TBP yield

% wt

TBP yield

% vol

Density at 15C

kg/l

0.9624

API Gravity at 60F

15.5

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

Viscosity at 20C

cSt

1585.94

Viscosity at 50C

VBN

0.08 0.16 0.28 0.40 0.55 1.25 1.45 1.81 2.45 3.29 4.37 7.91 10.47 13.33 16.82 20.67 26.75 31.34 34.69 37.39 39.83 56.16 61.89

Sulphur

% wt

6.70

Mercaptan sulphur

ppm

310

Hydrogen sulphide

% wt

0.001

Acidity

mgKOH/g

2.13

Paraffins

% vol

Naphthenes

% vol

Aromatics

% vol

N+2A index

Smoke Pt.

mm

Freezing Pt.

Cloud Pt.

Pour Pt.

15

Cetane index

Total nitrogen

% wt

Basic nitrogen

ppm

Nickel

ppm

75.8

Vanadium

ppm

105.3

P. Value

Asphaltenes

% wt

12.89

R.C.C.

% wt

8.66

Penetration at 25C

dmm

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

UOP K factor

12.05

33

34
Yields and characteristics of products
Gas Naphthas Kerosines Gasoils V. Dist.
370-530 370+ 410+ 530+ 21.06 17.42 1.0460 36.56 31.44 1.0061 42.61 37.15 0.9924 21.55 19.73 0.9452 80-180 160-230 13.40 14.70 0.7883 0.7945 0.8630 0.9170 10.23 24.78 5.71 9.39 24.72 5.05 17.79 20.66 0.7449 180-230 230-370 370-416

GEOSCIENCES

Table 1. Chemical and physical characteristics of three crudes originating in different fields

Crude B
Residues

Crude Oil
C1-C4 80-160 13.78 16.19 0.7368 1.11 4.39 5.74 0.6606 1.72 0.5549 C5-80

TBP range

TBP yield

% wt

TBP Distillation B
Cut point
C1

TBP yield

% vol

Density at 15C

kg/l

0.8652

%Wt Cum
0.00 C2 C3 0.02 0.30

API gravity at 60F

32.0

Viscosity at 20C 2.81 4.23 15.25 25.08 30.62 3.72 5.06 5.33 3.46 40.80 2.25 305 0.2250 306 38.57 0.0350 0.0380 0.0787 0.1850 241 224 264 186 46.70 6.44

cSt

8.77

Viscosity at 50C

VBN

Sulphur

% wt

2.80

Mercaptan sulphur

ppm

i-C4 n-C4

0.46 1.10 i-C5 n-C5 80 1.69 2.75 5.49 100 120 8.52 11.74 140 160 15.42 19.27

Hydrogen sulphide 0.11 0.14 0.07 0.08 0.06 90.4 74.2 70.5 15.7 13.8 21.0 21.9 43.3 26 25 48 4 9 49.7 0.00003 0.00003 0.00003 0.00003 190 302 0.14 1.9 11.2 6.9 4.8 1.32 0.44 10.31 25 24 52.2 30 54 13.4 12.4 38.2 7.4 2.3 11.9

% wt

0.001

Acidity

mgKOH/g

0.09

Paraffins

% vol

Naphthenes

% vol

Aromatics

% vol

N+2A index

Smoke Pt.

mm

Freezing Pt.

Cloud Pt.

180 33 67 210 230 0.2446 0.3329 250 270 13.0 8.0 5 1.54 12.28 2.67 20.87 22.6 13.9 290 320 350 370 390 410 12.02 11.73 11.89 530 550

23.28 29.06 32.67 36.39 40.26 43.82 48.13 54.21 57.39 60.45 63.44 78.94 81.90

Pour Pt.

15

Cetane index

Total nitrogen

% wt

Basic nitrogen

ppm

Nickel

ppm

4.8

Vanadium

ppm

2.9

P. Value

Asphaltenes

% wt

0.56

R.C.C.

% wt

4.39

Penetration at 25C

dmm

ENCYCLOPAEDIA OF HYDROCARBONS

UOP K factor

12.38

Table 1. Chemical and physical characteristics of three crudes originating in different fields

Crude C Yields and characteristics of products


Gas Naphthas Kerosines Gasoils V. Dist.
370-530 19.91 18.09 0.8926 0.9181 0.9320 0.9780 25.81 19.65 7.72 29.22 22.58 9.31 370+ 410+ 530+ 160-230 180-230 230-370 370-416 11.57 29.56 6.64 6.16 0.8741 28.66 0.8364 11.89 0.7894 16.25 16.84 0.7825

Crude Oil
Residues
C1-C4 70-160 17.57 19.49 0.7311 2.60 4.80 5.95 0.6544 3.86 0.5447 C5-70

TBP range

TBP yield

% wt

TBP Distillation C
Cut point
C1

TBP yield

% vol

Density at 15C

kg/l

0.8110

%Wt Cum

API gravity at 60F

42.9 C2 C3 0.98 i-C4 n-C4 0.04 0.04 0.03 93.7 74.3 21.0 4.7 9.1 30.4 32 30 45 2 5 61.3 0.0001 0.0001 0.0053 26 72.3 0.0274 104 0.0303 162 1.5 12.0 1.1 2.50 0.05 50.01 6.48 2.0 15.0 1.1 4.8 35.0 1.2 7.90 15.60 0.1300 0.1600 0.3200 30 33 39 51 8.5 4.1 2.2 8.5 i-C5 n-C5 70 100 120 140 160 180 210 230 250 270 290 320 350 370 390 410 12.00 11.90 11.90 530 530 0.05 0.90 1.67 2.60 3.66 5.03 7.40 11.87 15.74 20.24 24.97 29.65 36.99 13.33 41.22 45.73 50.56 54.80 61.69 67.05 73.99 77.42 90.69

Viscosity at 20C 4.99 15.56 24.28 28.43 32.96 0.62 0.68 0.46 0.42 35.52 0.19 0.0009 0.0006 0.0210 3 10 0.0156 42.67

cSt

3.56

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

Viscosity at 50C

VBN

Sulphur

% wt

0.24

Mercaptan sulphur

ppm

Hydrogen sulphide

% wt

Acidity

mgKOH/g

0.05

Paraffins

% vol

Naphthenes

% vol

Aromatics

% vol

N+2A index

Smoke Pt.

mm

Freezing Pt.

Cloud Pt.

Pour Pt.

36

Cetane index

Total nitrogen

% wt

Basic nitrogen

ppm

Nickel

ppm

0.5

Vanadium

ppm

3.4

P. Value

Asphaltenes

% wt

0.73

R.C.C.

% wt

1.46

Penetration at 25C

dmm

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

UOP K factor

12.15

35

GEOSCIENCES

paraffins

gases and oils constitute the main aspects in obtaining the representative overall composition of a reservoir fluid. For a discussion of the sampling and recombination procedures, see chapters 3.3 and 4.2.
naphtheno-aromatics

weight (%)

naphthenes

Analysis of hydrocarbon gases and oils

aromatics

heteroatomic compounds boiling point

Fig. 2. Distribution of various compound types throughout petroleum (Speight, 2001).

point (Fig. 2). As the boiling point of the petroleum fraction increases, not only will the number of constituents increase but so also will the molecular complexity of the constituents.
Non-hydrocarbon components (heterocompounds)

The non-hydrocarbon components of petroleum consist of: sulphur compounds commonly including thiols, sulphides, cyclic sulphides, disulphides, benzothiophene, dibenzothiophene and naphthobenzothiophene; oxygen compounds including alcohols, ethers, carboxylic acids, esters, ketones and furans; and nitrogen compounds including pyrrole, indole, carbazole, benzo(a)carbazole, pyridine, quinoline, indoline, and benzo(f)quinoline. Among the non-hydrocarbon species present in petroleum certain metals can also be found. These heterocompounds, as well as metals, are generally found only in the non-volatile portion of crude oil (Speight, 2001). The concentration of these heterocompounds is usually quite small, although it tends to increase with increasing boiling point (see again Fig. 2). Furthermore, their presence mainly influences the processibility of the crude oil and the quality of the petroleum products rather than the physical properties of petroleum. Compositional analysis Reservoir fluid samples at elevated pressure cannot be introduced into a Gas Chromatography instrument for direct analysis unless a special high pressure injection technique is employed. Therefore, the high pressure samples are generally separated into atmospheric gas and oil which are analysed separately. Sampling, recombination and analysis of hydrocarbon

The composition of hydrocarbon gases can be determined by Gas Chromatography (GC) and that of oils by True Boiling Point (TBP) distillation or simulated distillation. Furthermore, Mass Spectroscopy (MS) can be used together with GC for the most detailed analysis. Gas Chromatography. GC is a chromatographic method for separating volatile components using a gas as a mobile phase. A small amount of sample is introduced into the gas chromatograph through a heated injection port, where the liquid (if any) is vaporized. This sample is then transported by a carrier gas, e.g. helium, into a column coated or packed with a stationary phase. Components in the sample are retained by the stationary phase, then released and displaced forward by the upstream carrier gas, and, finally, elute in the reverse order of their affinity to the stationary phase, that is, with respect to the force with which they are retained. Gas chromatography can have up to one million theoretical equilibrium stages and thus has a high separation ability. The eluted components can be identified by their retention times, which are calibrated in advance, and quantified by different detectors. The two most commonly used detectors are the Flame Ionization Detector (FID) and the Thermal Conductivity Detector (TCD), where the former is usually employed for hydrocarbons and the latter for non-hydrocarbons. True Boiling Point (TBP) distillation. True Boiling Point distillation is traditionally used in analysing oil by fractionating it into relatively narrow fractions or cuts. These TBP fractions can then be treated as components with specific boiling points, molecular weights and critical properties, that is, in the same way as a pure component. Distillation separates molecules by their difference in volatility (vapour pressure, boiling point). However, it cannot be taken for granted that distillation is based on molecular weight difference. Although the volatility of hydrocarbons decreases with molecular weight within each homologous series, the difference between boiling points in the different homologous series is quite substantial. An increase in aromaticity and polarity decreases the volatility and increases the boiling point. TBP distillation consists of distillations both at atmospheric pressure and at reduced pressure. Atmospheric distillation is confined to a maximum

36

ENCICLOPEDIA DEGLI IDROCARBURI

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

reboiler temperature of 345C to avoid decomposition. Subsequently, vacuum distillation at pressures ranging from 5 to 200 mmHg (~0.7-27 kPa) is usually used to separate higher fractions. The boiling points determined at reduced pressure are converted to socalled Atmospheric Equivalent Temperatures (AET), which are hypothetical boiling points at atmospheric pressure if decomposition could be avoided. The standard American Society for Testing Materials (ASTM) D-2892 describes the TBP distillation procedure in detail. In practice, different variations can be implemented. In general, TBP distillation is a standard method of oil analysis, which fractionates oil truly according to their boiling points. Moreover, physical samples of each fraction can be obtained and further determination of their physical properties, e.g. molecular weight and specific gravity, is possible. However, the drawback of TBP distillation is that it requires large amounts of sample (1-10 litres) and a long analysis time (100 hours). Pedersen et al. (1989b) suggested using a mini-distillation apparatus which only requires 100 ml sample. The normal boiling points of paraffin plus 0.5C, as suggested by Katz and Firoozabadi (1978), are generally accepted as boundaries between different fractions, and the resulting narrow fractions are often called Single Carbon Number (SCN) fractions, to emphasize that they contain compounds whose boiling points fall within a narrow range. Each of these SCN fractions is named after the number of carbon atoms of the n-alkane in the fraction, e.g. C7, C8, etc. Simulated distillation. Simulated distillation is essentially a low resolution gas chromatography method. The method is based on the observation that hydrocarbons are sequentially eluted from a non-polar column almost in the order of their boiling points. After calibration with normal alkanes, the retention times for an oil sample can be readily converted to boiling points. In the case where complete elution of heavy ends is impossible, an internal standard must be used to determine the amount of eluted fractions. Simulated distillation is preferred to the TBP method since it is quick and requires only a small amount of sample. On the other hand, no physical fractions can be obtained with this method and therefore their molecular weights and specific gravities are unavailable. ASTM D-2887 is a simulated distillation standard up to approximately 540C (1,000F) atmospheric equivalent boiling point. However, recent efforts have focused on extending the range up to 800C (1,470F). Gas Chromatography coupled to Mass Spectrometry (GC-MS) and others. In the GC-MS method, a gas chromatography instrument and a mass

spectrometer are employed in series. GC serves as a separation procedure of different compounds, whereas MS provides the molecular weight, the chemical formula and the amount of each compound. The GC-MS analysis can be applied to gas or oil samples. Using this method, each individual substance in the mixture can, in principle, be detected quantitatively if the amount is above the detection limit. However, with increasing boiling point, the number of possible molecular structures increases dramatically, while the concentration of the compounds generally decreases, and thus the MS analysis becomes increasingly difficult. Therefore, a significant fraction of the oil will be unrecognized by the GC-MS and its amount will be determined by a mass balance in which all of the detected compounds are subtracted from the total oil. GC-MS can determine the distribution of paraffins, naphthenes and aromatics (indicated by the acronym: PNA) in a SCN fraction. Information about the PNA distribution can be used to improve the C7+ characterization. However, a complete PNA distribution of the whole oil is difficult to obtain and even if it is available, many uncertainties still exist in evaluating the properties of heavy components. In this sense, PNA analysis has limited usefulness in practical characterization. A more detailed discussion of different analysis methods, especially for heavy petroleum fractions can be found in Altgelt and Boduszynski (1994).

1.1.3 Physical properties of hydrocarbons


In quantitative simulations of the production and processing of petroleum fluids, two approaches to the description of petroleum fluids are usually employed: the black oil approach, where the fluid is described by only two components, oil and gas, and the compositional approach, where the fluid is described by a number of components. The black oil approach is directly based on Pressure-VolumeTemperature (PVT) experiments (see Chapter 4.2), however, its modelling of phase equilibria and physical properties is approximate. On the other hand, the compositional approach employs Equations Of State (EOS) and provides a more accurate fluid estimation. EOS modelling requires a detailed knowledge of the molar composition of the mixture and the critical properties and the acentric factor of all the components. The next section discusses how to obtain the required information either experimentally or through a characterization procedure.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

37

GEOSCIENCES

A major step in modelling the physical properties of reservoir fluids is to calculate volumetric properties and associated phase equilibria (PVT modelling) as well as other important physical properties such as viscosity and interfacial tension. A discussion of these issues is presented in the following. Physical properties of narrow distillation cuts and their characterization The compositional analysis cannot provide the required information for all of the real components in a petroleum mixture. The compositional information obtained varies with boiling point range. Hydrocarbon components up to C6 and non-hydrocarbon gases (N2, CO2 and H2S) can be quantified discretely and the properties of these well-defined components are readily found in handbooks. The low boiling range in the C7+ fractions (e.g. up to C30) can be fractionated in terms of narrow TBP fractions and their molecular weights and specific gravities can be experimentally determined. For the heavy end of the C7+ fractions, TBP residue, no further analysis of the molar distribution is made and only the molecular weight and specific gravity of the whole residue are experimentally determined. It is also very common that the only available information about the C7+ fraction is its molar composition, specific gravity and molecular weight. To create a set of components, which can be directly used by the EOS method, from the limited compositional information, proper characterization of the C7+ fraction is needed. This characterization roughly includes two steps: determination of the molar composition of all components (including the C7+ fractions) and estimation of the critical properties and the acentric factor (see below) of the C7+ fractions. In practice, the number of components in a characterized oil is too high for simulation purposes and lumping to a lower number of pseudocomponents is generally undertaken. Each pseudocomponent is a mixture rather than a pure compound, however, they are treated as pure compounds with specific physical properties in the calculations. The properties involved in characterization include the molecular weight, M, the normal boiling point Tb, the specific gravity g, the critical temperature Tc, the critical pressure Pc, and the acentric factor w. M is used to convert weight fractions from compositional analysis to molar fractions. Tb and g are used for estimating Tc, Pc and w, which are used almost exclusively for EOS modelling. In the following sections, the properties mentioned above will be defined and the experimental methods used to determine the measurable properties will be described;

moreover, correlations between these properties will be discussed, and different characterization methods will be introduced.
Properties

Molecular weight (M). For well defined molecular structures the molecular weight is easily calculated from the atomic masses of the constituents. In the case of the fractions, the average molecular weight can be measured by the following methods: vapour pressure osmometry, freezing point depression, boiling point elevation, gel permeation chromatography and nonfragmenting mass spectrometry (Speight, 2001). The different methods have advantages and drawbacks, which make them suitable for different molecular weight ranges. Vapour pressure osmometry, freezing point depression and boiling point elevation are all based on the assumption that the change in the corresponding properties (vapour pressure, freezing point, and boiling point) in a pure solvent caused by introduction of a solute at low concentration is directly proportional to the concentration of the solute. Gel permeation chromatography, also known as size exclusion chromatography, takes advantage of the difference in elution time between molecules with different sizes. Non-fragmenting mass spectrometry principally provides detailed information of the hydrocarbon types, the formulae and the concentration of all the components in a fraction. Normal boiling point (Tb ). The normal boiling point of a pure substance is the temperature at which the substance changes state from liquid to vapour at a pressure of 1 atm (0.1013 MPa). This property is known and tabulated for all the pure components normally considered in petroleum engineering except for the high boiling substances that do not have a boiling point at atmospheric pressure. For the distillation fractions the normal boiling point is usually chosen to be a certain average value of the temperature interval of the distillation fraction. Density (r) and specific gravity (g). The density r is defined as the mass divided by the volume at a certain temperature and pressure. This can be measured for pure components, fractions and petroleum fluids by different methods. The most widely used techniques are the following: Pycnometry, where the mass of a calibrated volume is determined before and after filling with the substance to be investigated. Pycnometers for high pressure and high temperature are available but have to be calibrated carefully at the relevant conditions. Oscillating tube or vibrating tube densimetry, where a tube of glass, for low pressure, and of metal, for high pressure, is exposed to forced

38

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

oscillations. The resonance frequency is measured and if the apparatus has been calibrated with a fluid of similar density, the resonance frequency of the sample can be converted into its density. The method is convenient for high pressure measurements since it can be connected directly to other high pressure equipment. It should be noted that for measurements at very high pressure and high temperature the calibration has to be repeated frequently. Displacement method, where a known volume is transferred into a vessel that is weighed before and after the transfer. This method can be used at high pressure and high temperature conditions. Specific gravity g, or relative density, is the ratio of the density of a material at temperature T and pressure P to the density of a reference material at reference temperature Tref and reference pressure P : ref [1] g r(P,T )/rref (P ,Tref) ref Specific gravities of oil and gas are measured at standard conditions of 14.7 psia (0.1013 MPa) and 60F (15.6C) with respect to water and air at the same conditions. In the petroleum industry, API gravity is often used: [2] API 141.5/g 131.5 o

where g is the specific gravity of the oil measured at o standard conditions. Figs. 3 and 4 show change of density and of molecular weight, respectively, versus carbon number for different oils. It can be seen that the density varies

considerably with the PNA distribution of the hydrocarbon fraction as compared to the molecular weight. The narrow range of molecular weight for a given carbon number has been used as a basic assumption in Pedersens characterization method (see below). However, other studies show that the molecular weight can also vary considerably (Dandekar et al., 2000). Critical temperature (Tc ). For a pure component the critical point is the point that terminates the vapour pressure curve, while the critical temperature T is the c highest temperature at which the substance can exist simultaneously as vapour and liquid. Therefore, at temperatures above Tc no phase transition can occur between vapour and liquid. For pure components, Tc can be measured in principle and compilations of Tc are readily available (Poling et al., 2000). Critical point measurement of heavy hydrocarbons is hindered by their decomposition temperature. Recently developed techniques, using extremely short exposure to high temperature (a few milliseconds), make it possible to measure the critical temperature of n-alkanes (Nikitin et al., 1994) up to tetracosane (n-C24). Fig. 5 shows how Tc increases with carbon number in the n-alkanes. Critical pressure (P ). Similarly the critical c pressure P of a pure component is the highest c pressure at which the pure component has a vapourliquid phase transition or, in other words, the highest possible vapour pressure of the component. For the same reasons explained above for the critical

Fig. 3. Density versus carbon

1 0.98 0.96 0.94 0.92

number for North Sea oils and condensates (Pedersen et al., 1989a).

density (g/cm3)

0.90 0.88 0.86 0.84 0.82 0.80 0.78 0.76 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 aromatic oils medium paraffinic oils paraffinic oils aromatics aliphatics Katz and Firoozabadi (1978)

carbon number

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

39

GEOSCIENCES

400 380 360 340 320

Fig. 4. Molecular weight versus carbon number for North Sea oils and condensates (Pedersen et al., 1989b).

molecular weight

300 280 260 240 220 200 180 160 140 120 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 aromatic oils medium paraffinic oils paraffinic oils aromatics aliphatics Katz and Firoozabadi (1978)

carbon number

temperature, the critical pressure cannot be measured for large hydrocarbon molecules. Pc decreases with carbon number in n-alkanes after ethane (see again Fig. 5). Acentric factor (w). The acentric factor w is defined as follows: [3] w log(Prsat )T
r

0.7

0.1

where Prsat is the reduced vapour pressure and Tr is the reduced temperature. The value of w reflects a deviation in the vapour pressure compared to the behaviour of ideal spherical molecules (e.g. noble gases) for which w is always zero. The factor w can be explained as the non sphericity of a molecule and generally increases with increasing molecular weight. Tc, P and w are the three basic input parameters for the c cubic equations of state which are widely used in the petroleum industry (see below). Critical volume (vc ). The critical volume vc is the molar volume of a component at its Tc and P . c Experimental values of vc are fewer and less accurate than Tc and P . The volume vc is generally not used as c an input parameter in the EOS method, but it is used by some property estimation semiempirical models, e.g. the Lohrenz-Bray-Clark (LBC) correlation for viscosity estimation (see below).
Correlations between properties

required for the characterization of a narrow petroleum fraction. Sometimes, experimental values of two of them (typically Tb and g, or M and g), are unknown and have to be estimated. TBP data for oils from a similar region can be used as a good reference. The generalized properties up to C45 proposed by Katz and Firoozabadi (1978) can be used as a rough estimation but caution must be taken since g is usually oil-specific. In any case, properties for C45 are still needed if the oil is heavy. Pedersen et al. (1989a, 1992) have assumed a very simple molecular weight correlation for SCN fractions: [4] MC 14i 4
i

where i is the carbon number. The SCN fractions

critical temperature (K)

800 700 600 500 400 300 200 100 5 10 15 20 25

40 30 20 10

carbon number
Fig. 5. Change of critical temperature

Estimation of the missing quantities: g, M and Tb. The specific gravity g, the molecular weight M, and the normal boiling point Tb form the minimum set

and critical pressure with carbon number in n-alkanes.

40

ENCYCLOPAEDIA OF HYDROCARBONS

critical pressure (bar)

50

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

are defined using Katz and Firoozabadis Tb up to C45 and extrapolating the successive values by adding 6 K for each carbon number from C45 to C80. Pedersen et al. also suggested a logarithm distribution for g: [5] gC C+Dlni
i

Characterization

For Cn where the molar distribution is unknown or incomplete, different distribution functions can be assumed. A simple approach adopted by Pedersen et al. (1989a, 1992) is to assume an exponential molar distribution: [7] zC Aexp(Bi)
i

where the two coefficients C and D are determined by a measured or assumed gC and the volume additivity n constraint on the experimental gC
n

[6]

zC MC gC
n n

1
i n

zC MC gC
i i

where i refers to the carbon number, and A and B are two constants determined by the two constraints on zC and MC :
n n

[8] [9]

zC

where z is the molar composition. Another approach to estimating g is to relate g to M through a unique characterization factor (Whitson and Brule, 2000), which can be determined by the constraint in (6). Several empirical correlations employed when estimating Tb are also reviewed in detail by Whitson and Brule (2000). Correlations between Tc , Pc and w. Abundant correlations exist for the estimation of critical properties. The most popular of these include the Lee-Kesler correlations (Kesler and Lee, 1976; Lee and Kesler, 1980), the Riazi-Daubert correlations (1980) and Twus perturbation expansion correlations (1984). All of these correlations express Tc and Pc in terms of Tb and g. At the same Tb, the difference in g reflects the difference in aromaticity. Furthermore, fractions with high aromaticity tend to have higher g, Tc and Pc. In Twus correlations, critical properties of normal alkanes are correlated only in terms of Tb, while g only appears in the perturbation step to correct for the deviation of the critical properties from those of n-alkanes. Acentric factors can be estimated using either the Lee-Kesler correlations or the Edmister correlation (1958). Pedersen et al. (1989a) argued that using Tb as an intermediate variable to estimate Tc and P is not c necessary and expressed Tc and P directly in terms of c g and M. Instead of correlating w, they directly correlate the m parameter in the SRK EOS (see below). Recently, another set of correlations for high temperature high pressure reservoirs were proposed by Pedersen et al. (2002). Due to the intrinsic limitation of any EOS, the true Tc , P and w will not necessarily give a c reasonable reproduction of the vapour pressure for heavy components. Soave (1998) proposed a set of correlations in which the values of Pc are adjusted so that boiling points at both 10 and 760 mmHg can be reproduced.

i n

zC

zC MC
n

i n

zC MC
i

Whitson (1983; Whitson and Brule, 2000) suggested a more general three-parameter gamma distribution. Therefore, after determination of the molar distribution, Tc, Pc and w of each SCN fraction can be evaluated using the correlations introduced in the previous section. The characterized SCN components are usually too numerous and thus are grouped into a number of pseudo components which are manageable for process or reservoir simulations. For reservoir simulations the number is currently limited to around 10 components, while for process simulations it is in general limited to around 50 components, even though for some process calculations it can be feasible to use up to 100 components for the description of the oil. Each lumped fraction can contain several SCN fractions and its properties are obtained using different averaging methods: from the simple weight or molar average to very sophisticated ones (Montel and Gouel, 1984; Leibovici, 1993). As a reversal process of lumping, delumping can be performed to retrieve the detailed compositional information after a simulation with lumped components (Leibovici et al., 1996). PVT experiments In the case of petroleum fluids in reservoirs under pressure, it is essential to know very early in the production phase how the fluid will respond to the pressure reduction that will occur when the reservoir is depleted. A set of simple, yet informative experiments are used to reveal the nature of the fluid. The experiments have the common name PVT experiments since they investigate the relationship between the pressure and the phase volumes at one or more temperatures. During each experiment the temperature is kept constant. Further details on these types of experiments can be found in chapter 4.2.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

41

GEOSCIENCES

PVT modelling
Empirical correlations

[13]

fi 1 ln/i ln 13 1 yi P RT

1 P ni

T,V,nj

RT 1 dV lnZ V

Empirical correlations were extensively used in calculating PVT properties before the widespread application of more theoretically sound models based on EOS. Unlike general EOS models, empirical correlations usually treat gas properties and oil properties as two separate categories. The properties of major concern include the volumetric property and the viscosity for both gas and oil, gas solubility, as well as bubble point pressure and surface tension in the case of oils. An extensive description of these empirical correlations can be found in chapter 4.2.
Basics of phase equilibrium

For a heterogeneous closed system consisting of Nc components and Np phases, it can be proven that the temperature T and the pressure P must be uniform throughout the system and that the chemical potential m of the individual components must be the same in all of the phases at equilibrium: [10] mij mik i 1,,Nc, j,k 1,,Np with j k

where Z is the compressibility factor given by Z Pv/RT, in which v is the molar volume. In principle, all of the thermodynamic properties can be calculated if the necessary volumetric data are available. However, experimental measurement of volumetric data over the full range of the integral in Eq. [13] is unrealistic in most cases. The necessary volumetric data are usually provided by an EOS which describes the mathematical relation between pressure, temperature, volume, and composition. Finally, it should be noted that criteria [10] and [12] are only necessary conditions (and not sufficient) for phase equilibrium since a stable equilibrium state is not only a stationary point but also a global extremum, e.g. dAT,V 0, dGT, P 0.
Equations of state (EOS)

The chemical potential of component i, mi, is usually expressed as ( G/ ni)T,P,nj i o ( A/ ni)T,V,nj i, where V is the total volume, ni is the molar number of component i, and G and A are the Gibbs energy and the Helmholtz energy respectively. In practical equilibrium calculation, a less abstract concept, fugacity, is often used instead of chemical potential. Fugacity, literally meaning escaping tendency, is related to chemical potential by the following definition: [11] mi mi mi fi fi fi RT ln 1 with 13 1 as P 0 fi yi P

Since the introduction of the van der Waals EOS in 1873, numerous EOS have been proposed. General reviews of EOS are readily available in the literature (Anderko, 1990; Wei and Sadus, 2000). The EOS encountered in modelling petroleum fluids roughly fall into three families: vdW (van der Waals) -type EOS, the virial EOS and its modifications, and EOS based on the Principle of Corresponding States (PCS). All vdW-type EOS consist of a repulsive term and an attractive term like their common predecessor van der Waals EOS: a RT [14] P 12 1 v b v2
repulsiv attractive term term

where and are the chemical potential and the fugacity of component i, respectively, at a reference state of the same temperature, and yi is the composition of component i in the mixture. Fugacity can be understood as a corrected pressure. The dimensionless ratio /i fi / yiP is called the fugacity coefficient. The ratio ai fi /fi is known as the activity of component i and gi ai /xi is called the activity coefficient where xi is the molar fraction. If the reference states for different components are at the same temperature, it can be shown that Eq. [10] is equivalent to the equality of the fugacities: [12] fij fik i 1,,Nc, j,k 1,,Np with j k

where b is a co-volume parameter and a/v 2 is an expression for the attractive pressure. Many vdW-type EOS, including the vdW EOS itself, take a cubic form in terms of the volume or the compressibility factor Z and are referred to as cubic EOS. Despite their simplicity and empirical nature, cubic EOS are the most widely used EOS in the petroleum industry. The two most successful examples are the Soaves modification of the Redlich-Kwong EOS i.e. the SRK EOS (Redlich and Kwong, 1949; Soave, 1972) and the PR EOS (Peng and Robinson, 1976), which will be discussed in detail below. The virial EOS can be expressed as an expansion in terms of molar density r: [15] Z 1 Br Cr2

The fugacity and other thermodynamic functions, such as the internal energy and the enthalpy, can be expressed rigorously in terms of volumetric properties, e.g.:

or pressure P: [16] Z 1 B P C P2

42

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

where coefficients B, C, etc. are called the second, third, etc. virial coefficients, and B , C , etc. are coefficients related to B, C etc. Since the third and higher virial coefficients are generally unavailable from experimental data, application of the virial EOS is limited to the low density region and, thus, is of little use in most practical applications. However, the rigorous statistical mechanics basis of the virial EOS has inspired many empirical extensions which are valid over wide ranges of density. Amongst these extensions are the Benedict-Webb-Rubin EOS i.e. BWR EOS (Benedict et al., 1940) and Starlings modification (1973) of this EOS (BWRS EOS). The third class of EOS is based on PCS. The classical two-parameter PCS assumes that two different systems have the same reduced pressure if they are at corresponding states, i.e. if they have the same reduced volume and temperature. Depending on the number of parameters used in defining the corresponding states, there are three-parameter and even four-parameter PCS. The PCS is actually implicated in most EOS, while it is in this third family that EOS are formulated explicitly using the PCS. The EOS in this family usually employ very accurate equations for suitable reference fluids and relate the final fluid properties with the properties of reference fluids by the shape factor method or the perturbation method. The main advantage of this family of EOS is that the properties of the fluid of interest can be accurately reproduced provided the substance is not very different from the reference fluids. The EOS proposed by Mollerup and Rowlinson (1974) for liquefied natural gas and low molecular weight hydrocarbon mixtures, and that proposed by Lee and Kesler (1975), for hydrocarbons and their mixtures, belong to this family. Cubic equations of state: the SRK EOS and the PR EOS. These EOS can be written in a general form: RT a(T) [17] P 11 1111121 n b (n d1b)(n d2 b) where a(T) and b are two EOS parameters. d1 and d2 are constants: d1 1 and6d2 1 for the SRK EOS, while 6 d1 1 2 and d2 1 2 for the PR EOS. Since the critical point of a pure component is an inflexion point of the pressure-volume isotherm, which requires the first and second derivatives to be equal to zero, the two parameters ac a(Tc ) and b can be determined through these two constraints and, finally, expressed in terms of critical temperature and pressure as follows:

[18] [19]

ac Wa R2Tc2/P c b Wb RT /P c c

where Wa 0.42747 and Wb 0.08664 for the SRK EOS, and Wa 0.45724 and Wb 0.07780 for the PR EOS. The temperature dependency of a(T) is expressed by the following functional form: [20] a(T) ac a(Tr,w) where the temperature function a(Tr,w) is assumed to be a function of the reduced temperature Tr and the acentric factor w has the following form: 23 [21] a(Tr,w) [1 m(1 Tr)]2 [22] [23] mSRK 0.480 1.574w 0.175w2 mPR 0.37464 1.54226w 0.26992w2

The function a(Tr,w) can be determined by fitting the pure component vapour pressures. Finally, Eq. (17) can be expressed in terms of the compressibility factor as [24] Z 3 [1 (1 d1)B]Z 2 [A d1B B2(d1 d2)]Z (AB d2 B2 d2 B3) 0 where d1 d1 d2, d2 d1d2 and two dimensionless parameters, A and B, are defined as: A aP/(RT)2 [25] B bP/RT Eq. [24] is in cubic form and can be solved analytically, which is generally believed to be an advantage of the cubic EOS. In fact, numerical solutions using the Netwon-Raphson method can be equally efficient. Quadratic mixing rules, also known as van der Waals one-fluid mixing rules or random mixing rules, are generally used for a and b in an Nc component mixture:
Nc

[26]

b
i 1

xi bi xi xj aij

Nc Nc

[27]

a
i 1j 1

5 where aij aiaj (1 kij) e kij is the binary interaction parameter with kii 0 and kij kji. The value of kij can be obtained by fitting binary vapour-liquid equilibrium data. Usually, kij 0 is a good approximation for most hydrocarbon-hydrocarbon pairs except for C1-C7+ pairs. Fugacity coefficients of component i in a mixture are given by:

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

43

GEOSCIENCES

[28]

fi bi ln/i ln 12 32 (Z 1) ln(Z B) yi P b
Nc

2 xj aij A j 1 1212333 12123 bi ln Z d1B 32 1212 (d2 d1)B a b Z d2 B Expressions for other thermodynamic properties are easily found in the original papers. The above expressions are commonly adopted in phase equilibrium calculations. However, a more systematic and efficient way to formulate these properties exists, especially when a large number of thermodynamic properties, including derivatives of fugacity coefficients, are required (Michelsen and Mollerup, 1986; Mollerup and Michelsen, 1992). The SRK and PR EOS give similar vapour-liquid equilibrium calculation results, however, the major improvement offered by the PR EOS is in the liquid density calculation. This can be attributed to the fact that the critical compressibility factor predicted by the PR EOS (0.307) is closer to the real hydrocarbon value (<0.290) than that determined by the SRK EOS (0.333). On the other hand, volume translation (see below) can improve the density prediction in the SRK EOS and, in many cases, it is required by both types of EOS to obtain acceptable volumetric results. In conclusion, the choice of the SRK or PR EOS is largely determined by personal preference. Improvement of cubic EOS. Modifications of cubic EOS have been performed mainly in two directions. The first is to introduce a better temperature dependency a(T) in order to improve the reproduction of the vapour pressure; the second is to modify the functional form of the EOS in order to improve liquid density prediction. It should be noted that Soaves modification of the RK EOS is actually a modification of the a(T) function. In a similar way Mathias and Copeman (1983), and Stryjek and Vera (1986), have proposed better a(T) functions for the RK and the PR EOS, respectively. Recently, Twu et al. (1995a, 1995b) proposed new a(T) functions for both the RK and the PR EOS. These modifications usually incorporate compound specific parameters which significantly improve the vapour pressure accuracy, especially for polar molecules. However, their effect on supercritical temperatures is still open to question. The PR EOS improves the density calculation by modifying the attractive form in the SRK EOS. Further improvement of density prediction can be achieved by introducing additional EOS parameters. The Schmidt-Wenzel EOS (1980) and the Patel-Teja EOS (1982) are two commonly used three-parameter EOS.

An alternative to three-parameter cubic EOS is the volume translation method which can be used to separate the application of an EOS to vapour-liquid equilibrium calculations from the application to density calculations. The method proposed by Martin (1979), and elaborated further by Peneloux et al. (1982), consists in translating the molar volume from the original EOS, vEOS along the volume axis as follows: [29] nCOR n EOS c

The mixture translation parameter c is calculated from the pure component translation parameters ci through a linear mixing rule:
Nc

[30]

c
i 1

xi ci

where ci can be determined by matching the measured saturated liquid volumes at Tr 0,7 or from different correlations (Peneloux et al., 1982; Jhaveri and Youngren, 1988). ci can also be treated as tuning parameters to match experimental densities. Since vapour volume is generally much larger than liquid volume, the above translation will mainly improve the liquid density without much affecting the gas density. The elegance of this method is found in the fact that the translation will increase the fugacity of component i in all the phases by the same factor exp( ci P/RT) and thus will not affect the phase equilibrium calculations. Also, the same formulation and code for the original EOS can be utilized directly.
Phase equilibrium calculation

Several frequently used phase equilibrium calculations are introduced below. More detailed discussion can be found in the work by Michelsen and Mollerup (2004). T-P flash calculation. In an isothermal flash (or T-P flash) calculation, a feed with a given composition is brought to a specified T and P, the resultant number of phases, and the amount and composition of each phase, then need to be determined. The isothermal flash is probably the most common and important equilibrium calculation. Therefore, robust and reliable algorithms are available for this type of calculation. In many practical calculations, it is either assumed or can be known in advance that there are, at most, one liquid phase and one vapour phase. The two-phase flash calculation for these situations consists of a stability analysis step and a phase split calculation step (Michelsen, 1982a and 1982b). Multiphase isothermal flash calculations also have similar steps, but they require more extensive stability analysis and more demanding phase split calculation. The discussion here is limited to two-phase equilibrium.

44

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

The stability analysis step is carried out as follows: a stable equilibrium mixture of molar composition z is subject to the following TPD (Tangent Plane Distance) criterion:
Nc

[31]

TPD(w)
i 1

wi [mi (w) mi (z)] 0

molar gibbs energy of mixing

which means that the appearance of any infinitesimal amount of a new phase of any molar composition w should always increase the Gibbs energy. Fig. 6 shows the geometric meaning of TPD(w). A tangent plane (line) to the Gibbs energy surface (curve) can be made at composition z, while TPD(w) is the distance from the tangent plane (line) to the Gibbs energy surface (curve) at composition w. Stability requires that all points on the tangent plane are lower than the Gibbs energy surface, in other words, that TPD(w) is always positive. In real calculations, it is impossible and also unnecessary to check all of the compositions. Only the stationary points (possible minima) of the tangent plane distances are checked. The search for stationary points should start from a set of initial estimates representing a possible incipient phase. At high pressures, a priori phase identification is difficult. Therefore, both vapour-like and liquid-like initial estimates are used in the stability analysis for a twophase vapor-liquid flash calculation. Once instability is found in the located stationary point, the fugacity coefficients at this point are used to generate the initial estimates for the equilibrium factors Ki in the phase split calculation step. Ki is defined as the mole fraction ratio of component i between vapour and liquid phases. Whereas the TPD criterion specified by Eq. (31) requires constraint minimization, in practice, it is modified to a form which only needs unconstrained minimization. The phase split calculation is usually started with a successive substitution algorithm. This algorithm consists of an inner loop, which solves for phase amounts and compositions at fixed Ki, and an outer loop, which updates the K-factors using the new phase compositions. The equation solved in the inner loop can be written as follows:
Nc

recommended to switch to a second order minimization of the Gibbs energy. Saturation points and the phase envelope. In the early days, saturation points (the bubble point and the dew point) calculations were thought to be simple and were even used to initiate a flash calculation by checking if the specified condition is situated in the two-phase region. The true situation, however, is just the opposite: calculations of bubble or dew points at elevated pressures are much more difficult. One obvious difficulty is that the number of solutions to the saturation point calculation is not known in advance. For example, there is no dew point at temperatures higher than the cricondentherm while the number of dew points increases to two below the cricondentherm. Furthermore, stability analysis cannot be performed in an easy and reliable manner as in the case of an isothermal flash calculation since a primary variable, T or P, is always missing. Currently, there is no entirely satisfactory method for calculation of saturation points at arbitrary conditions. With proper initial estimates, saturation point calculations can be performed using a partial Newtons method, in which saturation T or P converges faster than composition. Wilson K-factors can be used as initial estimates, however, their accuracy is poor at high pressure or near the critical point. Dew point calculation is generally more difficult than bubble point calculation due to the non-ideality of the incipient liquid phase. A search of the unstable P (or T) range by stability analysis can also be used to initiate the saturation calculation. This procedure
mole fraction

wsp

TPD(w) TPD(wsp)

[32]
i

zi(Ki 1) 11111 0 1 (Ki 1)b 1

and is known as the Rachford-Rice equation. b is the molar amount of vapour phase. The Rachford-Rice equation can readily be solved by the NewtonRaphson method. Successive substitution can be accelerated by the general dominant eigenvalue method. If the iteration does not converge after two or three accelerations, it is

Fig. 6. Illustration of Tangent Plane

Distance (TPD) concept. TPD at composition w for the feed composition z is shown as TPD(w). TPD at the stationary point wsp is TPD(wsp). The dashed tangent at wsp is parallel to the tangent at z.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

45

GEOSCIENCES

obviously needs more computation effort and a priori knowledge of a reasonable search region. A more reliable procedure to locate all saturation points is to perform phase envelope calculation. The phase envelope is a curve in the pressure-temperature plane showing the transition boundary between the vapour and the liquid phase. Further explanation of phase envelopes for multicomponent systems is given in Section 1.1.5. Michelsen (1980) developed an efficient algorithm to construct the entire phase envelope. The calculation starts at readily obtained low pressure saturation points, while the subsequent saturation points are generated by a full Newtons method using initial estimates generated from information obtained in earlier steps. This type of calculation can cross the critical point smoothly. Critical point. A preferable algorithm, proposed by Heidemann and Khalil (1980), expresses the critical point criteria in terms of the Helmholtz energy instead of the Gibbs energy, as had been done in earlier work (Peng and Robinson, 1977). After further improvement by Michelsen and Heidemann (1981), the critical point calculation algorithm is comparable to flash calculation in computation time.
PVT simulation and EOS tuning

All of the conventional PVT experimental results can be simulated based on an EOS model. Simulation of a specific PVT experiment is simply one of, or a combination of, several phase equilibrium calculations described above. Commercial PVT softwares are widely available nowadays. Calculations using cubic EOS with a proper characterization procedure can generally reproduce the experimental data qualitatively. Quantitatively, several percent errors are common in calculated saturation pressures, densities, and mole percent of key components (Whitson and Brule, 2000; Pedersen et al., 1989b). Near the critical region, it is likely that the EOS method will incorrectly identify dew points or bubble points even if there is only a small calculation error. The disagreement between the measured and the calculated results may originate from inaccurate properties deduced from characterization, limitations of the cubic EOS, incomplete compositional analysis, and last but not least, erroneous experimental data. Nearly all PVT reports are flawed to some extent, however, consistency checks before a PVT simulation can help to rule out some poor experimental data. The remaining deviations between predictions and measurements can be reduced further by adjusting the parameters entering the calculation model. This method is known as EOS tuning. Tuning of EOS parameters is not an exact science in the sense that there is no unique way of tuning, and

the engineer decides how much and which parameters to tune based on his experience. However, we attempt to give some guidelines to tuning below: The common tuning parameters are the properties of C7+ fractions including critical properties, acentric factors, their interaction coefficients with methane, and volume shift parameters. Coats and Smart (1986) suggested direct tuning of Wa and Wb, both for methane and the heaviest C7+ fraction. However, the magnitude of tuning is generally large and may cause undesirable results outside the regression region. The properties of the greatest interest in terms of further application and those of higher experimental accuracy should be assigned a higher weighting factor. Saturation pressure, densities, and gas/oil ratio are examples of important properties for reservoir simulation. Coats and Smarts suggestion (1986) of a weighting factor of 40 for saturation pressures, around 10 for densities and 1 for composition can serve as a rough guide to the setting of weighting factors. The number of parameters tuned and the extent of tuning should be as few as possible. Tuning should be focused on the most sensitive parameters, while the insensitive ones can be removed. Furthermore, one should be aware of the errors in experimental data and avoid overfitting. Although automatic non-linear regression is the prevailing technique, common sense should be the guide throughout and manual trial-and-error should also be attempted. The tuning of parameters should correspond to reasonable trends in properties, such as the increase of Tc, Pc with the carbon number. Pedersen et al. (1989b) suggested tuning of the C7+ molecular weight by arguing that the molecular weight measurement usually suffers from an error of 5-10%. The overall composition must be changed after C7+ tuning since the molar composition is actually converted from the original weight composition data using molecular weights. For a sample from a well-studied reservoir, however, the range of molecular weight is relatively certain. A consistent way of tuning kij between the C7+ fraction and methane is preferred, i.e. tuning of the coefficients of the empirical correlation which generates these kij instead of tuning them separately. It should be realized that kij tuning may cause spurious results at other temperatures. Volume shift parameters will not change phase equilibrium calculation. Also, the critical volume will only influence viscosity calculation through the LBC correlation (see below).

46

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

Viscosity, interfacial tension and diffusion


Viscosity

Viscosity is a measure of the resistance which a fluid exhibits to flow. The dynamic viscosity, h, of a Newtonian fluid is defined as the ratio of the local shear stress driving the fluid flow to the velocity gradient in its perpendicular direction. Viscosity is not an equilibrium property. However, the viscosity of a Newtonian fluid is still a state function and has a definite value once the state is fixed. The most common unit of h used in the oil industry is the poise (P), which is equivalent to 0.1 Pas, and its submultiple the centipoise (cP), 1 cP 1 mPas. The kinematic viscosity is the ratio of the viscosity to density and its unit is the stoke (St), where 1 stoke 10 4 m2/s. Fig. 7, a generalized viscosity plot calculated using Lucas corresponding-states viscosity correlation, and Fig. 8, the smoothed viscosity of propane, illustrate how fluid viscosity varies with pressure and temperature. An increase in pressure always increases the fluid viscosity; however, the influence of temperature is different for liquid and gas, and in the latter case depends on the pressure. For a liquid phase, or a gas phase at relatively high pressure, an increase in temperature will reduce the viscosity. For a gas phase at low pressure or a dilute gas, viscosity will increase with temperature. Abundant viscosity models, ranging from highly theoretical ones to simple empirical correlations, are available in the literature. Many of them are only suitable for predicting either the liquid or the gas phase viscosity. However, viscosity calculation for both hydrocarbon gas and liquid using a single model is often required in the petroleum industry, especially in processes involving high pressure phase equilibrium. The following discussion will focus only on this type of model. Reviews of other viscosity models can be found in the literature (Monnery et al., 1995; Poling et al., 2000). The Lohrenz-Bray-Clark correlation (LBC, Lohrenz et al., 1964) is traditionally used in the oil industry. Nevertheless, a number of new viscosity models appeared recently, including the model based on the corresponding states principle (AasbergPetersen et al., 1991), the model based on the graphical similarity between P v T and T h P plots (Guo et al., 2001) and the friction-theory (f-theory) viscosity model (Quiones-Cisneros et al., 2000; Quiones-Cisneros, 2001). Only the LBC correlation and the f-theory model are discussed below. Both the LBC correlation and the f-theory model incorporate the dilute gas viscosity as the lower density limit. Actually, only the viscosity of the dilute

gas can be evaluated strictly based on gas kinetic theory. On the other hand, viscosity models for dense fluids, including the LBC correlation and the f-theory, are generally semi-empirical in form. The Lohrenz-Bray-Clark (LBC) correlation. The LBC correlation is given by: [33] [(h h*)x 10 4]1/4 a0 a1rr a2 rr2 a3 rr3 a4 rr4

where a0 a4 are empirical coefficients, rr vc/v is the reduced density, x Tc1/6M 1/2Pc 2/3 is the viscosityhx
66 56 46 36 26 16 6
100

5
70 50 40 30

3
2.0 1.7 1.4 1.2 1.0 0.8 0.5 15 10 8.0 6.0 3.0 4.0

25 20

Pr 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5

reduced temperature hx
5 4 3 Pr 0 5 6 7 8 9
20 10 50 40 30 70 100

10

11

12

13

14

reduced temperature
Fig. 7. Generalized viscosity plot based on Lucas

corresponding states correlation (Poling et al., 2000). The vertical axis represents the product of the dynamic viscosity h and the parameter x 0.176 Tc1/6M 1/2P 2/3, c where h is given in mP, Tc in K, and P in bar. c The lower part of the figure shows the behaviour for high temperature values.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

47

GEOSCIENCES

reducing parameter and h* is the low-pressure gas mixture viscosity. When the LBC correlation is applied to a mixture, linear mixing rules are used for M, Tc, Pc and vc , and the following special mixing rule is applied to h*:
Nc Nc

EOS) by means of three temperature dependent friction coefficients as follows: [36] hf kr P krr P 2 ka P rep rep attr

[34]

h*
i 1

zi hi*Mi1/2/

i 1

zi Mi1/2

Quiones-Cisneros et al. (2001) further introduced the concept of corresponding states into the model and developed a general one-parameter f-theory. The one-parameter f-theory expresses the friction term of a pure component in the reduced form: [37] hf hf /hc

Lohrenz et al. (1964) also suggest a correlation for the critical molar volumes of the C7+ fractions. However, in practice, the C7+ critical volumes are generally treated as tuning parameters. As can be seen from Eq. [33], the LBC correlation actually expresses h as a 16-degree polynomial in the reduced density and thus is very sensitive to the density calculation results. It tends to underpredict viscosities especially for highly viscous fluids. The friction theory viscosity model (f-theory). The friction theory considers dense fluid viscosity as a mechanical property rather than a transport property. It is not until the dilute gas limit that the kinetic theory of gases becomes important. According to f-theory, the total viscosity h of dense fluids can be separated into a dilute gas term h0 and a friction term hf : [35] h h0 hf where the dilute gas model of Chung et al. (1988) is used to calculate h0. By analogy with the Amontons-Coulomb friction law, the viscosity friction term has been related to the repulsive pressure term P and the attractive term P rep attr in a vdW type EOS (e.g. the SRK EOS or the PR

where hc is the characteristic critical viscosity. hf is related to P and P in a way similar to that shown in rep attr Eq. [36]: [38] P rep hf kr 12 P c P 2 P rep attr krr 12 ka 12 P P c c

where kr, krr, ka are EOS-specific and temperature dependent coefficients. For mixtures, the h0 is calculated by:
Nc

[39]

lnh0

i 1

xi lnh0,i

and the hf is calculated using Eq. [36], while kr, krr, ka are calculated from the following mixing rules: Nc 1 xi hc,i kr,i Nc xi 111 12 [40] kr 0.3P 0.3 i 1 M i 1M c,i
Nc

[41]

ka krr

i 1 Nc

xi hc,i ka,i 111 0.3 M P c,i xi hc,i krr,i 111 0.3 2 M Pc,i

Nc i 1 Nc

xi 12 0.3 M xi 12 0.3 M

[42]

i 1

i 1

Fig. 8. Viscosity of propane at different reduced pressures and temperatures.

Tr 0.3 10,000 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.29

dynamic viscosity (mP)

1,000

100

Tc 369.82 K Pc 42.4953 bar


0 2 4 6 8 10

reduced pressure

48

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

Although the one-parameter f-theory loses some accuracy in comparison to the original f-theory, it reduces the required number of input parameters to a single compound-specific hc. For normal alkanes, hc can be evaluated by the following empirical equation: 1 MP 2/3 c [43] hc 7.9483010 4 1321 Tc1/6 where the units are M in g/mol, P in bar, Tc in K and hc c in cP. The f-theory has also been extended to model viscosities of petroleum fluids (Quiones-Cisneros et al., 2004). Interfacial tension originates from the unbalanced force at the interfacial layer between two phases. For molecules in the surface layer between a low dense gas phase and a high dense liquid phase, the attraction from neighbouring gas molecules is less than that from neighbouring liquid molecules. In macroscopic terms, the result is a tension at the boundary, pointing inside the bulk liquid phase. The tension between a pure liquid and its vapour (or air saturated with its vapour) is referred to as surface tension, while that between two liquids is referred to as interfacial tension. However, both terms have been applied to other situations and a consistent nomenclature does not seem to exist. Here, interfacial tension is accepted as a general term for all situations and surface tension specifically refers to a pure component. Interfacial tension is closely related to the flow behaviour of reservoir fluids at the pore level, e.g. the important capillary function is determined both by the interfacial tension and the pore characteristics. Surface or interfacial tension is interpreted as the tension force on a unit length on the surface, or the Gibbs energy change with unit increase in surface area. Traditionally, the units corresponding to the above two definitions are the dyne/cm and the erg/cm2, respectively. In SI unit, 1 dyne/cm=1 erg/cm2=1 mN/m. Interfacial tension is roughly related to the density difference between two phases and vanishes at the critical point. For a gas-oil system, interfacial tension usually decreases with increasing pressure, whereas the temperature effect depends on the position relative to the critical point. For example, for gas condensates, s is expected to decrease with decreasing temperature, while the opposite is expected for an oil sample. Interfacial tension at high pressure is commonly measured using the pendant drop method, which consists of measuring the shape of a suspended drop
Interfacial tension

of the heavier phase (or a standing bubble of the lighter phase) formed in its equilibrium phase, as shown in Fig. 9. The interfacial tension is related to the drop dimensions by: [44] Drgde2 s 111 H

where Dr is the density difference between two phases, de is the equatorial diameter, and 1/H is a tabulated function of ds /de, where ds is the diameter of the drop measured at the height de above the bottom of the drop (see again Fig. 9). In the case of small interfacial tension near the critical point, laser light scattering techniques are preferred (Fotland and Bjorlykke, 1989; Haniff and Pearce, 1990). Traditional interfacial tension modelling consists of relating surface tensions of pure compounds to various properties, such as density, compressibility and latent heat of vaporization and then extending this concept to mixtures by some arbitrary mixing rules. A more recent attempt is to use a theoretically sound gradient theory and its simplifications. Parachor method. The Macleod-Sugden equation (Poling et al., 2000), which is widely employed to estimate the vapour-liquid interfacial tension, is given by: [45] s1/4
L V P (rM rM ) s

L V where rM and rM are the molar densities of the liquid and vapour phase, respectively. [P ] is known as the s parachor, which is supposed to be temperature independent. Eq. (45) has been extended to mixtures by Weinaug and Katz (1943) using simple molar averaging for the parachor: Nc Nc V xi Ps i rM
i 1

[46]

L s1/4 rM

i 1

yi Ps i
L V Ps i (xi rM yi rM )

Nc
i 1

ds

de

de

Fig. 9. The principle of the pendant drop method

for the measurement of the surface tension.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

49

GEOSCIENCES

The above equation is widely used in the petroleum industry. Parachor values can be estimated by group contribution methods. Ali (1994) also reviews the parachor values reported by different investigators. For homologous hydrocarbons, an almost linear relationship with molecular weight is usually found. However, in real oils, the PNA distribution varies in various SCN groups and the linearity is distorted. Firoozabadi et al. (1988) give an equation that can be used to approximate the parachor of pure hydrocarbons from C1 through C6 and for C7+ fractions. It should be noted that since s is proportional to L V Ps 4(rM rM )4, the result obtained from Eq. [45] is very sensitive to the density calculation and the choice of parachor values. Other modifications of Weinaug and Katzs parachor method have been made, including modification of the exponent 1/4 on the left-hand side of Eq. [46] and improvement of the parachor mixing rule. However, Firoozabadi et al. (1988) comment that for reservoir fluid without asphaltene, the WeinaugKatz correlation is sufficient. Finally, it should be noted that generalized correlations are not expected to provide a reliable parachor value for the oil heavy end, which generally contains a high concentration of asphaltic and surface active materials. Therefore, it is advisable to determine it experimentally. The gradient theory and its simplification. The gradient theory (Carey et al., 1978, 1980; Guerrero and Davis, 1980) assumes that there is a continuous density variation for each component across the interface between two equilibrium bulk phases. The density profiles of all the components must take specific forms in order to minimize the Helmholtz energy at fixed T, V and N. Interfacial tension is by definition the partial derivative of the minimized Helmholtz energy with respect to surface area. After some derivation, it can be expressed in terms of the density profiles of all of the components:

proposed a linear gradient theory, which assumes that the number densities of each component are linearly distributed across the interface. These authors have applied this model to a variety of mixtures including hydrocarbon-water mixtures.
Diffusion and thermodiffusion

Diffusion describes the process of relative motion of different components in a mixture in the absence of mixing. Diffusion fluxes are determined as relative fluxes of different components in a mixture. For example, if Ji is the molar flux of component i in a mixture of Nc component, then one determines the molar convective flux using (Haase, 1969):
Nc

[48]

Jc

i 1

Ji

and the diffusive flux of component i using: [49] JD,i Ji zi Jc

where zi is the molar fraction of component i. The mass flux of each component can be similarly defined. Ficks law expresses the molar diffusion fluxes in terms of the gradients of the molar fractions:
Nc 1

[50]

JD,i r

k 1

Dik zk

[47]

s
i j

C(n)

dni 11 dx

dnj 11 dx

dx

where C(n) is the influence parameter for inhomogeneous fluid, ni is the number density of component i and x is the distance from the interface. When combined with a specific EOS to calculate , the density profiles are first calculated by solving a set of partial differential equations or algebraic equations. The interfacial tension is then readily obtained using Eq. [47]. To avoid the difficulty in the density profile calculation, Zuo and Stenby (1996a, 1996b, 1998)

where r is the molar density and Dik is the Fickian molar diffusion coefficient of component i. Although D12 D21 holds for binary systems, it is seldom pointed out that Dij is not necessarily equal to Dji for n 3. On the other hand, the study on multicomponent diffusion is rare, both experimentally and theoretically, in contrast to the fact that large databases of experimental values, and many models and correlations for diffusion coefficients exist for binary mixtures (Hsu and Chen, 1998; Poling et al., 2000). For dilute gases, diffusion coefficients can easily be evaluated based on the strict framework of gas kinetic theory. For liquids, only approximate models are available and their prediction ability is difficult to verify due to the lack of experimental data on diffusion coefficients in multicomponent mixtures. The free volume theory has been used in modelling liquid diffusion coefficients for a long time (Hirschfelder et al., 1954; Bondi, 1968) and recent models for multicomponent diffusion are available (Wesselingh and Bollen, 1997). In engineering practice, effective diffusion coefficients estimated from binary diffusion coefficients are often used (Whitson and Brule, 2000). Estimation of liquid diffusion coefficients is believed to have only orders of magnitude accuracy.

50

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

Thermodiffusion, or the Soret effect, which is induced by temperature gradients, is analogous to diffusion induced by concentration gradients. Thermodiffusion is often negligible but it can be important under specific situations, e.g. the compositional grading of thick oil reservoir (Georis et al., 1998). Accurate measurements of thermodiffusion are rare since they are easily ruined by the natural convection which occurs under the gravity environment.

Paraffin waxes in petroleum production

1.1.4 Heavy petroleum fraction


The most complex chemistry of petroleum is found in the heavy fraction. This part contains several classes of compounds. The detailed molecular structure cannot always be determined and it may be necessary to define a group of compounds as a solubility class rather than a class of molecular structures. From a technological point of view the most important heavy fractions are waxes and asphaltenes, but resins also play a significant role. These three classes of compounds will be defined and described below. Wax
Definition and chemical structure

Waxes are crystalline materials separated from petroleum by cooling a mixture of the oil, a ketone, and another polar or aromatic solvent. Several different solvent types and solvent pairs are used in the literature, such as acetone-toluene, methylisobutylketone (MIBK)-methylene chloride, etc. Due to the differences in the solvent pairs and other conditions the wax content will vary according to the method employed, and may even contain co-precipitated and entrapped oil and asphaltenes. As the separation temperature decreases the amount of solid material recovered increases. The term paraffins is often used synonymously with waxes since the long chain normal alkanes largely dominates the composition of waxes. Most waxes consist of normal alkanes with a carbon number in the range from 20 to 50, but there is discussion in the scientific community whether a different class of waxes exists, namely, the so-called microcrystalline waxes. To avoid confusion the term paraffin waxes will be used in the following. Branched alkanes do not fit into the crystal structures formed by the normal alkanes and, therefore, are not found in paraffin wax. As a result, the class of molecules are generally very well defined when paraffin wax problems have to be addressed.

The fraction referred to as paraffin waxes is important because it represents a potential production or transport problem in relation to oil and gas condensate production. As commercial products, paraffin waxes are valuable, for instance, as lubricants; however, this will not be covered further in this section. Paraffin wax formation in a petroleum fluid is observed as a solid phase that appears when the fluid is cooled down. The temperature at which the first paraffin wax is observed is called the Wax Appearance Temperature (WAT). In Fig. 10, a temperature-pressure diagram is shown for a petroleum fluid where both the solidliquid phase transition boundary and the vapourliquid phase envelope are reported (for a detailed discussion of the latter, see Section 1.1.5). The slope of the solid-liquid curve is rather steep, which indicates that the WAT is neither very sensitive to the pressure nor to the dissolved gas in the oil. If the WAT is higher than the process or transport temperature there will be precipitation and deposition of paraffin wax in the process equipment or transport pipelines. Therefore, the control of paraffin wax is covered by the term flow assurance. Since deposited paraffin wax can only be removed mechanically or by heating, it is crucial that the deposition of paraffin wax does not get out of control. Off shore pipelines are particularly vulnerable since the costs of remediation are considerable. The level of paraffin wax deposition in pipelines is controlled mechanically by pigging; a process in which a tool with a diameter slightly smaller than the inner pipe diameter also known as
700 600 500

pressure (bar)

400 300 200 100 0 0 100 200 300 400

model data

solid liquid solid liquid vapour

liquid

liquid vapour
500 600 700

temperature (K)
Fig. 10. Entire phase envelope

with consideration of the wax precipitation (Lindeloff et al., 1999).

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

51

GEOSCIENCES

the pig is pushed through the pipeline. The pig scrapes the paraffin wax off the pipe walls and pushes it to the end of the pipeline where it is collected. Another way to secure flow assurance is the use of chemical additives that affect the paraffin wax precipitation, or the growth or agglomeration of paraffin wax crystals. Such chemicals can keep the paraffin wax dispersed as small crystals that will be carried along with the flowing oil. The basic mechanisms are not completely understood and the use of these chemical additives requires experimental testing for each specific oil. Different oils will contain a varying amount of paraffin wax components but since accumulation over time will occur in pipelines, even small amounts of paraffin wax in an oil can represent a significant risk for the flow assurance.
Experimental determination of the wax appearance temperature

Modelling of wax precipitation

Accurate thermodynamic modelling of paraffin wax precipitation is possible. The requirement is that the balance between the paraffins, naphthenes, and aromatic compounds is known for the fraction of the fluid in which the paraffin wax components are present. The fugacity equality condition for the solid-liquid equilibrium between solid wax and liquid hydrocarbon can be written as [51] xiS giL fiL 1 1 12 exp xiL giS fiS niL niS 112 dP P RT
P

where xi is the mole fraction, giis the activity coefficient, fi is the fugacity at the reference state (T, P), and vi is the molar volume. The superscripts L and S represent the liquid and solid phases, respectively. The exponential term in Eq. [51] corrects the effect of pressure on the fugacity. The fugacity ratio fiL/fiS can be expressed as: [52] fiL ln 11 fiS Dh 11f RT T 1 1 Tf ht 11 RT T 1 1 Tt

A number of methods can be used to determine the WAT, including differential scanning calorimetry, visual observation using microscopy, cross polar microscopy, light transmittance or scattering, viscometry, cold finger, and filter plugging. Only the microscopy, light transmittance and filter plugging methods are briefly introduced below. In general, microscopy is only used at atmospheric pressure. A small fluid sample is studied under the microscope while being either heated or cooled. Since subcooling will occur, the most accurate measurements are obtained when a cold fluid sample is melted. The WAT can be identified as the temperature at which the last crystal melts. Since paraffin wax is a multicomponent mixture, the melting or freezing will take place over a range of temperatures and not at a single temperature as for a pure substance. Laser light of different wavelengths can easily be transmitted through petroleum fluids if a small light path is chosen. If the fluid is cooled uniformly, and is mixed well, the WAT can be determined by the fact that the transmitted light is disturbed by scattering from the paraffin wax crystal in the fluid. This method has been developed for high pressure applications. In the filter plugging method, petroleum fluid is pumped or circulated through a filter while the whole setup is cooled down in a controlled manner. The WAT can be detected by a sudden resistance to the flow through the filter. This method has also been developed for use at high pressure. It is possible to design the setup in such a way that the paraffin wax on the filter can be collected and its composition analysed.

Tf Tf C 1 p 1 1 ln1 222 R T T where Dhf is the molar heat of fusion or melting enthalpy at the melting temperature of the solute Tf , Dht is the molar enthalpy change of phase transition at the temperature of phase transition Tt and DCp is the difference between the heat capacity of the liquid phase and that of the solid phase. The Dht term is important in modelling wax precipitation since the solid-solid phase transition usually occurs below the melting temperature for heavy n-alkanes. The contribution from the DCp term is usually small and often neglected. Eq. [51] and Eq. [52] provide the framework of modelling wax precipitation in terms of the activity coefficients model. Early wax models (Won, 1986; Pedersen et al., 1991; Coutinho and Stenby, 1996; Coutinho et al., 1996) have been developed mainly for low pressure applications. The key problem lies in proper modelling of the activity coefficients in addition to reasonable estimation of Tf , Tt , Dhf and Dht for the n-alkanes in the oil. Coutinho et al. (1995) gave a comprehensive evaluation of the different activity coefficient models utilized in the prediction of alkane solid-liquid equilibria. To model the precipitation in live oils, where the effects of pressure and dissolved gas are present, it is convenient to modify the above framework by modelling the vapour and liquid phase using an EOS (Pedersen, 1995; Lindeloff et al., 1999; Pauly et al., 2000; Daridon et al., 2001). In this method the liquid

52

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

activity coefficients terms are replaced by liquid fugacities. Asphaltene and resin
Definition, composition and structure

Asphaltenes are operationally defined as the solid material precipitated from crude when mixed with an excess of a low boiling hydrocarbon, such as n-heptane or n-pentane. This fraction should be soluble in toluene or benzene (Fig. 11). The amount of asphaltene separated is a function of the n-alkane carbon number, the time of contact (especially for viscous heavy oils), the ratio of oil to solvent, and the temperature. Although some standards, such as the Institute of Petroleum (IP) 143, are prevailing in the determination of the asphaltene content of crudes, most of the above separation conditions are often varied in the literature on asphaltene research. This leads to some additional confusion regarding the actual behavior and nature of this fraction, as the gross composition of the asphaltene will vary considerably. The number of different compounds inside the asphaltene fraction is estimated to be over 100,000. These compounds are characterized by high molecular weight, polarity and aromaticity. Despite the complexity of asphaltene fractions, their ratio of hydrogen to carbon varies only in a narrow range, say, around 1.0 to 1.2. The heteroatoms N, S, O and trace metals are concentrated in the asphaltene fraction, while their quantity can vary from oil to oil in almost one order of magnitude. Sulphur is usually the most concentrated among all heteroatoms.

The molecular weight of asphaltene is directly related to the molecular size of asphaltene and is a crucial parameter in any asphaltene model. Unfortunately, its experimental determination is difficult. Asphaltenes tend to self associate even in dilute solution. The mechanism is not well understood but hydrogen bonding and charge-transfer are believed to be responsible. The association is influenced by solvent polarity, asphaltene concentration, and the temperature at which the determination is made (Speight, 1999). Methods like vapour pressure osmometry are only able to provide the molecular weights of aggregates and the measured values are dependent on the above mentioned factors. Other methods have their own specific problems. In general, none of the current technologies can answer the challenge of determining the true molecular weights of asphaltenes. Values from different methods differ significantly, even by orders of magnitude. Not much is known about the asphaltene structure. Investigations have shown that asphaltenes consist of condensed aromatic nuclei that carry alkyl and alicyclic systems with heteroelements scattered throughout in various locations (Speight, 1999). However, the formulation of the individual molecular structure is still impossible. An example of a hypothetical asphaltene molecule is given in Fig. 12. Different types of structures have appeared in the literature. Generally, resins refer to the more strongly adsorbed oil fraction in a deasphaltened oil which is subject to subdivision using a surface active

feedstock n-heptane

insolubles benzene or toluene

deasphaltened oil

silica or alumina insolubles asphaltenes 1. heptane 2. benzene or toluene aromatics 3. benzene and methanol resins

carbon disulphide or pyridine

carboids (insolubles)

carbenes (solubles)

saturates

Fig. 11. Simplified representation of the separation of petroleum into six major fractions according to the solubility in different

solvents (n-heptane, benzene, etc.) or the adsorption on solid substrates (Speight, 1999).

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

53

GEOSCIENCES

material, such as fullers earth or alumina. The absorbed resin is recovered by a more polar solvent and, therefore, eluted last (see again Fig. 11). Resins are believed to have molecular structures similar to asphaltenes but with longer alkyl non-polar chains and smaller aromatic rings. The molecular weight of resins is in the range of 100-600 u, according to vapour pressure osmometry measurements (Speight, 1999). These resins can be considered to be a transition type between asphaltenes and the simpler fractions, and hence an overlap between resins and asphaltenes is expected. Resins play an important role in dispersing asphaltenes in crude oil (Andersen and Speight, 2001). It is generally believed that asphaltenes are dispersed in crude oil as centres of micelles surrounded by resins and other lighter and less aromatic molecules. The transition from the asphaltene centres to the bulk phase oil is gradual, with the size and aromaticity decreasing outwards from the aggregate. The concept of Critical Micellar Concentration (CMC) has been widely used in describing the association of asphaltenes. However, recent studies suggest a step-wise mechanism rather than the formation of finite-size micelles (Andersen, 1994; Acevedo et al., 1999; Groenzin and Mullins, 2000).
Asphaltene precipitation

complex, and the possibility of precipitation is not necessarily proportional to the content of the asphaltene. For example, oils with a smaller percentage of asphaltene may be more likely to precipitate. Both depletion and gas injection can trigger precipitation of asphaltene. The reason is usually attributed to unfavourable changes in solvent properties with a change in pressure or the addition of injection gas. Experimental data on the onset of asphaltene precipitation are scarce in the literature. The data can be measured by different methods, e.g. visual observation through microscopy, light transmission or scattering, electrical conductivity, viscometry, and capillary flow measurements. The modelling of asphaltene precipitation is less developed in comparison to the modelling of wax precipitation. Among the major difficulties is the lack of understanding of the actual mechanisms responsible for the precipitation. Moreover, little is known about the true dispersion state of asphaltene in oil, while data on the physical properties of asphaltene are virtually unavailable. There is even debate on whether the precipitation is a reversible process. The limited experimental data make it more difficult to evaluate the existing models.

1.1.5 Reservoir fluids


This section describes the different types of petroleum reservoir fluids which are commonly encountered by exploration. A proper classification of a reservoir is important since the fluid type is a deciding factor in the production scheme. The properties of interest also differ between different types of reservoirs. As will be seen below, the classification of reservoir fluids is closely related to the Pressure-Temperature (P-T) phase diagram of multicomponent mixtures. The P-T phase diagram of a multicomponent system
mole S C 84.9% H 8.2% N 1.0% O 1.2% S 4.7% H/C 1.15 molecular weight: 1,370 Fig. 13 shows a typical pressure-temperature phase diagram of a multicomponent system with a specific overall composition. In the phase diagram, the two-phase region is enclosed by a continuous boundary called the phase envelope. This phase envelope consists of a bubble point branch and a dew point branch, with the two branches connected at the critical point where the coexisting liquid and vapour become identical in terms of intensive properties. In contrast to a single component system, the critical point of a multicomponent system is generally not

The importance of asphaltene is unfortunately justified by the many problems it causes. Apart from poisoning of catalysts and destabilization of products in the refining process, the major problem caused by asphaltene in oil production and transport is its precipitation. Asphaltene precipitation is
OH N S

Fig. 12. A hypothetical asphaltene molecule (Speight, 1999).

54

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

the maximum temperature and the maximum pressure of the two-phase region. The maximum temperature point and the maximum pressure point are called the cricondentherm and the cricondenbar, respectively. In the two-phase region, a set of quality lines (dashed lines) can be drawn to indicate constant percentage liquid volume. The bubble point and the dew point branches correspond to the two special quality lines with 100% and 0% liquid, respectively. All of the quality lines converge to the critical point. The single phase region can be further divided into a liquid region and a gas region, where the former refers to the region above the bubble point curve with temperatures lower than the critical temperature and the latter refers to the remaining part of the diagram. A special phenomenon called retrograde condensation can be observed for mixtures due to the existence of two dew point pressures at temperatures higher than the critical temperature but lower than the cricondentherm. As shown in Fig. 13, both a low pressure dew point D1 and a high pressure dew point D2 exist at temperature T1. For a depletion process A D2 B C D1 at constant temperature T1, an infinitesimal amount of liquid will form when the system reaches D2, and a further decrease in pressure will increase the liquid volume fraction until it reaches its maximum at B. The increase of the condensed liquid (usually called the liquid dropout) with decreasing pressure is contrary to the common intuition and named retrograde condensation. It should be noted that B is only the maximum in the liquid volume fraction rather than the absolute
Fig. 13. Typical pressure-

amount of the liquid dropout. The absolute liquid dropout can still increase after B but the increase is not enough to counteract the expansion effect so that the liquid volume fraction decreases. After C, further decrease in pressure will reduce the absolute liquid dropout due to the re-vaporization. The decrease continues until an infinitesimal amount of liquid is left at D1. Classification of reservoir fluids Petroleum reservoirs are described or classified based on the composition of the petroleum fluid, and the temperature and pressure prevailing in the reservoir. However, the dimensions of a petroleum reservoir are large and the composition of the petroleum fluid varies depending on the location in the reservoir due to the influence of gravity, the geothermal gradient and the geologic history of the reservoir. The classification, into the categories described below, is therefore a simplification which should be kept in mind. When a petroleum reservoir is discovered, it is common practice to classify the reservoir into one of the following types: natural gas, gas condensate, volatile oil or black oil, based on the location of initial reservoir conditions with respect to the phase envelope of the reservoir fluid. Natural gas is further subdivided into wet gas and dry gas, while black oil will, here, also cover heavy oil and extra heavy oil. The classification has a direct impact on the economical and technical evaluation, and thereby on the production strategy. For example, some aspects of the fluid properties, such as volumetric properties, are

temperature phase diagram for a multicomponent system. The quality lines can be more evenly spaced for other systems.

single phase region (liquid)

single phase region cricondenbar (gas)

A D2

critical point pressure % liquid


100 80 30 20 15 10 5 2 1 0

B C two-phase region cricondentherm

D1

temperature T1

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

55

GEOSCIENCES

Table 2. Typical molar composition of petroleum reservoir fluids Component N2 CO2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ C7+ molecular weight C7+ specific gravity Dry gas 0.0292 0.0107 0.8763 0.0477 0.0185 0.0049 0.0058 0.0020 0.0020 0.0029 Wet gas 0.0150 0.0100 0.9010 0.0420 0.0210 0.0020 0.0015 0.0021 0.0024 0.0020 0.0010 128 0.750 Gas condensate 0.0050 0.0393 0.7257 0.0850 0.0493 0.0102 0.0105 0.0035 0.0045 0.0050 0.0620 182 0.807

Volatile
oil 0.0161 0.0217 0.6065 0.0796 0.0471 0.0210 0.0220 0.0210 0.0110 0.0190 0.1350 199 0.802

Black oil 0.0070 0.0200 0.3300 0.0660 0.0700 0.0240 0.0280 0.0180 0.0170 0.0290 0.3910 249 0.853

important for the estimation of the petroleum reserves. Others are of importance for the flow assurance. Potential production problems have to be addressed as early as possible. Table 2 shows the typical molar composition of petroleum reservoir fluids and the corresponding calculated phase envelopes are shown in Fig. 14. It should be noted that the same initial reservoir conditions and separator conditions are used for all fluids in Fig. 14, whereas in fact, they vary from field to field.
Dry gas and wet gas

very common in reality, and at low temperature one or more solid phases will usually form. These points are not considered in the two-phase envelopes in Fig. 14.
Gas condensate

For both dry gas and wet gas, the initial reservoir temperature is higher than the cricondentherm of the reservoir fluid and the reservoir conditions never fall inside the two-phase region during a depletion procedure (the vertical line in Fig. 14). Furthermore, for a dry gas, the process conditions are typically outside the two-phase envelope. Compared with a dry gas, a wet gas contains a larger fraction of C2-C6 components, and its phase envelope is larger and reaches higher temperatures. The separator conditions are inside the two-phase region for a wet gas. The typical gas/oil ratio of a wet gas is 60,000-100,000 scf/stb (11,000-18,000 sm3/m3) and its stock tank oil is usually water-white with a specific gravity higher than 60 API. It can be seen that there is no gas-liquid critical point on the phase envelope of the wet gas in Fig. 14. This is due to the three-phase equilibrium at the low temperature. Actually, phase envelopes in real hydrocarbon mixtures can be much more complicated than the phase envelops shown in Fig. 13. For example, multiphase equilibria as well as multiple critical points or no critical point are

The initial temperature of a gas condensate reservoir lies between the critical temperature and the cricondentherm of the reservoir fluid. In such a case the fluid exists as a gas at initial conditions, but during pressure depletion, liquid will drop out due to retrograde condensation below the high pressure dew point. A typical gas condensate has a gas/oil ratio in the range 8,000-70,000 scf/stb (1,400-12,000 sm3/m3). The stock tank oil is usually slightly coloured and its API gravity is usually greater than 50.
Volatile oil and black oil

In the case of both volatile oil and black oil, the initial reservoir temperature is below the critical temperature of the reservoir fluid. The fluid therefore exists as a liquid at initial conditions, however, on pressure depletion, the bubble point will eventually be reached. Volatile oil differs from black oil quantitatively only in terms of having a relatively larger fraction of light and intermediate hydrocarbon components. As a result, its critical temperature is much closer to the reservoir temperature. During the depletion, volatile oils show a larger shrinkage factor due to the large amount of gas released. The typical gas/oil ratio for a volatile oil is 2,000-3,500 scf/stb (360-620 sm3/m3). Its stock tank oil is usually greenish to orange in colour with an API gravity between 45 and 55.

56

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

The gas/oil ratio for a black oil is usually below 700 scf/stb (120 sm3/m3) and the stock tank oil colour can range from brown to black with an API gravity between 15 to 40. Black oils with a gas/oil ratio less than 200 scf/stb (36 sm3/m3) and an API gravity less than 15 are sometimes known as low shrinkage oils.
Heavy oils and extra heavy oils

condensate, as shown in Fig. 15, a small pressure drop below the dew-point will give rise to a large amount of liquid dropout. Compositional grading A reservoir at static equilibrium can show significant compositional variations known as compositional grading. Compositional grading is mainly due to gravity, as the temperature gradient in a reservoir is usually small (0.025 K/m), and therefore plays a less important role. However, the temperature can significantly influence the fluid distribution in some specific situations (Georis et al., 1998). The isothermal compositional grading can be described by [53] mi (P , zref ,T ) Mi ghref mi (P, z,T ) Mi gh ref i 1,, Nc

The definition of heavy oil is quite arbitrary. Generally, it refers to petroleum with an API gravity of less than 20 and usually, but not always, a sulphur content higher than 2 wt% (Speight, 1999). The term heavy oil has also been arbitrarily used to describe both heavy oils that require thermal stimulation of recovery from the reservoir and bitumen in bituminous sand (tar sand) formations from which the heavy bituminous material is recovered by a mining operation. Speight (1999) gives a detailed discussion of the difference between the terms heavy oil, bitumen, mineral wax, asphaltite, asphaltoid, and bituminous sand.
Near critical fluid

A reservoir fluid with its critical temperature very close to the reservoir temperature is called a near critical fluid. Such a near critical fluid can be either a gas condensate or a volatile oil, and is characterized by a dramatic change in the liquid volume fraction just below the saturation pressure. For the near critical gas

where mi is the chemical potential of component i, Mi is the molecular weight, zref is the known composition of a single phase fluid at reference depth href with reference pressure P , and P and z are the pressure ref and composition at depth h. In general, the heavy components tend to accumulate at the bottom while the light components tend to accumulate in the top. Therefore, saturation pressure and solution gas/oil ratio can change considerably as they are associated with compositional change. Fig. 16 shows two typical

50

reservoir conditions gas condensate phase envelope gas-liquid critical point three-phase point

40

Fig. 14. Phase envelopes for different types of reservoir fluids. The dashed line indicates the depletion line for an isothermal reservoir. Fluids with critical points to the right of the depletion line belong to oil reservoirs, while the rest belongs to gas reservoirs.

pressure (MPa)

30

volatile oil

20

wet gas
10

black oil separator conditions reservoir temperature

dry gas
0 100 0 100 200 300 400 500

temperature ( C)

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

57

GEOSCIENCES

profiles of saturation pressure and reservoir pressure. On both profiles, the dew point pressure increases with depth (i.e. with an increasing amount of heavy components) and the bubble pressure decreases with depth (i.e. with decreasing methane content). The reservoir pressure increases with depth and shows different slopes in the gas and oil regions. The difference between two profiles is that in the left-hand profile, the saturation pressure reaches the reservoir pressure at a depth where a clear Gas Oil

Contact (GOC) is formed, whereas in the right-hand profile, the saturation pressure is always below the reservoir pressure and the transition from oil to gas takes place through a local critical point. Compositional change is discontinuous at the GOC in the left-hand profile while that in the right-hand profile is always continuous.

1.1.6 Formation water


Water is always associated with petroleum hydrocarbons in their production. Not only is water found almost invariably in reservoirs in appreciable quantities as connate, interstitial, or formation water, but it is also deliberately injected to improve oil recovery. Water interacts with hydrocarbon fluids in terms of flow, expansion, and sometimes dissolution and release of soluble gases. Furthermore, solid precipitation including scaling and hydrate formation due to temperature and pressure change may cause considerable damage during the production process. Knowledge of the physical properties of the formation water and its phase equilibrium is therefore important to petroleum engineers. Some of these aspects are discussed in the following paragraphs. Further details on the characteristics of the layer waters (salinity, solubility of gas in the water, volumetric parameters and viscosity) are reported in chapter 4.2.
Composition

2,600 2,400

pressure (psi)

2,200 2,000 1,800 1,600 1,400 60

100% % 90% 80% 70% 60% 50% 40%

d by liqui

volume

critical point

0%

30% 20% 10%

80

100 120 140 160 180 200 220 240 260

A a 2,700

temperature ( F)
166 F

171 F

169 F

2,600
19

5 F

2,500 2,400 2,300 2,200

14

21 2

18 1 F

11

2 10

F F 85

pressure (psi)

2,100 2,000 1,900 1,800 1,700 1,600 1,500 1,400 1,300 1,200 0 10 20 30 40 50 60 70 80 90 100

Formation water contains dissolved salts (primarily sodium chloride) and dissolved gases (primarily methane and ethane). It bears little relationship to sea water although both are generally called brine or salt water. The formation water has much higher salinity, ranging from 200 to 300,000 ppm (around saturation), compared with around 35,000 ppm of sea water. The cations dissolved in formation waters usually include Na , Ca2 and Mg2 , and occasionally K , Ba2 , Li , Fe2 and Sr2 . The anions commonly 2 2 include Cl , SO4 and HCO3 , while CO3 , NO3 , 3 and S2 are also often present. Br , I , BO3 Finally, the concentration of salts in formation water is called salinity and can be specified using different quantities. Table 3 gives definitions of some of the most commonly used quantities.
Volumetric property of formation water

B b

percent liquid by volume

Fig. 15. P-T phase diagram for

a near critical gas condensate fluid (A) and its corresponding volume isotherms (B) (Katz et al., 1959).

Unlike water density, which is only a function of temperature and pressure and can be easily determined with high accuracy, formation water density is also influenced by the amount and composition of dissolved salts and gases. The large range of dissolved

58

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

Fig. 16. Typical saturation

pressure and reservoir pressure change with depth. depth

gas

w de

gas

depth

gas/oil contact
re

res

su re

su

bb le p

bu bb le

bu

pre s

components brings difficulty in developing a general model. Fortunately, the volumetric property of formation water varies within a relatively small range, with the isothermal compressibility being typically 2-4 10 6 psi 1 (3-6 10 10 Pa 1) and the formation volume factor, which is defined as the volume of the formation water at reservoir conditions divided by the volume of the water produced from the formation water at surface conditions, typically ranging from 1.00 rm3/m3 at high pressure to 1.07 rm3/m3 at low pressure. Engineering correlations (McCain, 1990; Whitson and Brule, 2000) can give reasonable estimations in practice when experimental data are not available. These engineering correlations often simplify the dissolved salts to NaCl or neglect the salt effect on the formation volume factor.
Mutual solubility of formation water and hydrocarbons

The solubility of hydrocarbons in water is very low, while that of methane in water is no more than 1 mol% at common reservoir conditions. From methane to propane, the solubility decreases by a factor of two to three on going from one hydrocarbon gas to the next

Table 3. Different commonly used expressions

w de re ssu pre
res
oil

re ssu pre
critical point

o ir p r e s er v
r e ss

er es pr ir vo re su

ur e

oil

pressure

pressure

of salinity.
Term Molality Molarity Weight percent Parts per million (ppm) Milligrams per liter Definition moles of solute mass of pure water (kg) moles of solute volume of brine (l) mass of solute 100 mass of brine mass of solute (g) mass of brine (t) mass of solute (mg) mass of brine (l)

heavier one, and the solubility of heavier liquid hydrocarbons is much smaller. Furthermore, the solubility of hydrocarbons increases with pressure, this effect being more prominent for lighter hydrocarbons and at low pressures. On the other hand, at constant pressure, the solubility of hydrocarbons has a temperature minimum, before which it decreases with temperature and after which it increases. Within common reservoir temperature and pressure ranges, an increase in hydrocarbon solubility with temperature is usually observed. Finally, the solubility of CO2 in water is almost one order of magnitude larger than that of methane. The equilibrium water content in hydrocarbons is generally larger than the corresponding hydrocarbon solubility in water, especially at low pressure and high temperature. Therefore, in the case of liquid hydrocarbons, the water content can be orders of magnitude higher. However, since the water content decreases with pressure and increases dramatically with temperature, the water content in light hydrocarbon gases can be lower at low temperature and high pressure. Fig. 17 illustrates the mutual solubility of methane and water within the usual reservoir temperature and pressure ranges. The presence of salt reduces both the solubility of the hydrocarbons and the equilibrium water content. The plot of methane solubility in water prepared by Culberson and McKetta (1952) is commonly used to estimate the solubility of natural gas in water. Application of the salinity correction can be performed subsequently by introducing salting out coefficients (Whitson and Brule, 2000). The solubility of CO2 in brine can be estimated by the empirical correlation of Chang et al. (1996), while that of water in natural gas and hydrocarbon liquids can be estimated using empirical charts (Hoot et al., 1957; McKetta and Wehe, 1962; GPA, 1980). Cubic EOS with the conventional quadratic mixing rules cannot simultaneously correlate the

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

59

GEOSCIENCES

10-1

mole fraction

200 C 150 C

10-2 104.44 C 10-3 71.1 C 37.78 C 10-4 0 200 400 600 800 1,000

pressure (bar)
Fig. 17. Solubility of methane in water

(solid lines) and water content in methane (dashed lines) (Olds et al., 1942; Culberson and McKetta, 1951; Sultanov et al., 1971; Sultanov et al., 1972).

inclusion, using Debye-Hckel terms, is implemented in other engineering models (Zuo and Guo, 1991; Zuo et al., 1996). In the production process, the hydrocarbons and water may coexist with glycols, which are added for the purposes of inhibiting hydrate formation, depressing water freezing point, and dehydrating natural gas. Modelling glycol-hydrocarbon-water systems is theoretically difficult due to the self association and cross association in these systems. The Cubic-Plus-Association (CPA) EOS (Kontogeorgis et al., 1999), a recent EOS, developed by combination of the SRK EOS and the Wertheim association term, has succeeded in the description of this system. The most impressive aspect of the CPA EOS is its ability to simultaneously describe hydrocarbon-water mutual solubility using a single interaction parameter and to accurately predict multicomponent phase behaviour (Derawi et al., 2003).
Viscosity

composition both in the hydrocarbon phase and in the aqueous phase. Therefore, when the water content in the hydrocarbon phase is accurately correlated, the solubility in the acqueous phase will often be underestimated by orders of magnitude. A conventional engineering solution to the above problem is to use different sets of interaction parameters for the hydrocarbon phase and the aqueous phase, where the interaction parameter in the aqueous phase is usually assumed to be temperature dependent. Sreide and Whitsons modification (1992) of the PR EOS is such an example. Unconventional mixing rules can also be applied to the problem (Kabadi and Danner, 1985). Sreide and Whitsons model only accounts for the salt effect empirically by constricting the interaction parameters to be salinity dependent. More rigorous
75

The viscosity of formation water at reservoir conditions is virtually less than 1 cP, generally lower than the viscosity of oil. Within the normal reservoir temperature and pressure ranges, the viscosity of formation water increases with pressure and decreases with temperature. Furthermore, an increase in salinity will increase the viscosity, while the effect of dissolved gas is also believed to increase the viscosity. To account for this Whitson and Brule (2000) recommend a modified version of the empirical correlation by Kestin et al. (1981) for brine viscosity calculation.
Interfacial tension between hydrocarbon and formation water

interfacial tension (mN/m)

70 65 60 55 50 45 40 35 30 0 200 400

Sachs and Meyn (1995) 25 C Jennings and Newman (1971) 23.3 C 106.0 C 176.7 C

600

800

1,000

pressure (bar)
Fig. 18. Pressure and temperature dependency

of interfacial tension in the methane-water system (Jennings and Newman, 1971; Sachs and Meyn, 1995). Obvious disagreement can be observed between the data at 23.3C and 25C.

The interfacial tension of water/hydrocarbon systems varies from approximately 72 mN/m for water/brine/gas systems at surface conditions to 20 to 30 mN/m for water/brine/stock-tank-oil systems at reservoir conditions. Fig. 18 shows the experimental data on the interfacial tension between methane and water, which is largely representative of the case of reservoir gas and formation water. The interfacial tension decreases with pressure and temperature in the tested region. Due to the experimental difficulty, the published data are often in significant discrepancy as indicated by Fig. 18. The interfacial tension between liquid hydrocarbon and water decreases with temperature but slightly increases with pressure, while it generally increases in the presence of salts. The parachor method is not suitable to model interfacial tension between hydrocarbons and water.

60

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

On the other hand, Firoozabadi and Ramey (1988) suggested a graphical relation with limited success. The empirical correlation uses density difference and reduced temperature as correlation parameters. Finally, more recently, Zuo and Stenby (1998) applied the linear gradient theory to the modelling of interfacial tension between hydrocarbon and formation water and obtained fairly good results.
Mineral scaling

References
Aasberg-Petersen K. et al. (1991) Prediction of viscosities of hydrocarbon mixtures, Fluid Phase Equilibria, 70, 293-308. Acevedo S. et al. (1999) Thermo-optical studies of asphaltene solutions. Evidence for solvent-solute aggregate formation, Fuel, 78, 997-1003. Ali J.K. (1994) Prediction of parachors of petroleum cuts and pseudocomponents, Fluid Phase Equilibria, 95, 383-398. Altgelt K.H., BODUSZYNSKI M.M. (1994) Composition and analysis of heavy petroleum fractions, New York, Marcel Dekker. Anderko A. (1990) Equation of state methods for the modelling of phase equilibria, Fluid Phase Equilibria, 61, 145-225. Andersen S.I. (1994) Concentration effects in HPLC-SEC analysis of petroleum asphaltenes, Journal of Liquid Chromatography, 17, 4065-4079. Andersen S.I., Speight J.G. (2001) Petroleum resins: separation, character, and role in petroleum, Petroleum Science and Technology, 19, 1-34. Atkinson G., Mecik M. (1997) The chemistry of scale prediction, Journal of Petroleum Science and Engineering, 17, 113-121. Atkinson G. et al. (1991) The thermodynamics of scale prediction, in: Proceedings of the Society of Petroleum Engineers international symposium on oilfield chemistry, Anaheim (CA), 20-22 February, SPE 21021. Benedict M. et al. (1940) An empirical equation for thermodynamic properties of light hydrocarbons and their mixtures, methane, ethane, propane and n-butane, Journal of Chemical Physics, 8, 334-345. Bondi A. (1968) Physical properties of molecular crystals, liquids, and glasses, New York, John Wiley. Carey B.S. et al. (1978) On gradient theories of fluid interfacial stress and structure, Journal of Chemical Physics, 69, 5040-5049. Carey B.S. et al. (1980) Semiempirical theory of surface tension of binary systems, American Institute of Chemical Engineers Journal, 26, 705-711. Chang Y.B. et al. (1996) A compositional model for CO2 floods including CO2 solubility in water, in: Permian basin oil and gas recovery. Proceedings of the Society of Petroleum Engineers conference, Midland (TX), 27-29 March, SPE 35164. Chung T.H. et al. (1988) Generalized multiparameter correlation for nonpolar and polar fluid transport properties, Industrial and Engineering Chemistry Research, 27, 671-679. Coats K.H., Smart G.T. (1986) Application of a regressionbased EOS PVT program to laboratory data, Society of Petroleum Engineers Reservoir Engineering, 1, 277-299. Coutinho J.A.P., Stenby E.H. (1996) Predictive local composition models for solid/liquid equilibrium in n-alkane systems. Wilson equation for multicomponent systems, Industrial and Engineering Chemistry Research, 35, 918-925. Coutinho J.A.P. et al. (1995) Evaluation of activity coefficient models in prediction of alkane solid-liquid equilibria, Fluid Phase Equilibria, 103, 23-29. Coutinho J.A.P. et al. (1996) A local composition model for paraffinic solid solutions, Chemical Engineering Science, 51, 3273-3282.

Scale formation, the precipitation of inorganic minerals from brine, either in production facilities or in reservoir pores, is a serious problem in oil production. As an example, Fig. 19 illustrates the result of serious scale formation in a well tube. The most common types of scale found are CaCO3, BaSO4, SrSO4, CaSO4 and CaSO42H2O. Formation of these scales is usually caused by temperature and pressure change during production, which not only induces supersaturation of one or several dissolved minerals, but also changes the solubility of CO2 and H2S, and thus modifies the pH in the aqueous phase. The solubility of minerals such as carbonates is dependent on the pH and, therefore, scale may be formed after a change in the pH. Another common reason for scale formation is the mixing of incompatible water, which can happen during sea water injection. Thermodynamic modelling of scale formation mainly concerns whether, and how much, scale will form under certain conditions. Most of the modelling work uses a general model for electrolyte solutions (Atkinson et al., 1991; Yuan and Todd, 1991; Haarberg et al., 1992; Atkinson and Mecik, 1997; Kaasa, 1998). Scale formation modelling of sulphates is relatively simple while that of carbonates needs additional consideration of the pH and the phase distribution of CO2.

Fig. 19. Scale formation in a well tube

(Kaasa, 1998).

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

61

GEOSCIENCES

Culberson O.L., McKetta J.J. Jr. (1952) Phase equilibria in hydrocarbon-water systems III. The solubility of methane in water at pressures to 10,000 psia, American Institute of Mining, Metallurgical and Petroleum Engineers Transactions, 192, 223-226. Dandekar A.Y. et al. (2000) Compositional analysis of North Sea oils, Petroleum Science and Technology, 187, 975-988. Daridon J.L. et al. (2001) Solid-liquid-vapor phase boundary of a North Sea waxy crude. Measurement and modeling, Energy & Fuels, 15, 730-735. Derawi S.O. et al. (2003) Extension of the cubic-plusassociation equation of state to glycol-water crossassociating systems, Industrial and Engineering Chemistry Research, 42, 1470-1477. Edmister W.C. (1958) Applied hydrocarbon thermodynamics. Part 4: Compressibility factors and equations of state, Petroleum Refiner, 37, 173-179. Firoozabadi A., RAMEY H.J. JR. (1988) Surface tension of water-hydrocarbon systems at reservoir conditions, Journal of Canadian Petroleum Technology, 27. Firoozabadi A. et al. (1988) Surface tension of reservoir crudeoil/gas systems recognizing the asphalt in the heavy fraction, Society of Petroleum Engineers Reservoir Engineering, 3, 265-272. Fotland P., Bjorlikke O.P. (1989) Phase behaviour and interfacial tension as functions of pressure and temperature in a ternary alkane system, Journal of Physical Chemistry, 93, 6407-6413. Georis P. et al. (1998) Measurement of the soret coefficient in crude oil, in: Proceedings of the Society of Petroleum Engineers European petroleum conference, The Hague, 20-22 October, SPE 50573. GPA (Gas Processor Association) (1980) SI engineering data book, Tulsa (OK), Gas Processors Suppliers Association. Groenzin H., Mullins O.C. (2000) Molecular size and structure of asphaltenes from various sources, Energy & Fuels, 14, 677-684. Guerrero M.L., Davis H.T. (1980) Gradient theory of surface tension of water, Industrial and Engineering Chemistry. Fundamentals, 19, 309-311. Guo X.Q. et al. (2001) Equation of state analog correlations for the viscosity and thermal conductivity of hydrocarbons and reservoir fluids, Journal of Petroleum Science and Engineering, 30, 15-27. Haarberg T. et al. (1992) Scale formation in reservoir and production equipment during oil recovery. An equilibrium model, Society of Petroleum Engineers Production Engineering 7, 75-84. Haase R. (1969) Thermodynamics of irreversible processes, Reading (MA), Addison-Wesley. Haniff M.S., Pearce A.J. (1990) Measuring interfacial tension in a gas-condensate system with a laser-light-scattering technique, Society of Petroleum Engineers Reservoir Engineering, 5, 589-594. Heidemann R.A., Khalil A.M. (1980) The calculation of critical points, American Institute of Chemical Engineers Journal, 26, 769-778. Hirschfelder J.O. et al. (1954) Molecular theory of gases and liquids, New York, John Wiley; London, Chapman & Hall. Hoot W.F. et al. (1957) Solubility of water in hydrocarbons, Petroleum Refiner, 36, 255-256.

Hsu Y.D., Chen Y.P. (1998) Correlation of the mutual diffusion coefficients of binary liquid mixtures, Fluid Phase Equilibria, 152, 149-168. Jennings H.Y. Jr., Newman G.H. (1971) The effect of temperature and pressure on the interfacial tension of water against methane-normal decane mixtures, Society of Petroleum Engineers Journal, June, 171-175. Jhaveri B.S., Youngren G.K. (1988) Three-parameter modification of the Peng-Robinson equation of state to improve volumetric predictions, Society of Petroleum Engineers Reservoir Engineering, 3, 1033-1040. Kaasa B. (1998) Prediction of pH, mineral precipitation and multiphase equilibria during oil recovery (Ph.D. Thesis), NTNU (Norges Teknisk-Naturvitenskapelige Universitet), Trondheim. Kabadi V.N., Danner R.P. (1985) A modified Soave-RedlichKwong equation of state for water-hydrocarbon phase equilibria, Industrial and Engineering Chemistry Process Design and Development, 34, 537-541. Katz D.L., Firoozabadi A. (1978) Predicting phase behavior of condensate/crude oil systems using methane interaction coefficients, Journal of Petroleum Technology, 20, 16491655. Katz D.L. et al. (1959) Handbook of natural gas engineering, New York, McGraw-Hill. Kesler M.G., Lee B.I. (1976) Improve predictions of enthalpy of fractions, Hydrocarbon Processing, 55, 153-158. Kestin J. et al. (1981) Tables of the dynamic and kinematic viscosity of aqueous sodium chloride solutions in the temperature range 20-150 and the pressure range 0.1-35 Mpa, Journal of Physical and Chemical Reference Data, 10, 71-87. Kontogeorgis G.M. et al. (1999) Multicomponent phase equilibrium calculations for water-methanol-alkane mixtures, Fluid Phase Equilibria, v. 158-160, 201-209. Lee B.I., Kesler M.G. (1975) A generalized thermodynamics correlation based on three-parameter corresponding states, American Institute of Chemical Engineers Journal, 21, 510-527. Lee B.I., Kesler M.G. (1980) Improve vapor pressure prediction, Hydrocarbon Processing, 59, 163-167. Leibovici C.F. (1993) A consistent procedure for the estimation of properties associated to lumped systems, Fluid Phase Equilibria, 87, 189-197. Leibovici C.F. et al. (1996) A consistent procedure for pseudocomponent delumping, Fluid Phase Equilibria, 117, 225-232. Lindeloff N. et al. (1999) Phase-boundary calculations in systems involving more than two phases, with application to hydrocarbon mixtures, Industrial and Engineering Chemistry Research, 38, 1107-1113. Lohrenz J. et al. (1964) Calculating viscosities of reservoir fluids from their compositions, Journal of Petroleum Technology, October, 1171-1176. McCain W.D. Jr. (1990) The properties of petroleum fluids, Tulsa (OK), PennWell. McKetta J.J. Jr., Wehe A.H. (1962) Hydrocarbon-water and formation water correlations, in: T.C. Frick, R.W. Taylor (edited by) Petroleum production handbook, Dallas (TX), Society of Petroleum Engineers, v. II, 22-31. Martin J.J. (1979) Cubic equations of state-which?, Industrial and Engineering Chemistry. Fundamentals, 18, 81-97. Mathias P.M., Copeman T.W. (1983) Extension of the Peng-

62

ENCYCLOPAEDIA OF HYDROCARBONS

COMPOSITION AND PHYSICAL PROPERTIES OF HYDROCARBONS

Robinson equation of state to complex mixtures. Evaluation of the various forms of the local composition concept, Fluid Phase Equilibria, 13, 91-108. Michelsen M.L. (1980) Calculation of phase envelopes and critical points for multicomponent mixtures, Fluid Phase Equilibria, 4, 1-10. Michelsen M.L. (1982a) The isothermal flash problem. Part 1: Stability, Fluid Phase Equilibria, 9, 1-19. Michelsen M.L. (1982b) The isothermal flash problem. Part 2: Phase split calculation, Fluid Phase Equilibria, 9, 21-40. Michelsen M.L., Heidemann R.A. (1981) Calculation of critical points from cubic two-constant equation of state, American Institute of Chemical Engineers Journal, 27, 521-523. Michelsen M.L., Mollerup J.M. (1986) Partial derivatives of thermodynamic properties, American Institute of Chemical Engineers Journal, 32, 1389-1392. Michelsen M.L., Mollerup J.M. (2004) Thermodynamic models. Fundamentals & computational aspects, Holte (Denmark), Tie-line. Mollerup J.M., Michelsen M.L. (1992) Calculation of equilibrium properties, Fluid Phase Equilibria, 74, 1-15. Mollerup J., Rowlinson J.S. (1974) The prediction of densities of liquefied natural gas and lower molecular weight hydrocarbon mixtures, Chemical Engineering Science, 29, 1373-1381. Monnery W.D. et al. (1995) Viscosity. A critical review of practical predictive and correlative methods, The Canadian Journal of Chemical Engineering, 73, 3-40. Montel F., Gouel P.L. (1984) A new lumping scheme of analytical data for compositional studies, in: Proceedings of the Society of Petroleum Engineers annual technical conference and exhibition, Houston (TX), 16-19 September, SPE 13119. Nikitin E.D. et al. (1994) Critical constants of n-alkanes with from 17 to 24 carbon atoms, Journal of Chemical Thermodynamics, 26, 177-182 Olds R.H. et al. (1942) Phase equilibria in hydrocarbon systems. Composition of the dew-point gas in methanewater system, Industrial and Engineering Chemistry, 34, 1223-1227. Patel N.C., Teja A.S. (1982) A new cubic equation of state for fluids and fluids mixtures, Chemical Engineering Science, 37, 463-473. Pauly J. et al. (2000) Prediction of solid-fluid phase diagrams of light-gases-heavy paraffin systems up to 200 MPa using an equation of state-GE model, Fluid Phase Equilibria, 167, 145-159. Pedersen K.S. (1995) Prediction of cloud point temperatures and amount of wax precipitation, Society of Petroleum Engineers Production and Facilities, February, 46-49. Pedersen K.S. et al. (1989a) Characterization of gas condensate mixtures, in: Chorn L.G., Mansoori G.A. (edited by) C7+ fraction characterization, New York, Taylor & Francis, 137-152. Pedersen K.S. et al. (1989b) Properties of oils and natural gases, Houston (TX), Gulf. Pedersen K.S et al. (1991) Wax precipitation from North Sea crude oils. 4: Thermodynamic modelling, Energy & Fuels, 5, 924-932. Pedersen K.S. et al. (1992) PVT calculations on petroleum reservoir fluids using mesured and estimated compositional

data for the plus fraction, Industrial and Engineering Chemistry Research, 31, 1378-84. Pedersen K.S. et al. (2002) Cubic equations of state applied to HT/HP and highly aromatic fluids, in: Proceedings of the Society of Petroleum Engineers annual technical conference and exhibition, San Antonio (TX), 29 September2 October, SPE 77385. Peneloux A. et al. (1982) A consistent correction for RedlichKwong-Soave volumes, Fluid Phase Equilibria, 8, 7-23. Peng D.Y., Robinson D.B. (1976) A new two-constant equation of state, Industrial and Engineering Chemistry. Fundamentals, 15, 59-64. Peng D.Y., Robinson D.B. (1977) A rigorous method for predicting the critical properties of multicomponent systems from an equation of state, American Institute of Chemical Engineers Journal, 23, 137-144. Poling B.E. et al. (2000) The properties of gases and liquids, New York-London, McGraw-Hill. Quiones-Cisneros S.E . (2001) One parameter friction theory models for viscosity, Fluid Phase Equilibria, 178, 1-16. Quiones-Cisneros S.E. et al. (2000) The friction theory (f-theory) for viscosity modeling, Fluid Phase Equilibria, 169, 249-276. Quiones-Cisneros S.E. et al. (2004) PVT characterization and viscosity modeling and prediction of crude oils, Petroleum Science and Technology, 22, 1309-1325. Redlich O., Kwong J.N.S. (1949) On the thermodynamics of solutions. 5: An equation of state. Fugacities of gaseous solutions, Chemical Reviews, 44, 233-244. Riazi M.R., Daubert T.E. (1980) Simplify property predictions, Hydrocarbon Processing, 59, 115-116. Sachs W., Meyn V. (1995) Pressure and temperature dependance of the surface tension in the system natural gas/water. Principles of investigation and the first precise experimental data for pure methane/water at 25C up to 46.8 Mpa, Colloids and Surfaces, A. Physicochemical and Engineering Aspects, 94, 291-301. Schmidt G., Wenzel H. (1980) A modified van der Waals type equation of state, Chemical Engineering Science, 135, 1503-1512. Soave G.S. (1972) Equilibrium constants from a modified Redlich-Kwong equation of state, Chemical Engineering Science, 27, 1197-1203. Soave G.S. (1998) Estimation of critical constants of heavy hydrocarbons for their treatment by the Soave-Redlich-Kwong equation of state, Fluid Phase Equilibria, 143, 29-39. Speight J.G. (1999) The chemistry and technology of petroleum, New York, Marcel Dekker. Speight J.G. (2001) Handbook of petroleum analysis, New York, John Wiley. Starling K.E. (1973) Fluid thermodynamic properties for light petroleum systems, Houston (TX), Gulf. . Stryjek R., Vera J.H. (1986) PRSV An improved Peng-Robinson equation of state for pure compounds and mixtures, The Canadian Journal of Chemical Engineering, 64, 323-333. Sultanov R.G. et al. (1971) Moisture content of methane at high temperatures and pressures, Gazovaia Promyshlennost, 16, 6-8. Sultanov R.G. et al. (1972) Solubility of methane in water at high temperatures and pressures, Gazovaia Promyshlennost, 17, 6-7.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

63

GEOSCIENCES

Sreide I., Whitson C.H. (1992) Peng-Robinson predictions for hydrocarbons, CO2, N2, and H2S with pure water and NaCl brine, Fluid Phase Equilibria, 77, 217-240. Twu C.H. (1984) An internally consistent correlation for predicting the critical properties and molecular weights of petroleum and coal tar liquids, Fluid Phase Equilibria, 16, 137-150. Twu C.H. et al. (1995a) A new generalized alpha function for a cubic equation of state. Part 1: Peng-Robinson equation, Fluid Phase Equilibria, 105, 49-59. Twu C.H. et al. (1995b) A new generalized alpha function for a cubic equation of state. Part 2: Redlich-Kwong equation, Fluid Phase Equilibria, 105, 61-69. Wei Y.S., Sadus R. (2000) Equations of state for the calculation of fluid-phase equilibria,American Institute of Chemical Engineers Journal, 46, 169-196. Weinaug C.F., Katz D.L. (1943) Surface tensions of methanepropane mixtures, Industrial and Engineering Chemistry, 35, 239-246. Wesselingh J.A., Bollen A.M. (1997) Multicomponent diffusivities from the free volume theory, Transactions. Institution of Chemical Engineers, 75, 590-602. Whitson C.H. (1983) Characterizing hydrocarbon plus fractions, Society of Petroleum Engineers Journal, August, 683-694. Whitson C.H., Brule M.R. (2000) Phase behaviour, Richardson (TX), Society of Petroleum Engineers. Won K.W. (1986) Thermodynamics for solid-solution-liquid-

vapor equilibrium. Wax phase formation from heavy hydrocarbon mixtures, Fluid Phase Equilibria, 30, 265-279. Yuan M.D., Todd A.C. (1991) Prediction of sulphate scaling tendency in oilfield operations, Society of Petroleum Engineers. Production Engineering, 6, 63-72. Zuo Y.X., Guo T.M. (1991) Extension of the Patel-Teja equation of state to the prediction of the solubility of natural gas in formation water, Chemical Engineering Science, 46, 3251-3258. Zuo Y.X., Stenby E.H. (1996a) A linear gradient theory model for calculating interfacial tensions of mixtures, Journal of Colloid and Interface Science, 182, 126-132. Zuo Y.X., Stenby E.H. (1996b) Calculation of surface tensions of polar mixtures with a simplified gradient theory model, Journal of Chemical Engineering of Japan, 29, 159-165. Zuo Y.X., Stenby E.H. (1998) Calculation of interfacial tensions of hydrocarbon-water systems under reservoir conditions, In Situ, 222, 157-180. Zuo Y.X. et al. (1996) Simulation of the high-pressure phase equilibria of hydrocarbon-water/brine systems Journal of Petroleum Science and Engineering, 15, 201-220.

Erling Halfdan Stenby Wei Yan


Department of Chemical Engineering Technical University of Denmark Lyngby, Denmark

64

ENCYCLOPAEDIA OF HYDROCARBONS

Anda mungkin juga menyukai