Sei sulla pagina 1di 38

5.

Treatment plants for oil production

5.3.1 Introduction
This chapter will cover all surface facilities (referred to as such to distinguish them from production wells) used to gather reservoir fluids, separate crude oil from other phases, and ensure sufficient treatment of the latter to render it transportable. According to this scheme, the surface facilities used for separation and treatment, known as treatment plants, are those comprised between the production wellhead (excluded) and the crude oil storage facilities (included). Finally, the surface plants used for secondary recovery projects will be considered. This discussion will also cover the equipment needed to inject water and gas, and their pumping and compression systems, up to the head of the injection well (excluded). For a description of chemical and physical properties of the reservoir fluid produced and carried to the surface by production wells, see Chapter 4.2. Except in some unusual cases, the fluid produced by an oil field reaches the production well head as a combined result of two or three phases. The main liquid phase is the crude oil itself, saturated in dissolved light hydrocarbons under wellhead conditions (delivery pressure and temperature). The second phase of the well head fluid is associated gas, which is in thermodynamic equilibrium with the liquid phase described above. Consequently, the gas phase associated with the crude oil is saturated in heavy hydrocarbons (hydrocarbon dew point) at the delivery pressure and temperature mentioned above. During the production of a field, the wellhead fluid also frequently consists of a third phase: free water. This phase, as well, is in equilibrium with the two phases described above, and the associated gas is therefore saturated in water (water dew point). Depending on its origin, the free water phase may consist of formation water (i.e. coming directly from the productive

formation where it is already present as a free phase), or supersaturation water. The main property of formation water is its high salt content, in some cases up to 340 g/l of TDS (Total Dissolved Solids). When salinity is very high, this is mainly due to chlorides, though carbonates, bicarbonates and sulphates may also be present in quantities close to their saturation levels. Given these properties, formation water represents one of the main contaminants of crude oil production, since its presence as a dispersed phase within the oil gives the latter a salinity that causes problems in field treatment and later in refining processes. In other words, salinity must be reduced with suitable treatments (see Section 5.3.4) to ensure that the crude oil can be transported and commercialized. When an oil field comes into production, formation water is not present, except in unusual cases. In later years, and especially during the advanced stage of reservoir exploitation, the percentage of water in terms of volume produced by the wells (water cut) may become extremely high so high that it is not unusual to have production wells with a water cut above 50%. It is easy to perceive that limited values for water cut correspond to a dispersed water phase in a dispersing crude oil phase. On the other hand, when values are high, formation water becomes the dispersing phase and oil the dispersed phase. The properties of these emulsions are discussed in Section 5.3.4. The free water phase may also be devoid of salinity. This occurs when, instead of coming directly from the productive formation, free water forms due to the oversaturation of either the liquid hydrocarbon phase or the associated gas phase, or both. Oversaturation is caused by the cooling that the fluid delivered by the reservoir undergoes along the well string. The solubility of water, both in the oil phase

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

643

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

and in the associated gas phase, drops markedly as temperature decreases (Fig. 1). The water separated from the two hydrocarbon phases (liquid and gaseous) is thus devoid of salinity. Both formation water and oversaturation water contain suspended solids. Where present, the free water phase tends to stratify, since it is heavier than the crude oil phase; as a result, most of the solid impurities dragged from the reservoir or produced due to corrosion in the pipelines tend to accumulate in the water phase. When referring to a crude oil, it is common practice to consider residual water content and sediments as a single parameter: BS&W (Bottom Sediments and Water). The two hydrocarbon phases (liquid and vapour) enter the production facilities as a mixture. Frequently, these hydrocarbons are found in a single phase within the reservoir from which they are produced; in other words, the pressure of the productive formation is either equal to or higher than the bubble point of the mixture concerned at formation temperature. This circumstance occurs very frequently in oil reservoirs under original conditions. Often, during the depletion of the reservoir, pressure tends to drop rapidly; this leads to the creation of two separate phases within the productive formation itself. The facts outlined above are intended to highlight the common origin of the two phases, and thus their shared chemical nature. It is common practice to study and characterize an oil field on the basis of the chemical composition of the reservoir fluid; this is referred to as recombination, since it is obtained by recombining the two phases, produced and sampled separately, according to their original proportions. The recombination of a petroleum reservoir, and of an oilfield in particular, is a mixture of hydrocarbons with a variable light and heavy hydrocarbon content. In light hydrocarbons methane predominates, while ethane, propane and butane are also present, though in more modest and decreasing quantities. Given their volatility, these components are predominantly present in the vapour phase as they enter production facilities. Heavy hydrocarbons present enormous variability. Hydrocarbons of varying molecular weight exist, from that of pentanes (with a value of 72.17) up to values that are an order of magnitude higher. As already mentioned in Chapter 1.1, oil reservoirs contain almost the entire range of saturated and unsaturated hydrocarbons, paraffins, naphthenes and aromatics (olefins, however, are absent). As far as paraffins, or aliphatics, are concerned, both linear and branched chain types are present. The physical properties (the density and viscosity of the various fractions of a crude) depend on the variable presence of the components mentioned. Therefore, it is common

0.5

water solubility (lb water/100 lb of hydrocarbon)

0.1

0.05

0.01
heptene-1 hexadiene-1.5 butadiene-1.3 benzene butene-2 butene-1 i-butene styrene i-pentane i-butane n-heptane propane n-hexane n-butane n-pentane n-octane cyclohexane lube-oil

0.005

0.001 5 15 25

temperature (C)

35

45

55

65

75

Fig. 1. Solubility of water in liquid hydrocarbons

(GPSA, Gas Processors Suppliers Association).

practice to characterize crudes according to their predominant paraffin, naphthene or aromatic content. This characterization, like specific density (or API gravity), is of considerable practical interest in the refining industry and for the commercialization of crude oil. It is of minor relevance, however, in the evaluations needed to determine the type of gathering and treatment procedures used for crude oil. In the past, numerous attempts were made to correlate the chemical-physical properties of crude oils and their cuts in a simple way. The most interesting, though rarely used in production activities, is the UOP (Universal Oil Product) characterization factor. This factor, indicated with the symbol K, relates the mean boiling point at atmospheric pressure of a petroleum cut to its relative density, K[(460tf)d ](1/3), where tf expresses temperature in degrees Fahrenheit, and d is the specific density of crude cut relevant to water at 60C. For many crudes, this factor remains relatively constant for most cuts, with the exception of both very light and especially very heavy cuts. The UOP factor can also be directly correlated with the degree of hydrocarbon saturation of which the crude is composed. The lower the value of K, the lower the hydrogen/carbon ratio, and vice

644

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

versa. Consequently, crudes with a high aromatic content have a relatively low value for K (10-11), whereas those with a paraffinic base have considerably higher values (12-12.5). A reservoir fluid, and the crude oil deriving from it, do not only contain hydrocarbons. These fluids also contain other elements, such as nitrogen, mainly in the form of N2, but also nitrogen compounds, sulphur in the form of various sulphur compounds such as hydrogen sulphide (H2S), mercaptans, ranging from the lighter methyl and ethyl mercaptans present in light cuts (benzene) to heavier types, as well as ternary aromatic sulphur compounds. Oxygen is also present, mainly as carbon dioxide or water. In the reservoir fluid dissolved oxygen, O2 is not present. After storage at atmospheric pressure, the stabilized crude may contain dissolved oxygen, since it has come into contact with air during the storage process. The presence of oxygen accentuates chemical reactions, with the partial oxidation of high-boiling and highly unsaturated compounds. Finally, a crude oil may contain helium (He), elementary mercury or its compounds, such as mercury sulphide (HgS) and others; other heavy metals such as chrome, vanadium and their salts may also be present in the reservoir fluid. Sulphur compounds have a significant impact on oil refining treatment, and an extremely negative influence on the commercialization of the crude. In simple terms: total sulphur contents above 1.5% in weight in a crude oil considerably penalize the sales price. By contrast, crudes with a low sulphur content have a higher price. In the field production and treatment of oil, it is not possible to alter the total sulphur content significantly; however, it is of fundamental importance to reduce both the H2S and mercaptan content. The first of these compounds is a lethal gas even at very low concentrations; it is extremely dangerous since, being heavier than air, it tends to stratify in the environment. The reduction of mercaptans is necessary only for those which are volatile and more aggressive (methyl and ethyl mercaptans); this reduction must be carried out during the treatment phases if their content exceeds allowed values.

5.3.2 Characterization of oils


The development and exploitation phase of an oil field requires an extensive knowledge of the chemical physical properties of the crude produced, and those of the fluid contained within the reservoir. Therefore, it is necessary to examine the chemical composition and thermodynamic behaviour of the recombined reservoir

fluid. This information is obtained with PVT (Pressure, Volume, Temperature) analyses carried out in the laboratory, and presented in a document known as a PVT report. This information is essential to draw up an optimal development plan for the reservoir, including the location and number of production wells and their production capacity. On the basis of these facts, reservoir studies (see Chapters 4.5. and 4.6) are used to define a production profile, representing the basis for the entire development of the field, and in particular for the design of treatment facilities. Although mainly relevant to reservoir studies and development techniques, PVT analysis is also the principal source of the basic data needed to design surface facilities. The fluid samples used for PVT analysis (see Chapter 4.2) are frequently taken during production tests using a test separator (see Section 5.3.3). Separator tests make it possible to sample and analyse separately the associated gas produced by the well under examination, and the liquid phase (or phases: hydrocarbons and water), to measure their relative flow rates and thus obtain the composition of the recombination. The data obtained during these tests are then processed in the laboratory to obtain all the data needed to identify the behaviour of the reservoir fluid under reservoir conditions during its productive life. Thus, one obtains the compositions, the relative flow rates of the gas under the various separation conditions, the GOR (Gas-Oil Ratio), and the properties of the oil under standard conditions, in other words, stabilized at atmospheric pressure with several flash stages. The latter make it possible to evaluate the density and, above all, the viscosity of the stabilized oil experimentally. While a detailed knowledge of the reservoir fluid chemical composition is sufficient to determine the compositions and flow rates of the individual separated phases, and, more generally, to reconstruct a complete balance, experimental data are however, needed to determine transport properties, in particular viscosity. The data are usually defined using PVT analysis at two different temperatures in order to construct a complete diagram of the viscosity patterns of the stabilized oil. Flash separation enables one to reconstruct the compositions of the saturated liquid under various conditions using thermodynamic equilibrium calculations, and to easily obtain viscosity values for the liquid and gas. Evaluating the viscosity of a crude is fundamentally important to correctly define all the phases of gathering, separation, storage and transportation of the same. This aspect often requires in-depth investigation and, in some cases, a complete rheological study.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

645

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

As far as chemical composition is concerned, it is well-known that a crude oil may contain high concentrations of paraffins even in intermediate cuts. When these high-boiling paraffins are of linear chain type, their chemical structure increases the force of attraction between the molecules, which can be seen very simply in their physical state: they are solid at ambient temperature. If they are sufficiently diluted in other hydrocarbons, branched paraffins, naphthenes and light aromatics remain in solution even at relatively low temperatures. On the other hand, if they are present in high concentrations, they cause the formation of solid crystals suspended in the liquid, and at lower temperatures the separation and deposition of the paraffins themselves; these phenomena are more accentuated the lower the temperature. Above the point at which microcrystals of paraffins form (cloud point), the crude behaves like a Newtonian fluid, whereas below this point it becomes a pseudoplastic fluid, whose rheological properties are difficult to identify. The study of fluid motion under these conditions is extremely complex; the viscosity of the crude under examination is no longer a constant governed only by temperature and by the composition of the fluid, but varies depending on flow conditions in the pipeline. It has a maximum value at zero velocity (yield value), and tends to decrease as the flow velocity increases (apparent viscosity). Unfortunately, during the exploitation of a reservoir, this situation is fairly common, and therefore must be taken into consideration especially when transportation of the stabilized crude is concerned. It is evident that, whenever economically feasible, non-Newtonian flow in the crude should be avoided by keeping the temperature above the cloud point. This situation mainly concerns treatment and storage; the equipment involved, being concentrated inside a

restricted area, can be heated and thermally insulated, thus avoiding problems caused by the deposition of solid paraffins. In the case of gathering networks used to channel oil production from individual wells to the treatment plant, this is not always possible. In these circumstances, the support of a complete rheological study and a correct assessment of all transportation parameters are needed. In particular, the pressure required to move the gel piston in the section of pipeline to be started up must be calculated with accuracy. This calculation is extremely simple when there is a known value of the shear rate for the temperature at which the pipeline concerned must be started up. To obtain more detailed data on the characterization of the oil, a so-called crude assay is needed. This report represents a full and exhaustive analysis of the stabilized crude, and is aimed at its commercialization. As far as crude oil production is concerned, it is important to know a series of properties and specifications required for its commercialization, referring to the crude sample in its entirety. Table 1 lists some of the most important, with the relevant reference standard. Below, the properties and specifications required for the commercialization of crude oil will be described.
Density and API gravity

Density and API gravity are properties that cannot be significantly altered with treatment. In the past, there was a close correlation between the API gravity value and the sales price; therefore, it was sufficient to increase the API degrees even by a few decimals to obtain an increase in price. The treatments which increase the light ends recovery lead to a density reduction and a corresponding increase of crude API gravity; thus, these treatments are advantageous

Table 1 Characteristics Density and API gravity Reid vapour pressure H2S content Mercaptan (methyl and ethyl) content Total sulphur content Water content and salinity Paraffin content (wax) Upper and lower cloud point Pour point Viscosity Total acid number Heavy metal content Unit of measurements API psia ppm by weight ppm by weight % by weight % by volume and ptb % by weight C C cP mg of KOH/g ppm by weight Reference standard

ASTM D 323 ASTM D 3227 ASTM D 3227 ASTM D 4294 ASTM D 4006, IP 265 BP 237

ASTM D 445 ASTM D 664 IP 288

646

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

although less profitable than in the past, since the crude most in demand, with few exceptions, is now medium rather than light.
Reid Vapour Pressure (RVP)

The crudes vapour pressure is a measure of its degree of stabilization. The lower the pressure, the more stabilized the product; this specification is thus essential for the definition of treatment processes for the crude. It is important to remember that even products like gasoline and LPG have very strict constraints on vapour pressure. Since they are stored in atmospheric tanks, both crude oil and the gasoline obtained from it must have a vapour pressure below the boiling point at storage temperature. The greater the difference between atmospheric pressure and the vapour pressure at storage temperature, the smaller the losses due to evaporation, which are a significant cause of environmental pollution. Where no systems are applied to collect and recover displaced vapours, a higher degree of stabilization considerably reduces the pollution of the atmosphere due to non-combusted hydrocarbons. For gasolines, vapour pressure specifications are clearly defined, and reflect particular environmental conditions (specifically, two values: for summer and winter), whereas crude oil regulations are fairly flexible. According to the Reid method, vapour pressure is a standardized measurement of the pressure created inside a cylinder of standard size, immersed in a thermostatic bath at 100F (37.8C) after being partially filled with the liquid sample to be tested. However, this method involves a systematic error in measurement, since the vapour pressure of gasoline depends largely on its butane content. At 100F, normalbutane and isobutane have vapour pressures of about 56 psia and 60 psia respectively; as a result, they contribute substantially to vapour pressure. This means that the loss of light hydrocarbons through evaporation, which occurs during the measurement and sampling process itself, significantly reduces the value for vapour pressure of the liquid examined. For gasolines, this deviation may be close to 20%; in other words, a Reid vapour pressure of 10 psia corresponds to a True Vapour Pressure (TVP) of about 12 psia. For crude oil, the situation proves even worse; the contribution to the mixtures vapour pressure is provided by dry gases (methane and ethane), which, even if present in very modest concentrations, may supply 50% of the total pressure. As pure components, these have extremely high vapour pressures (respectively 5,000 and 800 psia at 100F). If 0.2% in volume of the sample is lost through evaporation during sampling, since this loss is principally due to methane, ethane and LPG, the vapour pressure of the crude

presents a deviation so large that it even influences the order of magnitude. For example, values of 3 psia for Reid vapour pressure, relevant to 17 psia for true vapour pressure, are common when the crude is stabilized using the multiflash system (see Section 5.3.3). From a commercial and regulatory point of view, the RVP should still be considered the only standard measurement used to evaluate the transportability of a crude oil, even though the considerations outlined above indicate that it may present notable deviations. From both a theoretical and practical point of view, it would be more sensible to refer to the true vapour pressure. Using modern computer calculation techniques, if the composition of the light hydrocarbons in a crude sample is known (e.g. up to the pentanes), the true vapour pressure can be calculated with very modest deviation and the expected operating margin and evaporation losses can be established more precisely.
Hydrogen sulphide (H2S) content

Although hydrogen sulphide is not the main contaminant of crude oils, it is certainly the most dangerous. As a result, both regulations and the plant types used to guarantee acceptable values must be evaluated with great care, standardizing the specifications for the maximum admissible H2S content as much as possible. The maximum value allowed may vary depending on the type of stabilization chosen. For example, if for commercial reasons it is decided to stabilize the crude in an extreme way, an H2S content with values above those allowed is less dangerous. If the process of sweetening the crude is carried out with stabilization by fractionation (see Section 5.3.4) rather than with more or less heated flash separations (see Section 5.3.3), a good stabilization always leads to the reduction of H2S to negligible values. As far as standard specifications are concerned, only the GOST (Gas-Oil Transport Specifications applied in Eastern European and ex-Soviet countries) specify a single value for transportation and commercialization. This value is 20 ppm wt. In western countries, and particularly in Europe, there is no universally accepted standard, but the companies operating in this sector have frequently self-imposed values of 50 ppm in weight, and sometimes higher (70 ppm). In recent years, however, a greater sensitivity towards the problems linked to pollution and safety have led to these values being lowered considerably.
Mercaptan content (methyl and ethyl)

The mercaptan content is limited for various reasons. The most obvious is that the presence of a few parts per million of these compounds is sufficient to

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

647

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

confer an extremely strong and unpleasant odour on the vapours given off by stock tanks. Processing oil with a high mercaptan content should therefore be avoided, since refineries are often located near densely populated areas. Mercaptans are no more dangerous to human health than the corresponding paraffin, but they are fairly dangerous for atmospheric storage. Since they are relatively aggressive, they attack the metal wall of the tank at the vapour-liquid interface, causing corrosion and the production of an iron sulphide patina. Therefore, the problem caused by this aggression is not due to the corrosion as such, which is fairly modest, but to the fact that iron sulphides become pyrophoric when in contact with oxygen. Since H2S has the same property, the presence of both compounds has a cumulative effect. In this same context, the GOST are extremely restrictive, and set out a specific concentration standard of 60 ppm by weight (as methyl and ethyl mercaptan). In the western world, this restriction has always been ignored, at least in the past. In some cases, however, it is indirectly applied, since the total organic acidity content is extremely important in the commercialization of crude oil. H2S and mercaptans react with the titration substance (a KOH solution) and may thus raise the total acidity considerably, significantly penalizing the price of the crude oil. Demercaptanization is even more important when, as in many recent large projects, it is decided to transport the oil produced by various fields in a cooperative manner, using a single shared infrastructure. If crudes with a high mercaptan content were sent into the oil pipeline without prior processing, or with overly lax specifications, they would inevitably contaminate the entire production transported by the oil pipeline in question, leading to a drop in price.
Total sulphur content

separators). Only in unusual cases, with very dense and viscous oils, is specific treatment required. The salt content of the crude oil depends exclusively on the presence of formation water. For a long time, both the water and salt content were considered a secondary problem in the production of crude oil, to be resolved indirectly using equipment designed for other purposes, such as stock tanks. This attitude was the result of two factors. Firstly, the oil produced is mainly transported by sea with oil tankers; since sea water with a salinity of 35 g/l is used to ballast them, the crude transported is in any case contaminated with salt water. The second more serious reason is that the refining, which the crude later undergoes, includes an extremely severe desalting of feed to be treated. This occurs because during processing extremely high temperatures are reached at which chlorides dissociate, causing the formation of hydrochloric acid. To limit this effect, the salinity of oil in the refinery is reduced to values below 3 ptb (pounds per thousand barrels), equivalent to about 8.6 g/m3. During production activities, this value is uncommon: the most frequently used is 20 ptb, corresponding to 57.5 g/m3. If the two specifications (water content and salinity) described above are combined, it is possible to determine the maximum acceptable salinity of residual water in the crude, equal to 11.5 g/l. To conclude, in many cases, since this standard is not universally accepted, one may come across far more lax specifications, for example 60 ptb or even 200 g/m3.
Cloud point, pour point, paraffin content

Sulphur content is extremely important for the commercialization of the product, but is of minor relevance in field processing. The treatements outlined in the previous paragraph do not usually influence the total sulphur content, firstly since mercaptans contain a negligible proportion of total sulphur, and secondly because the treatment used to extract mercaptans, though available, is not used in field production activities.
Water content and salinity

Water is present in crude oil production, and must be removed for commercialization. The allowed residual water content for a treated crude is universally set at 0.5% in volume. Water is easily removed during the process of separating oil and gas (three-phase

The cloud point is the temperature at which microcrystals of paraffin begin to form in the crude oil production. If the temperature drops further, the crude loses its original properties and turns into an increasingly dense gel. It becomes extremely viscous, and can no longer flow through the pipeline. The temperature corresponding to this limit is known as the pour point. It is worth recalling that PVT analysis often fails to identify these values, which instead are shown generically by the crude assay. Given its complexity, this study is not always available during the initial phase of a field development project; therefore, it is not improbable that development may begin without these evaluation. The definition of the parameters, as they are quantified in the crude assay, only indicates the existence of a problem linked to the presence of paraffins, and provides the most important data to evaluate its extent. In order to resolve the problem, the rheological study mentioned above is also necessary.
Viscosity

The viscosity values found during the crude oil characterization study are only approximate, and

648

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

usually very pessimistic. When studying the treatment of the oil, direct measurements on the fluid produced and separated at the wellhead are therefore needed. The data obtained with PVT analysis are generally more accurate.
Total acidity

The origin of total acidity has already been mentioned above. In general, it is important to remember that crudes with high acidity have a lower sales price, thus making it advisable to reduce this value. In the discussion of H2S and light mercaptans, and therefore volatile acidity, it became clear that it is possible and advisable to alter this value. By contrast, it is not feasible to alter acidity due to high-boiling compounds (naphthenic acids). An order of magnitude commonly accepted without penalties for total acidity is 0.25 mg of KOH/g.
Heavy metal content

The heavy metal content has no impact on field treatments. A desalting process that involves washing with water may reduce their presence in the form of salts, although this is not a significant objective in field treatment. In some very rare cases, when it is convenient to use the stabilized crude as a fuel to generate electrical energy, the heavy metal content is reduced only in the amounts used for this purpose, as specified by the suppliers of engines and turbines.

to the preceding reservoir study, and the subsequent refining process. Equilibria and phase curves are commonly used in thermodynamic reservoir study (see Chapters 1.1 and 4.2). Although conceptually identical to those used for treatment plants, they are characterized by the extreme pressure conditions, under which normal equations of state are difficult to employ. For this reason, experimental studies are necessary. This is not true under the most common handling and treatment conditions for crude oil. In this sector of the petroleum industry, as will become clear from the discussion below, most of the processes involved are not based on chemical reactions, and do not involve the use of catalysts, frequently used in refining. Consequently, both the design and operation of treatment plants can be easily schematized and broken down into a series of few operational units, linked basically to transformations of a physical status. As such, it is much easier than in other sectors to develop extremely reliable and realistic material and energy balances on the basis of the most common equations of state, without the need for empirical data. The reliability of the above-mentioned methods, and the refinement of simulation programmes, have shortened the time needed to design crude oil treatment processes, and optimized the performance of even very complex plants.
Oil-gas separation

5.3.3 Separations
Before describing the various separation methods, it is worth noting some of the peculiarities characteristic of oil treatment in the field and the use of thermodynamic equilibria involved, both with respect
100

The first treatment process undergone by the crude oil is based on a simple physical separation of the main phases. The fluid produced by the reservoir through a system of gathering networks is transported from individual wells to one or more treatment centres. Transport occurs under oil-gas phase flow conditions; as a result, the separation system, or more

bubble NLvL(rL/gs)1/4
10

bubble slug

vL, liquid velocity vG, gas velocity rL, liquid density g, acceleration of gravity s, superficial tension NL, liquid velocity number NG, gas velocity number

slug wave
1.0

wave stratified annular


0.1 0.1

annular stratified

1.0

10

100

1,000

NGvG(rL/gs)1/4
Fig. 2. Two-phase horizontal flow regimes

(Katz et al., 1959).

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

649

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

accurately its first stage, is heavily influenced by the flow conditions in the gathering system. Fig. 2 shows the different types of horizontal two-phase flow patterns. If the flow is stratified, the two phases are already uniformly separated when they enter the separation system; therefore, all that is required is a small vessel through which the two different streams are delivered into their respective collectors. However, under the flow conditions described as slug flow, the situation is completely different. The vessel that makes separation possible must first uniform the flow entering the plant, and therefore be considerably larger; its size basically depends on fluid mechanical conditions up-stream the vessel. It should be noted that, while the optimal condition of stratified flow hardly ever occurs, slug flow is very common. In practice, it is almost impossible to obtain a uniform flow entering the gasoil separation unit. This is due to design choices made in sizing the gathering system, the typology of the fluid transported (gas-oil ratio, density and viscosity of the liquid phase, etc.) and, finally, to the altymetry of the pipe route where two-phase transportation occurs (a sequence of small dips and rises). To conclude, in addition to the flow rate, the gas-oil ratio and the properties of the phases, the sizing of the separation system must also take into account the type of incoming flow and its basic parameters. A second factor connecting separation with the preceding production system is represented by test equipment, essential to establish the optimal exploitation conditions for the field during its entire productive life. This knowledge is obtained with periodic tests carried out on individual producing wells. The simplest and most common way of carrying out these production tests is to have two manifolds at the entrance to the plant: one for production, into which all the wells flow to feed the production line, and one for testing (test manifold), to which individual wells are connected in turn. This solution enables the use of a single test separator, working in parallel with the production separators; however, as a result, the gathering network for this use must be constructed with individual flowlines, well by well, until these reach the test separator. When exploiting very large reservoirs, it is possible to collect several wells from a single area and channel them into the centralized treatment centre through a single flowline. In this case, the satellite centre (or centres, if there are more than one) must in turn have a test manifold and test separator; alternatively, alongside the production line, a single flowline for testing can be used, linking the satellite in question to the main oil station. With these preliminary considerations in mind, a complete oil-gas separation system can be analysed in

greater detail. The purpose of the latter is to separate the two phases, producing a gas stream and a stabilized oil, in other words, two intermediate products, since both will undergo further transformation before they can be considered final products. As far as the main product is concerned (i.e. the crude oil), it has already been noted that treatment mainly involves rendering it stable under storage and transportation conditions. Since the latter two operations are usually carried out at ambient temperature and atmospheric pressure, stabilizing the oil involves separating it from associated gas to guarantee a vapour pressure lower than or equal to atmospheric pressure. This is usually done with a multiflash system, involving a multi-stage separation at decreasing operating pressures, from the pressure on arrival at the treatment centre to the final stage at atmospheric pressure. This is the simplest treatment process, and remains fundamental in the separation and stabilization of oil. In the past, its simplicity was increased by the fact that only a small portion of the gas streams produced at decreasing pressures was used for field utilities, such as producing the electric energy needed to operate the treatment plant, for oil transportation and any heating required. Under these conditions, the treatment system was basically limited to a series of separation stages, usually three (maximum four), and the flares used to dispose of the associated gas produced. Since this type of plant has not been completely abandoned, it is important to stress that a simplified treatment has a detrimental effect on environmental pollution and energy conservation. A further effect should also be considered: the efficiency implied in this type of choice. If all the gas produced by separation were collected, compressed and sent for treatment and subsequent use, the efficiency of separation would be less important: anything not recovered as crude oil would be recovered as gas. On the other hand, if the gas is burned (or, as is more frequent today, reinjected into the reservoir), a modest, though not negligible, amount of light liquid products is burned or reinjected into the reservoir. It is clear that the main purpose of a multiflash separation system, or other type of stabilization system, is to maximize the recovery of light hydrocarbons. This increases the production of crude oil, and increases, though only slightly, the products API gravity. Generally speaking, in order to increase the recovery of liquids, the number of separation stages is increased. Consider an oil stabilization process with only two stages: one under pressure and one at atmospheric pressure. This solution is adopted for production tests with fluid sampling for PVT analysis.

650

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

Production tests are carried out at sufficient pressure to obtain the production parameters of the wells. In order to translate well production into stock tank conditions (i.e. atmospheric pressure), the liquid phase collected by the test separator is operated at atmospheric pressure. The resulting product is stabilized oil under stock tank conditions. On the other hand, the two gas streams, at the separator and at the atmospheric stage, are extremely rich in light compounds, which, if a different method were employed, may remain dissolved in the oil, thus increasing its quantity and quality. If the delivery pressure of the wells, and consequently that on arrival at the gathering centre, is sufficiently high, it is preferable to pass from two to three, or even four stages (three plus an atmospheric stage). Since recovery or losses, depending on the point of view, are linked to associated gas in the first stage, and to dissolved gas in the other stages, it is preferable to use several stages in order to free the latter. Considering that, all other parameters being equal, dissolved gas is directly linked to pressure, the higher the pressure during the first stage of separation, the greater the dissolved GOR in the liquid separated, and thus the higher the recovery of liquids if several stages of equilibrium are used at intermediate pressures between the original pressure and final atmospheric pressure. It is important to stress that the recovery of liquids is not alone in benefiting from the subdivision into several separation stages; this is also true of any separated gas that anyhow must be recompressed from the working pressure at which it was produced to the required final pressure through a series of compression stages. Another important parameter in the stabilization of oil by separation is the working temperature of the atmospheric stage. Under normal conditions, a value of 40C, giving a true vapour pressure slightly below atmospheric pressure, reduces losses from evaporation to a minimum (see Section 5.3.6). Moreover, in storage tanks the Reid vapour pressure is very low, in line with common transportation regulations. It can be observed that vapour pressure is mainly due to a few components, ranging from methane to the butanes. If, for the sake of simplicity, Raoult and Daltons laws (valid for perfect gases) are applied to the hydrocarbon mixtures under consideration, and one considers the vapour pressures of their pure components, thus follows PP i Xi; where P is the vapour pressure of the mixture, P i Xi is the partial pressure of the i-th component, Xi is its molar fraction, and P i the pure component vapour pressure. Since the interaction between the components is modest, using the laws cited above does not result in significant deviations (more conservative values are generally

obtained than those resulting from bubble point calculations using the adequate equations of state). Considering the case of two-stage separation, it is evident that the amount of methane and ethane present in the stabilized oil is not negligible. The contribution made by these components to vapour pressure is even more significant. Assuming a loss in the order of 0.2% in volume during storage, this loss, translated into the percentage of moles, becomes far more significant. In this case, the oils final dry gas content (i.e. methane and ethane) is evidently subject to considerable variation. Similarly, the final vapour pressure of the crude undergoes a significant variation. Basically, the stabilized and stored crude may reach an acceptable vapour pressure during storage itself, which thus improperly becomes the final stage of oil stabilization prior to transportation and commercialization. This mechanism explains why in many cases, especially for relatively light oils (API gravity36) with low viscosity, it is unnecessary to control the final separation temperature to obtain adequate stabilization, since this occurs during the period in storage. In practice, separation thus occurs at the same temperature as that of arrival at the gathering centre and treatment. This practice is not always advisable, in part since the temperatures of the fluids delivered by the wells may be very high, and remain so in the gathering system. This occurs relatively frequently in the development of onshore fields. If the productive horizon is fairly deep, for example 4,000 m, as is often the case for recently discovered oil fields, the temperature may be over 120C. In unusual cases it may be as high as 150C at depths only slightly above 4,000 m. Under these conditions, even taking into consideration the cooling along the well string (not significant) and the more significant cooling along the buried pipeline, the fluids may reach treatment plants at a temperature of about 100C. This temperature is not at all ideal for multiflash stabilization. The opposite condition prevails in offshore field development. In this case, the fluid may cool very significantly inside the subsea pipelines, and the temperature of the oil gathered at the entrance to the separation plants may be too close to that of the seafloor (12C in temperate seas, such as the Mediterranean). At this temperature, it may become difficult to separate the two phases due to the increased viscosity of the liquid. If the temperature at the final stage is not controlled, the vapour pressure of the stabilized product will be too high. Even without taking into consideration cooling due to adiabatic flashes, at the atmospheric stage the oil would have a true vapour pressure at 30C far in excess of an atmosphere (1.8 absolute bar), and the Reid vapour

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

651

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

pressure would be too high. In this case, the temperature of the final separation stage must be raised. Until now, the final stage has been considered atmospheric; actually, if one wanted to recover the gas produced by this separator through compressing it, one can maintain working pressure at values above atmospheric pressure. Even when the gas is burned in a flare, the working pressure is greater than or equal to 1.2 absolute bar. If this new operating parameter is taken into consideration, it is easy to calculate the temperature required to obtain the desired degree of stabilization. In many cases, the optimal pressure value in the separator is in the order of 1.7 absolute bar; for the purposes described above, the working temperature needed to obtain a Reid vapour pressure of 10 psia is 75-80C. This means that significant heating is required before the final stage. As the diagram in Fig. 3 shows, the gas from the final and intermediate stages is recompressed and mixed with the stream from the preceding stage. In the final stage, the gas is cooled before the relevant compression stage and the condensed liquids are mixed with the stabilized crude. As far as the first separation stage is concerned, and frequently the intermediate stage as well, no heating is involved unless there are further problems, which will be discussed later (dehydration, desalting). The low temperature of the first two stages leads to a reduction in the loss of light hydrocarbons in their respective gas streams. The gases are separated under conditions of thermodynamic equilibrium with the liquid under the

relevant temperature and pressure conditions. In conclusion, a multiflash oil separation-stabilization system produces maximum recovery if the temperatures of the first stages are kept as low as possible, and the temperature and pressure of the final stage are appropriately regulated by heating. That which occurs when oil arrives at the separator at moderate temperature (i.e. lower than or equal to that of the environment) was previously addressed. However, when oil enters at very high temperatures as described above, the contrary may occur; in other words, it may be necessary to cool all three gas streams produced by the separators, or even to cool the whole stream entering the separation unit. Clearly, the condensates produced by the cooling of gas, including that from the first stage, are recycled in the separators themselves (in the subsequent stage). This condition is not unusual; in many cases, alongside simple cooling, a true condensate removal system by refrigeration may be used for all of the associated gas (see Chapter 5.4). The liquids produced in this way can be treated and fractionated as described in the chapter cited, and may be commercialized separately if they are produced in sufficient quantities to render this operation economically viable. In many cases, the condensate obtained by its removal from gas can be recycled and commercialized through the crude oil itself, by stabilizing the entire stream. This solution is useful when the gas is later reinjected into the reservoir, either to save energy or due to productive problems with the reservoir. In this case, condensate removal serves to

gas recovery and/or injection

associated gas

MP compressor

reservoir fluid I stage separator

II stage separator wash water

LP compressor III stage separator

oily water treatment

stabilized and desalted oil

Fig. 3. Three-phase separation with heating and desalting. MP, medium pressure; LP, low pressure.

652

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

HP manifold

gas recovery and reinjection MP gas compression LP gas compression

MP manifold

oily water treatment


Fig. 4. Three-phase

stabilized oil

separation with three delivery pressure levels. HP, high pressure.

LP manifold

maximize the immediate recovery of liquids, and there is thus no need to meet restrictions on the dew point of the gas. By contrast, where it is preferable to commercialize the associated gas, a more stringent condensate removal process may be necessary, with the production of liquefied petroleum gases as a third product. The LPGs extracted cannot always be mixed with the stabilized oil stream. In essence, it is not always possible to comply with two conflicting restrictions without producing a third product. It is evident that, starting from a very simple scheme, a larger and fairly sophisticated stabilization system can be obtained. Today this type of situation is common. A so-called ideal separation or stabilization of a reservoir fluid involves subdividing the recombination into two streams: one of vapour and the other of liquid. The latter contains all the liquid components which are liquid at ambient conditions (i.e. the C5), while the former contains all of the volatile components (i.e. methane, ethane, nitrogen and CO2). Regarding propane and butane, which are the key components of the fractionation, the liquid hydrocarbon stream contains normal butane isobutane until the acceptable vapour pressure (RVP0.7 absolute bar) is reached. In many cases, to obtain this value, the above-mentioned stream also contains some of the propane present in the recombined reservoir fluid, while the remainder is found in the gas stream.

This method allows two different types of analysis to be carried out. First, it is possible to calculate simply and effectively the actual performance of a separation system by comparing the oil stream stabilized by two or three separation stages with the liquid stream of an ideal stabilization system; a stabilization system can thus be considered optimal the closer it comes to the solution described. Second, ideal separation makes it clear from the first data available (reservoir recombination) whether and to what extent the reservoir fluids can be subdivided into only two streams, providing a gas and a stabilized liquid oil that meet the basic restrictions for transportation and/or commercialization. This condition, which has enormous economic benefits (smaller investments) and is of great operational simplicity, can be obtained as long as the stabilization of the crude is carried out in an adequate way. In other words, it must be as close as possible to a fractionation, and not a rough separation obtained by flash. As the examples provided show, the latter produces a stabilized liquid stream whose vapour pressure is mainly due to the presence of methane and ethane. These two components can be reduced to a minimum, so that they contribute only negligibly to the vapour pressure of the stabilized oil. This does not require an ideal or overly sophisticated plant (see Section 5.3.4). Clearly, a significant percentage of butanes, except in very unusual cases, can be contained in oil with a

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

653

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

RVP below or equal to 10 psia. Often a non-negligible amount of propane may also be acceptable. In evaluating the multiflash separation system, for the sake of simplicity it was assumed that the production of the field under examination formed a single stream. All of the wells, regardless of operating conditions, thus enter a single manifold at the end of the gathering system, which collects and uniforms the entire production. Note that in this case the entire production is treated under the well conditions with the lowest working pressure when it reaches the gathering centre. In the initial phase of reservoir exploitation, this condition does not cause significant problems since all of the wells generally deliver at similar pressures. However, if these pressures vary greatly, as is the case for some producing fields, several production manifolds operating at different pressures are needed (e.g. high 80 absolute bar, medium 30 absolute bar, low 9 absolute bar). For wells producing at high pressure, the first stage at 80 absolute bar is followed by three further separation stages at medium, low and atmospheric pressure. The production at medium pressure by-passes the first stage and enters directly into the second, where it is mixed with the saturated liquid from the first stage. The production at low pressure enters the third stage and is followed by only a single stage at atmospheric pressure (Fig. 4). Basically, the example given shows that separation parameters, and more generally treatment parameters, must above all be adapted to field production parameters. Furthermore, when associated gas is recovered or injected, a compression system for the latter is needed. Often it is the compression of gas at the various stages of separation (which is far more expensive than separation) that dictates the optimal operating parameters. Taking this into consideration, it is clear that the operating parameters for separation often differ from the optimal parameters for maximum recovery. As far as construction materials are concerned, for a simple gas-oil separation carbon steel is used for all the vessels, even in the presence of corrosive compounds dissolved in the liquid. This choice is motivated by the significant filming tendency of crude oil, which create an effective self-protection of all the surfaces that it wets. Regarding the internals, the wire mesh pad and straightening vanes, stainless steel is generally used (with a variable nickel-chrome content, e.g. AISI 304). In this case, the metal wire mesh pad must be perfectly electrically isolated from its supports, which are welded to the vessel. This avoids the formation of micro-cells due to the different electrochemical potential of the two materials and the consequent passage of a current, which would cause significant corrosion of the separator wall.

Oil-water separation

Until now, only aspects concerning the oil-gas phase equilibrium and the final stabilization of the crude have been considered. Field separation processes also have another aim, that of separating the water phase potentially present in the reservoir fluid. The specification to be met is 0.5% in volume of residual water cut. Under normal separation conditions, when the oil has low viscosity and medium-low density, the separation of water (thus meeting this specification) does not require special features. It is sufficient to use three-phase gas-oilwater separators instead of simple two-phase gas-oil separators; these allow the two liquid phases to be separated as well. This introduces a new topic, concerning the typology and characteristics of the separators used in oil fields, where the most important phase treated is the liquid hydrocarbon phase. For a more detailed gas-liquid separation description, see Chapter 5.4. The separation mechanism used for drops of liquid suspended in the gaseous phase is identical; the coalescence systems for the drops dispersed in the gas phase are also similar, as are the internals used, identical also in terms of installation. The configuration of oil separators often differs from that of gas separators. Except in rare circumstances, the most frequently used separator in oil fields is of horizontal type, whereas vertical separators are more common for the separation of gas. The reason for this is clear: since the liquid phase is the most important, the oil separator must hold the liquid for a period of time sufficient for the gas phase dispersed in the liquid to be separated and reach the interface between the two fluids as the liquid passes through the separator. Adding to this the need to separate the two liquid phases of water and hydrocarbons, it is evident that the holdup time required may increase considerably. Basically, the sizing and/or the operational evaluation of a single separator can be represented by the semplified movement of particles, in this case spheres, suspended in a dispersing fluid (drop dynamics). There are several equations that represent the decantation of particles in a fluid, each with its own field of application; below, that most commonly used for a liquid-liquid separation will be examined. The motion of a drop of liquid dispersed in a gas or in a dispersing fluid, like that of a bubble of gas dispersed in a liquid, reaches a steady condition when the resultant of external forces (force of gravity, buoyancy, the resistance of the medium) is nil. This condition, i.e. in the field of relatively low Reynolds numbers; as for example that relating to liquid-liquid

654

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

600 400 200

viscosity of saturated oil (cP)

100 60 40 20 10 6 4 2 1 0.6 0.4 0.2

500 300 200 100 70 50 40 30 20 15 10 7 5 4 3 2 1.5 1.0 0.7

a bso

l u te

v is c o

s it y o

f gas f ree cru de

at rese r

voir tem p

e rature

Fig. 5. Viscosity of saturated oil as a function of solution GOR (Katz et al., 1959).

0.1 0 100 200 300 400 500 600 700 800 1,000 1,200 1,400

gas in solution (ft3/bbl)

separation and, in this case, the separation of oil and water, can be represented by Stokes well-known law: Vt 1,488gDp2(r1r2)/18m2 where Vt is the settling velocity in ft/s; Dp is the diameter of the particles in ft; g is the acceleration of gravity (32.2 ft/s2); r1 is the density of the dispersed liquid in lb/ft3; r2 is the density of the dispersing liquid in lb/ft3; m2 is the viscosity of the dispersing fluid in cP (1cP 103 Pas). The formula under examination does not apply to the separation of drops of liquid dispersed in a gas since the Reynolds number, in this case, is always very high; however, it is extremely relevant to the analysis of the separation of drops of water suspended in a liquid hydrocarbon phase. Its field of application, concerning low Reynolds numbers, is compatible with the size of the suspended particles present (up to 100 mm). The equation highlights the three essential parameters of the analysis being carried out: the diameter of the particle (which has the greatest impact on the phenomenon, being raised to the second power), the difference in density between the two liquids, and the viscosity of the dispersing fluid (i.e. the oil). It is important to note that these physical properties must be referred to as the temperature and pressure conditions of the separator. In the case of the hydrocarbon phase, this means that density and viscosity must refer to the saturated liquid at both temperature and pressure conditions, which differ considerably from those of the oil under stock tank conditions made available through analyses. At a fixed temperature, the solution GOR of each separation stage

is quite proportional to the operating pressure which is at its maximum in the first stage. This fact leads to a considerable reduction of both the density and the viscosity of the saturated oil. Since the density of the two liquid phases is of the same order of magnitude, even a small variation in the density of the oil may increase the difference in density between the two liquids significantly, and thus the corresponding settling velocity. The data from PVT analysis or simple flash calculations make it possible to determine the solution GOR and the density of the saturated liquid. As far as the viscosity of the latter is concerned, Fig. 5 allows one to calculate the reduction in viscosity in passing from stock tank oil to saturated oil as a function of the solution GOR. For example, assuming an oil having the following characteristics: stock tank density of 0.842, viscosity of 10 cP at 40C to be treated at a working pressure of 35 absolute bar and a temperature of 40C, corresponding to a solution GOR of about 250 SCF/stb (from the material balance), with the help of Fig. 5 the viscosity of the dispersing fluid can be determined, which turns out to be 3.3 cP. The variation in viscosity alone leads to a settling velocity three times higher in the first stage separator than in the atmospheric separator. Given an identical temperature, the effect of variations in density must also be added; though modest, these have a similar effect; corresponding density of saturated oil with a solution GOR of 250 SCF/stb becomes 0.8 ca. Dr19.86 and Dr212.48 lb/ft3. Overall, the ratio of the two settling velocities at identical temperature is 3.83. It is easy to see that the holdup time in the separator can be reduced proportionally in order to obtain the same degree of

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

655

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

separation. Alternatively, taking into account the impact of the diameter of the particles, with an identical holdup, drops with a diameter 3.8321.96 times smaller can be decanted. The above example makes it easy to understand the influence of the main operating parameters on the separation of water dispersed in oil. To obtain more detailed calculations, and to calculate the size of a separator guaranteeing the required residual water content, the distribution of the particles in a water oil emulsion must be known. Although numerous experiments have been carried out on this topic, it is extremely difficult to obtain precise data: it is not easy to quantify the degree of emulsion of water in a crude. This clearly depends on the properties of the two fluids, but it is certainly not sufficient to know their density and viscosity to quantify the degree of dispersion of the water drops. This depends on the path followed by the emulsion from the reservoir, through the well and its pressure control systems at delivery, up to arrival at the gathering system, and on the pressure drop during the stages of separation. In a three-phase gas-oil-water separation, some pieces of information, dictated by operational experience, allow this problem to be overcome in part. A liquid-liquid separation, unless anomalous factors such as those relating to extremely viscous oils intervene, may be carried out quite simply. It is common knowledge, supported by numerous field tests, that the size of the water drops to be separated in order to obtain an adequate dehydration must be in the order of 100-150 mm. This means that, if the separator is designed so that drops with a diameter larger than or equal to the predetermined diameter can make the vertical journey from the gas-oil interface to the oil-water interface, a residual water content lower than the required 0.5% in volume can be obtained. Suppose, for example, that the cylindrical part of the horizontal separator has a diameter and length of 2.5 and 10 m respectively, and that the gas-oil interface is
momentum absorber sive tray demister

maintained at the middle of the vessel. Consider that the oil-water interface is kept 450 mm above the bottom of the tank and that the holdup time for the oil is about 5 minutes. The water particle to be decanted must cross a maximum layer of 800 mm in 5 minutes in other words, it must have a settling velocity of 160 mm/min (0.52 ft/min). Applying Stokes law for a particle diameter of 100 mm would result in a velocity eight times lower than this, and, therefore, it would be impossible to reach the required value for residual water cut. On the other hand, if one calculates backwards to obtain the diameter of the drops that can be separated, a value of about 300 mm is obtained, which unfortunately provides an inadequate separation. At this point, coalescence devices should be introduced. In the case of drops suspended in a gas, a wire mesh pad is inserted in the vertical separator used to decant the drops to obtain a good separation. This simple and cheap device guarantees an adequate coalescence of the drops of liquid suspended in the gas, forcing them to collide in the obligatory path through the mesh itself. A similar result is obtained in normal liquidliquid separation, as in the case of light condensate, LPG and gasoline, by inserting a coalescence device identical to that used for gas into the liquid-liquid separator. With oil, on the other hand, it is impossible to use this type of coalescence device, suitable for light and extremely clean fluids. Oil requires a coalescence system that, working with a fluid with a high content in sediments of various type, must provide large spaces through which to pass. This problem is tackled by inserting a coalescence section consisting of straightening vanes, suitably distanced and placed at different angles (45 or 60) parallel to flow inside the separator (Fig. 6). Usually this occupies the whole section through which the three phases of gas, oil and water must pass; the straighteners also facilitate separation in the gas phase, and in the bottom water phase. In order to facilitate deposition, and thus the

gas straightening vanes oil buffle water vortex breaker

three-phase separator A B

Fig. 6. Internal view of a three-phase separator (A) and straightening vanes (B).

656

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

removal of suspended solids (sand, colloidal clay, etc.), it is sometimes preferable to leave the lower part, devoted to the water phase, free. The distance that the drop must cross is reduced to extremely low values. If we have an angle of 45, the vertical distance between two vanes is 22i, where i is the distance between the two vanes (50-75 mm). If the whole separator is filled with the pack of straightening vanes at 50 mm intervals from one another, this leads to a reduction in the diameter of the separated particles of (1.450/800)20.3; in other words, a value slightly below the required value of 100 mm. When the drops, whose size is larger than that calculated, reach the surface of the vane, the wall effect causes them to coalesce and the separation reaches the required value. However, the example given is a limited case; usually heating is used to improve separation. Longer holdup times can also be applied, in part because it is not practically viable to fill the entire cylindrical part of the separator with straighteners. If the gas flow rate is very low, as in the two stages following the first stage, the level of the liquid can be kept above the centre line to increase the holdup time without changing the size of the separator. It is important to note that this solution is effective if straightening vanes are present. If the inside of the separator is empty, increasing the holdup by increasing the height, and thus the vertical journey, is not particularly useful; as a result, vertical separators are rarely used for this purpose. For the separation of drops of liquid from gas see Section 5.4.2. Oil-water separation is not significantly different from gas-oil separation in terms of the choice of materials. The straightening vanes described above are also made of extremely thin sheets of stainless steel (AISI 304, or other even more valuable types). If the associated gas and the oil contain acid gases, these concentrate in the water phase proportionally to their fractional pressure. If this pressure is high (greater than or equal to 1 bar), the watery solution becomes acidic, and thus corrosive. The protective film formed by the oil also disappears. During the pressurized separation stages, the part of the tank below the water-oil interface is protected by clading with stainless steel (AISI 316-L or duplex steel 3 mm thick). Finally, it should be noted that corrosion in the water phase is often increased by the high concentration of salts.

5.3.4 Treatments
Desalting
Desalting by washing with water

The above discussion makes it possible to stress that a good three-phase separation is able not only to

stabilize the oil, but also to dehydrate it to the required values. Returning to the properties of formation water and particularly its salinity, as said earlier, it is often necessary to desalt the crude by reducing the salinity of the emulsified water. A desalting system consists of a mixer-settler type system which allows the salinity to be diluted by mixing the oil with washing water with a low salt content (1 g/l). The diluted water is then separated out by decanting, thus bringing the oil to the same or lower value for water cut as before mixing. This equipment can be inserted downstream of separation as an additional unit. In this case, the intake to the desalting system is a separated and stabilized crude, but with excessive salinity. Consider, for example, a crude with a water cut of 0.5% in volume and a water salinity of 50 g/l. The salinity of the oil is thus about 87 ptb, whilst the value to be guaranteed is 20 ptb. Assuming an identical final water cut, the system therefore requires an effective dilution of about 1/5. Since the quantity of water to be diluted is modest, injecting an amount 3% in volume or greater is sufficient. Often dilution is obtained by injecting water into the feed line. The most common mixing device is a lamination valve which, by producing sufficient turbulence, facilitates contact between the two phases of oil and water. Using this system leads to a considerable loss of head; in the case under consideration, downstream an atmospheric separator, a feed pump will be needed to transfer the intake from the separator to the desalter. The mixing efficiency is obviously less than one. If we use a mixing valve and have a very low washing water/crude ratio, 3% as in the example given, the efficiency of the mixing process is not much above 70%. Since the treated product has a fixed water cut of 0.5% by volume, a simple in-out balance for salt indicates that the content is 10.4 g/l. The salinity of the crude is thus 18.1 ptb, meeting the required value. It is important to note that in the example given it was assumed that the washing water has a salt content equal to or less than 1 g/l. If brine with higher salinity is available, it is sufficient to raise the washing water/feed ratio slightly. In this type of equipment, using a water flow rate below 3% leads to inefficient mixing; it is therefore pointless to improve the degree of liquid-liquid separation to a water cut below 0.5% by volume. The separator used thus works exactly like that described above. It is obvious that a multiflash separation as described above allows this new function to be integrated without the need for additional equipment; the washing water is simply injected before the final separator. As such, the first separation stage functions as a dehydration stage, reducing formation water to 0.5% in volume as seen above.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

657

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

Washing water can also be mixed by other devices, such as a static mixer. This equipment closely resembles structured packing, and supplies the necessary surface area to create close contact between the two phases of oil and water. Basically, a true stage of equilibrium in equicurrent is obtained. The latter characteristic affects both efficiency, which becomes very close to one, and the loss of pressure, which becomes negligible. Another advantageous property of this equipment is that it mixes by contact over a large surface area, which does not cause excessive dispersion and the consequent stabilization of the emulsion, as is the case if a valve is used. This phenomenon also occurs in the multiflash separator described above, where the oil passes through the level control valve, rendering subsequent decanting less efficient. In the example analysed, desalting conditions are relatively favourable. We have considered a crude of medium density, but with relatively high viscosity (in most cases viscosities are lower, and densities more favourable). The treatment described can therefore be used with excellent results and modest water consumption. When desalting treatment is carried out offshore, it is obvious that the dilution water used cannot be seawater as it is. Treating the crude therefore becomes more expensive since the seawater must first be desalinated, using one of a variety of methods. The simplest and most frequently used is based on reverse osmosis and can supply water with negligible salinity by using two stages of separation with semipermeable membranes. Since the crude is usually very salty, it is not necessary to completely desalinate the washing water. A single stage of purification is therefore enough, supplying water with a residual salinity lower than or equal to 1 g/l, sufficient for this purpose. Under other conditions, such as in the desert, it is possible to use brackish water with a salinity of 2-3 g/l without notably increasing the amount of washing water. The waste waters from crude oil dehydration and/or desalting are oily waters containing large amounts of suspended solids, and therefore require suitable treatment before they can be discharged. In liquid-liquid separation the watery phase at the bottom of the separator also behaves like an emulsion in which the dispersed liquid is the oil itself, and the dispersing phase is water. In this case, during the holdup in the separator, the hydrocarbon drops dispersed in the water follow an inverse process to that of the overlying oil phase. In the example separator described above, the bottom section devoted to water is kept at a predetermined level (450 mm). By varying this parameter, the water can be given the minimum holdup time required to obtain the desired purity. To avoid overloading the subsequent water deoiling treatment excessively, the degree of separation most

frequently required from separators is 500-1,000 ppm of oil dispersed in water, in other words 0.05-0.1%. This specification is more restrictive than that for oil. However, it should be remembered that the viscosity of the dispersing fluid is very low (0.6-0.4 cP) under the temperature conditions of the first and third stage respectively. By applying Stokes law as above, since the difference in density is identical, and viscosity 10 times lower, a suspended particle of identical diameter has an ascending velocity ten times higher. Adding to this the fact that the distance to be covered is shorter, slightly less than 5 minutes are needed for a particle to travel the entire vertical distance. As it is difficult to maintain holdup times lower than 3 minutes and obtain a good level control the required clarification of the water inside the separator can therefore be obtained without the need to use straighteners or changing the holdup time. The water decanted from the separator is first degassed in a flash drum, and then deoiled with a primary treatment consisting of a horizontal settling tank (API skimmer) or a CPI (Corrugated Plate Interceptor), and with secondary treatment (filtration or floatation). For offshore treatment, and recently for onshore treatment as well, the API skimmer has been replaced by cyclones based on the centrifuge principle, which allow acceptable values for suspended oil (30 ppm) to be obtained. International standards for open sea conditions specify a limit value of 40 ppm, whereas for inland and coastal waters the most frequently required value is 10 ppm. In this case, deoiling with secondary treatment, the most common of which is floatation, is needed.
Electrostatic desalting

This treatment process is common practice in refineries where the feed to the primary treatment of crude fractionation (topping) requires a salt content below 3 ptb. In refineries it is more important than in field treatment to keep the flow rate of washing water to a minimum, and this can be done by minimizing the salinity of the crude oil entering the system, following the criteria described above, and above all by keeping residual water after separation to values of 0.1-0.2% in volume. These specifications are easy to meet with medium-light or light crudes, even in field treatment, especially when the viscosity of the stabilized oil is low (1-2 cP at ambient temperature). In simple multiflash separation, as a result of gas saturation in the first stage and heating in the following stages, extremely low viscosities, below 1 cP, can be obtained; these allow water contents far lower than the required 0.5% in volume to be obtained with only minor modifications to the sizing of equipment. The greater ease of separating

658

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

emulsified water in the field as opposed to the refinery can be explained by the different properties of the crude. When the oil exits the wells it has not undergone oxidation processes. Under original conditions its highboiling components, especially asphaltenes, are those found in the reservoir. On contact with air, which occurs in the stock tank as a result of breathing, and which continues in the oil tanker used for transportation, the crude undergoes a weathering process which changes it profoundly, increasing the presence of strongly surfaceactive compounds, which dissolve or accumulate as colloids suspended in the drops of water. This change in composition, added to the reduction in light compounds which have a fluidifying effect, causes a marked increase in the strenght of the emulsion, although the increase in viscosity is not equally significant. A practical and simple system for evaluating the effects of the weathering of the crude and the stabilization of emulsions is that known in technical jargon as a bottle test. These tests involve taking liquid samples which are heated and centrifuged in a standardized equipment for the time required to separate out the water. The tests are also carried out with the addition of various demulsifying compounds in order to evaluate their effectiveness. If a bottle test is carried out at the mouth of the well, even if degassing is carried out in a very rough way, the results, compared to those which can be obtained in the laboratory with a weathered sample, are completely different and better. In short, the temperature at which the emulsion breaks is lower, and the separation time is lower. For this reason, refineries use an electrostatic desalting system. This is carried out hot, in other words the treatment forms part of the system used to preheat the feed to high temperatures (100-130C), even for crudes which are not particularly viscous or dense. Heating and the electrostatic coalescence of the suspended drops of water is often coupled with the use of demulsifying chemicals, which ease the process of breaking the emulsion. The principle on which electrostatic coalescence is based is easy to understand. The emulsified water drop has a strong surface tension which prevents, or hinders, the aggregation of several drops, which would allow these to reach the settling velocity required for separation. This is due to the presence of modest quantities of surface active compounds. The drop of salty water, being a polar substance, behaves like a dipole when subjected to an electrostatic field. The application of a high voltage alternating electric field (10,000 volts or above) in addition to orienting the dipoles, elongates the drops, deforming their spherical shape into a ellipsoid, in which the larger diameter becomes ten times larger than the original diameter. This alternate stretching has two positive effects: the film covering the particle is thinned,

facilitating its rupture, and above all, it eases the collision of the drops, which find themselves at a distance of less than ten diameters from one another. This causes an enlargement of the drops, which thus settle more easily; the mean free path between the various particles becomes ten times the diameter of the particles. The latter consideration makes it easier to understand why the distribution of the drops created in the suspension is extremely rarefied, and allows us to guarantee a final content of 0.1-0.2% by volume. Whereas in refineries, due in part to the higher flow rates, electrostatic desalters are used for mediumlow viscosity oils, in the field this type of treatment is used only for extremely heavy and viscous crudes. It should be stressed that applying an electric field of several thousand volts to an emulsion of water in oil is possible because of the high resistivity to the passage of electric currents the emulsion has. This means that the difference in voltage applied to the grids, which causes the effects described above, can be maintained with a very modest consumption of electric power. For this to occur, the emulsioned water must not be too fragmented, and the amount of water in the emulsion must not exceed 15% by volume. With higher water cuts, or because of too high mixing valve pressure drop, the electric conductivity of the bath will exceed allowed values and the electrical field short-circuits. To avoid this condition, the emulsion must undergo suitable pretreatment to reduce the water content to the value mentioned. A second characteristic of the use of the electrostatic coalescence separator is the absolute guarantee that no gas bubbles will form inside it. Many vendors of this equipment prefer to feed the electrostatic coalescer with a pump, so as to ensure an operating pressure in the vessel of at least a bar above the bubble point of the treated fluid. Usually, electrostatic treatment is carried out in combination with heating; it is therefore necessary to evaluate the vapour pressure of the hydrocarbons carefully. At high temperatures, water also has a significant vapour pressure which is added to that of the hydrocarbon phase. The vessel in which electrostatic coalescence is carried out is therefore pressurized. Pressure increases as the working temperature rises, and decreases the more the oil to be treated has been stabilized during prior treatment. This is why this treatment is often inserted downstream the final separation stage. On offshore platforms where equipment are distributed on several floors, the desalter is usually placed immediately beneath the final degassing stage. This avoids the need to use a feed pump for the crude. Although electrostatic coalescence facilitates separation, it is good practice to use heating as well, in order to shorten the holdup time, or to produce the

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

659

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

water treatment and reinjection free water KOD compression and collection

stabilized and desalted oil fuel gas

inlet separator

treated oil flash drum

electrostatic coalescer II stage

electrostatic coalescer I stage

flash drum

wash water water to treatment

Fig. 7. Oil desalting (20API) with water separation at the wellhead and two-stage desalter at the gathering and treatment centre.

required water cut. To return to the example of desalting with multiflash, it was demonstrated that modest heating allows the crude to be stabilized at higher than atmospheric pressure. If the electrostatic solution following the third separation stage is chosen, the heating used for stabilization facilitates desalting at no further cost. When a holdup time above 20 minutes is used for electrostatic treatment, the maximum viscosity at which a good water separation (0.5% in volume) can be obtained is 10 cP. This value makes it clear that recourse to this process is limited to very viscous and very heavy crudes (20-25API). Sometimes an electrostatic coalescer is used cold to avoid heating.
Multistage desalting

Washing water represents a significant cost in the desalting of crude oil, and it is thus preferable to keep this to an absolute minimum. When the formation water is extremely salty, it may be helpful to use a two-stage treatment system. This unit often consists of two electrostatic coalescers. The first, which receives the maximum water intake, acts as a dehydrator; the second carries out the desired dilution with a water flow rate sufficient to guarantee the requisite oil salinity.

Consider, for example, an oil with formation water whose salinity is 300 g/l, and specify as 20 ptb the salinity of the treated oil. If the first stage functions simply as a dehydrator, the lowering of salinity will depend exclusively on the final water cut (0.2% in volume), and not on the quantity of water entering the system. Assuming a maximum value for water of 15%, this gives a reduction of 96%. Nevertheless, salinity remains very high. The second stage therefore involves dilution with washing water. Assuming an amount of washing water equal to 3% and a mixing efficiency of 70%, the salinity of the water remaining in the crude is about 26 g/l; with a residual water cut of 0.2% a value of 18 ptb is thus obtained, in other words the requisite oil salinity. Taking into consideration the fact that the desalting system forms part of the stabilization process, it is obvious that a good gas-oil separation can reduce the water content significantly, at least to the value of 15% needed for the electrostatic separator to function. In many cases, during the productive life of the reservoir, the crude produced at the wellhead may contain 50% or more water. Under these conditions the well can no longer produce spontaneously, and it will often be necessary to use an artificial oil lift system with submerged pumps installed inside the well. In this case

660

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

the emulsion of the formation water phase in the oil reaches its maximum levels. The emulsification due to the passage of the reservoir fluid through the pump must therefore be reduced by injecting demulsifiers directly into its intake. This is followed by a gravity separation directly at the wellhead, without the help of an electrostatic field. This operation is aided by several factors. The first is the delivery condition of the well itself. Often a rough separation of associated gas is carried out at the intake to the pump, before the reservoir fluid enters the pumping system. The pumped liquid is thus at a pressure above its bubble point and remains so at the wellhead even when the delivery pressure is relatively low. If we assume 15 absolute bar as a pressure value under these conditions, the oil does not liberate gas for the reasons outlined above, but has a solution GOR which, though modest, renders the fluid less viscous and less heavy. The second factor facilitating separation at the wellhead is the working temperature, which is fairly high since the heat exchange along production tubing does not lead to significant cooling. Assuming a flow temperature of 90C, an extremely common value for an oil well, and adding together the effects of the dissolved gas and temperature, it is clear why both viscosity and density are much lower than those of the dead oil. In conclusion, it can be stated that a significant amount of the water produced can be separated without heating and/or electrostatic treatment. The separated water, given the quantities involved and its properties, is collected and reinjected into the reservoir following pretreatment with cyclones and filtration. The oil treated in this way at the wellhead still has a high water content (5-10%), which can nevertheless be guaranteed with modest holdup times and simple separator internals (straightening vanes). The oil is transported with this residual content to the gathering centre, where it is treated using the systems described above. Sometimes it may be convenient to accept a higher water cut, up to the standard 0.5% by volume, and to use a more sophisticated washing system (Fig. 7). Two-stage desalting is carried out with two washing stages in counterflow, using the water recovered from the second desalter in the first stage. Assume a salinity of 180 g/l and a water/oil feed ratio of 5% by volume; with a residual water cut of 0.5% it is sufficient to inject 5% of water. This solution, assuming a mixing efficiency of 80% in the second stage, gives a dilution ratio of 1/9, with a decrease in salinity to about 11% of the intake value. This gives a salinity of residual water of about 10.5 g/l, guaranteeing the required salt content. The water decanted with a lower salinity (8.5 g/l) can therefore be used to reduce the salinity of the first stage from

180 to about 95 g/l. Using this system, the desired result can be obtained with a small amount of dilution water. Adopting the basic scheme of dehydrator plus desalter, twice the amount of water would be needed. This result is obtained by adding a pump to recirculate the decanted water from the second to the first stage. As far as the choice of materials for desalting is concerned, the same considerations apply as for oil-water separation, with particular emphasis on high salt concentrations, and their negative impact on corrosion, further accentuated when a high working temperature is needed. Sweetening When an oil has a high hydrogen sulphide content, the gas-oil separation and stabilization processes must necessarily take due account of this. The acceptable contents were indicated above. In the past, a reduction of H2S in an oil which did not have particularly high contents of this contaminant was obtained simply by increasing the temperature of the final atmospheric separation stage sufficiently. This solution can be considered makeshift, since hydrogen sulphide has a volatility very close to that of propane. With a specification of 50 ppm by weight, in any case an extremely high value, the oil must be stabilized excessively, increasing the loss of light products, not only propane but also butanes and gasolines. This procedure may be defined as a makeshift solution. With flash distillation it is not possible to obtain a precise fractionation of key components. In other words, this solution may be acceptable when the H2S content is relatively low, and therefore does not affect the stabilization parameters for the oil. In all other cases, a specific treatment known as sweetening is needed. In the past, the cold stripping method was commonly used. This system uses a stripping column working at close to atmospheric pressure, in which a gas stream with a low-medium hydrogen sulphide content removes this component from the oil stream to be treated (Fig. 8). This operation is known as cold stripping since it is unnecessary to heat the oil in order to carry out the removal required. Were it necessary to previously sweeten the gas used for stripping to obtain the desired effect, this solution would no longer be simple and cheap. In a stripping column the result of this process can be described by the stripping factor SKiV/L, where Ki is the equilibrium constant of the i-th component, in this case the H2 S, V and L are the molar flow rates of the vapour (stripping gas) and liquid (oil to be treated) respectively. The equilibrium constant, as is well-known, expresses the ratio between the molar

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

661

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

gas collection/reinjection

gas/oil I stage separator gas/oil II stage separator

cold stripper

flare gas atmospheric separator stabilized and sweetened oil

Fig. 8. Gas-oil

separation and sweetening by cold stripping with associated gas.

fraction of the component in the vapour and in the liquid phase, in other words Ky/x. Fig. 9 shows values of K for H2S plotted on a graph using the convergence pressure method. This method can be used for low pressures, whereas it is of little use for high pressures, close to the convergence pressure itself. Assume a working temperature of 38C and a pressure of 1.5 absolute bar (150 KPa): Fig. 9 gives K23. Therefore the concentration of oil in equilibrium with the stripping gas is xy/K; in other words, at equilibrium the molar fraction of H2S in the liquid phase becomes 23 times lower than that of the stripping gas. Assume a concentration in the stripping gas of 0.6% mol, thus quite high, and consider an oil with a mean molecular weight of 225. The concentration of oil in equilibrium for H2S is Xe0.6/230.03% mol. Since the molecular weight of H2S is 34, this gives Xe40 ppm in weight. It is therefore possible to sweeten the oil to the required standard of 50 ppm. The difference between the value at equilibrium and the guaranteed value makes the number of stages in counterflow needed to obtain the desired result acceptable. In conclusion, it is possible to sweeten an oil with a relatively high H2S content by using the gas associated with the crude itself. The latter is a proportion of the gas from the stage at highest pressure with a lower H2S content. At a pressure of 15 absolute bar and at an equal temperature, the equilibrium constant K becomes about 2.3; this explains the lower concentration of H2S in first stage gas. This type of oil treatment, which is extremely simple and effective, has lost its appeal as oil prices have risen. Stripping gas is poor not only in H2S, but also in higher hydrocarbons. During the mass transfer

with the crude, the gas removes not only H2S, but also propane, butane and gasolines, since with an intake saturation at 15 absolute bar, it exits in equilibrium with the crude at 1.5 absolute bar and at the same temperature. This unwanted stripping of light hydrocarbons represents the main limitation of this treatment system. A second limitation is linked to the final hydrogen sulphide content of 50 ppm by weight. This value, though acceptable from the point of view of safety and corrosion control, is not sufficient to
1,000

100 60 40 30 20 10

26 te 0 m pe 12 ra 1 66 20 tu re 4 9 ( 3 38 C) 14 16 9

Ky/x
1.0 0.1 100 300 500 1,000 3,000 10,000

pressure (kPa)
Fig. 9. Equilibrium constant

for hydrogen sulphide (GPSA).

662

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

guarantee good total acidity levels in the crude, and thus causes a significant drop in price. All sweetening processes, given the particular chemical composition of the fluids treated, present problems of corrosion. In this type of treatment, working at ambient temperature, carbon steel with suitable corrosion allowances (3-5 mm depending on the concentration of acid gases) can be used for all the equipment wet by the oil, and particularly the stripping column. When the gas injected is saturated in water at column bottom conditions, the latter must be protected with a stainless steel plating. Stabilization and stripping with a cold column The paragraph on separations described a criterion allowing the efficiency of a stabilization system to be determined. Below, we will describe the plant that comes closest to the ideal distillation objective and allows for considerable savings, since it eliminates the necessity for one of the two stages of compression of the separation gas required in order to recover the latter. The stabilizing column operates under pressure, and the top gases are thus recovered directly without the need for compression. Fig. 10 shows a diagram of how this unit works. The oil from an initial separation stage at medium pressure (21 absolute bar) is fed into the top of the

column working at a pressure of 7 absolute bar. At the bottom of the column is a reboiler allowing it to function. Controlling the temperature of the bottom product ensures that the required crude bubble point is reached, and thus the required stabilization. In order to keep the vapour and liquid profiles in the column as uniform as possible, a second source of heat is included, with the insertion of an intermediate reboiler. This reboiler is fed, on the process side, by withdrawing all the liquid using a chimney tray inserted at the chosen point in the column. The heat required for this service is supplied by the stabilized product stream exiting the bottom of the column, which thus cools. This system has replaced the traditional preheating of the feed with the bottom product. The most interesting result of this solution is precisely the ability to keep the feed cold, thus eliminating the need for the top condenser present in normal fractionating columns. This makes the unit far simpler and cheaper. The main purpose of the stabilizing column is to guarantee the vapour pressure of the crude, for example RVP0.7 absolute bar, whilst operating at pressure. Though it is impossible to develop the entire thermodynamic calculation for the column with simplified methods (this can be done better with the use of a suitable simulation programme), it is possible to provide a series of initial calculations which allow the basic parameters of the process to be identified.

overhead compressor gas reinjection dehydration and degasolination MP compressor

stabilizer

stabilized oil side separator

reservoir fluid

gas/oil/water separator

oily water

oily water heating fluid


Fig. 10. Stabilization with a fractionation column and degasolination.

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

663

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

10,000

1,000

vapour pressure (kPa)

100

10
carbon dioxide ethane ammonia propane freon 12 i-butane n-butane freon 11 i-pentane

critical point extended beyond critical point n-pentane n-esane n-eptane water n-ottane n-nonane n-decane n-undecane n-dodecane

0.1 0 25 50 75 100 125 150 175 200 225 250 275 300

temperature (C)
Fig. 11. Vapour pressure of light hydrocarbons (GPSA).

At an operating pressure of 7 absolute bar, using the diagram of vapour pressures for individual components as a function of temperature (Fig. 11), the bottom temperature can be roughly calculated. Bearing in mind the type of stabilization underway, the product is free of dry gases, methane, ethane and light inert gases, and has extremely modest H2S content, such as to contribute only in an insignificant way to the vapour pressure. The latter therefore depends mainly on the propane and butane content. The true vapour pressure which we wish to obtain in this case is 0.98 absolute bar at 37.8C. If this value is entered in the diagram mentioned, tracing a line parallel to the normal butane curve, we can determine the temperature needed to reach the 7 absolute bar of the fixed bubble point, that is the temperature at the bottom of the column. This value is roughly 150C. The temperature difference between the feed and the bottom product is considerable, sufficient to create a significant reflux inside the column. In practice, a saturated gas under feed tray conditions liberates at the top; if the stabilization required is not excessive, this can be kept at a temperature close to that of the feed itself (for example 45C). The result of this reflux is that the bottom product does not contain methane and ethane, but has a significant propane and butane content. This shows that the cut obtained between the components is very clear. In fact, if the aim is merely to stabilize the oil, its purity is of less interest than the maximum recovery

compatible with the desired vapour pressure. If the ideal fractionation described above is compared with the true fractionation obtained with this stabilization, it is evident that the recovery is extremely high (about 99% wt). Any losses can basically be ascribed to the first stage of flash separation. If the gas stream is subjected to condensate removal, recovery becomes almost total (99.8% wt); on the other hand, if we examine the sweetening of the oil, that is the removal of H2S, the cut obtained ensures a purity far higher than that obtained with the cold stripping system described earlier (6 ppm as compared to 50 ppm). This solution thus guarantees a good lowering of hydrogen sulphide content even from highly contaminated intakes, without compromising hydrocarbon recovery or altering the parameters required for simple stabilization. As well as being advantageous under normal conditions, it is obvious that this type of stabilization becomes the optimal solution when the crude must also be sweetened, since it does not entail any additional expenditure. Obviously the heating of the crude in the reboiler is different from that required for multiflash stabilization, both in terms of duty and temperature levels, but is not much higher than that needed to desalt a heavy crude. The design with an intermediate reboiler and total withdrawal has another peculiarity. The stream under examination is at an intermediate temperature between the top and the bottom (about 70C), thus not

664

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

particularly high. Most importantly, due to the reflux, the sidedraw has very different properties from the feed crude, in fact it has a much higher light hydrocarbon content in propane, butanes and pentanes than the initial intake. If the original crude has medium-high viscosity, for example 5 cP at 40C, the feed to the column has a lower viscosity (about 3 cP), due to the saturation gas. Flushing with a very light product, which has the physical properties of a LPG and therefore low viscosity (less than 0.1 cP under side separator conditions) and a volume ratio of about 25%, causes a very marked reduction in the viscosity of the stream under examination, which becomes about 1 cP. This example demonstrates that the liquid from sidedraw is much lighter, and has low viscosity, even starting from a medium crude with high viscosity. If the intake to the column contains a residual water cut, this can thus be decanted very easily by sidedraw. To obtain this result it is important to insert a liquid-liquid separator before the side reboiler, allowing the residual water in the side stream reentering the lower section of the column from the reboiler to be reduced to less than 0.1% in volume. This unit must have a side separator since all the water present in the feed, being unable to exit with the bottom product or the top product, accumulates in the side separator. This makes it possible to combine stabilization with a more effective desalting system, which is far cheaper and less bulky than what could be obtained with an electrostatic coalescence stage. If this solution is completed with an appropriate dilution with washing water, salinities far lower than those required for field treatment can be obtained, comparable to those specified for refining. Since the operating temperature of a stabilizing column is very low compared to that of a topping column, the normal specification of 20 ptb or the 10 ptb which can easily be guaranteed with this method, are more than enough to allow the lower part of the column below the side separator to function optimally. However, in order to reduce fouling and excessive maintenance work, it is essential for both reboilers, at the side and bottom, to be protected from the wash in question. It is thus unnecessary to use an electrostatic treatment stage before the column. The waste water from washing and separation thus obtained can be recycled into the separation stage which feeds the column. In many cases, there are more than one preceding separation stage, especially when the delivery pressure of the wells is very high. In this case, the first high pressure stage facilitates a reduction in formation water due to simple gravity separation (residual content below 2% in volume). The second stage can ensure initial dilution and desalting using water recycled from the side separator. Currently, all the gases associated with

oil production must be recovered, including those liberated during the low pressure flash stages. The compression and cooling of a rich gas cause water and light products to condense; these are recycled into the crude. Given their high volatility, these light hydrocarbons remain in the crude only partially, largely re-evaporating due to the presence of methane and ethane in the degassed oil. During the stabilization process a significant quantity of the light hydrocarbons and LPG recovered from associated gas remains in the oil, especially when the required vapour pressure is relatively high. In practice not only the liquids resulting from simple compression and consequent cooling, but also the condensates removed from all associated gas can be recovered in the crude oil stabilization system, obviously without excessively increasing recovery. This solution allows a partial commercialization of the LPG present in the gas without the costs of separate production, which are frequently excessive. In conclusion, stabilization with a column allows for considerable operational flexibility and further expansions of the field development plan which were not initially foreseen. Many units for the stabilization of oil and light condensates use a furnace as a bottom reboiler. According to many engineers this solution is more compact and economical. Essentially, it derives from a simple transposition of a classic refining design, the topping heater. In field treatment, given the far lower working temperatures, it is neither necessary nor useful to adopt a solution which creates significant safety problems. This type of installation also requires pumps to withdraw the liquid to be reboiled from the bottom of the column and circulate it in the furnace. As it returns to the column, the partially vaporized product is separated, and the liquid becomes the stabilized product; the furnace thus acts as the final stage of stabilization. In order to adopt this solution, the bottom of the column contains a divider; on one side the liquid from the last tray of the column is withdrawn with a pump; on the other side is the vessel for the stabilized bottom product. In addition to the inherent problems of the furnace itself, this solution requires circulating pumps to withdraw boiling liquid. This entails a need to raise the column floor sufficiently. The pumps and furnaces mentioned above need sophisticated regulation and a series of protective measures which render the control system extremely complex. A different solution is to use a reboiler heated by an intermediate fluid, hot oil or vapour at medium pressure. If the final stabilization temperature is 150C, and the temperature at entry to the reboiler is just above 100C it is not necessary to have a high temperature heating fluid. It is not even convenient to

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

665

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

save on the exchange surfaces by using a very hot fluid in the case of hot oil. If vapour is used, which has negligible resistance to heat exchange, and we have a good exchange coefficient on the fluid process side, linked to a good vaporization of the crude, it is sufficient to maintain the temperatures on the hot side in the order of 175-180C, corresponding to a saturated vapour at about 9 absolute bar. All that is needed is therefore a simple water-tube boiler to generate saturated steam at medium pressure; this is a standard and very compact equipment, transportable even for large treatment capacities. For a treatment line of 50,000 bbl/d, a 20 t/h boiler or just above is sufficient. Considering that the limit for transportability is above 40 t/h, lines with a higher stabilization potential, up to 100,000 bbl/d, can be used, fed by a single boiler which is easy to install and inexpensive. With a hot oil system, still in the context of transportable equipment, the potential must be reduced (a 6 Gcal/h heater, corresponding to 12-13 t/h of saturated vapour, is at the limit of transportability). For this reason, using this type of heating, the stabilization line would have a maximum treatment capacity of about 30,000 bbl/d. Similar potentials can be obtained by using a direct heater as reboiler. In many units, kettle-type reboilers are used. This solution is not optimal under plant operating conditions. The oil to be heated is not a very clean fluid. A considerable amount of heat is lost as sensitive heat, and the vaporized oil hardly ever exceeds 15% in weight of the intake at the reboiler. For this reason it is preferable to use a once through natural circulation reboiler. Obviously this solution should also be extended to the side reboiler. From the sidedraw tray the liquid feeds the reboiler (horizontal shell and tubes) installed on the level below. The liquid column differential created between the descending stream of saturated liquid, and the return stream, two-phase with vaporized, is sufficient to ensure natural circulation through the reboiler. Obviously, the design of the reboiler must take into account a very low loss of pressure on the process side (0.05-0.1 bar). Since re-entry to the column is at a lower level, natural circulation occurs in a simple way, and does not require a large surge on the reboiler feed. Usually the draw off tray, and especially that which feeds the intermediate reboiler, is empty. The regulation required to ensure correct functioning is based on controlling the two temperatures of the intermediate sidedraw and the bottom product. The latter obviously determines the level of stabilization and usually governs the flow rate of the heating fluid. The other regulation is apparently less important, but in all stabilization columns the side

reboiler must also be controlled (through by-passing part of the hot product). Excessive heating would lead to a considerable increase in the flow rates of liquid and vapour over the sidedraw tray, which is almost always the most loaded and therefore the most critical in the entire column. The type of trays and/or packing used under various conditions will be described below. As far as the materials used are concerned, the observations made earlier remain valid, taking into account that in this case the corrosiveness resulting from the presence of hydrogen sulphide is increased by the higher relative temperature. All parts of the unit which come into contact with gas saturated in water must be made of stainless steel AISI 304 or 316-L. Demercaptanization This fairly sophisticated process, with characteristics typical of the refining industry, is also becoming established in field treatments. When an oil with a high H2S content is produced, it frequently also contains mercaptans, and the sweetening process described above has only a minimal impact on their presence in the stabilized crude. If the methyl and ethyl mercaptan content after stabilization falls, as is often the case, within the limit of 60 ppm in weight, and sometimes higher, the removal of these compounds is left to the refinery. This solution is usually adopted when the producer of the crude and the refiner belong to the same company, and when there are no specific transportation restrictions. In many cases (for example the oils produced in the Caspian Sea area) the light mercaptan content is extremely high (1,500-2,000 ppm in weight in the stabilized crude); in order to handle this oil safely, mercaptan must therefore be neutralized by transformation into disulphides. The most commonly used process is that known as Merox, developed by UOP (Universal Oil Products). The description of this unit is limited to identifying its impact on the overall crude treatment system. The removal of H2S has only an extremely modest impact on the removal of mercaptans. However, the sweetening process must be as effective as possible, since even modest traces of residual H2S in the oil cause significant problems in the subsequent process of mercaptan neutralization. The term neutralization is used since in this case no extractive process takes place, although it is available, but merely neutralization based on transformation into disulphides. First, it should be remembered that the volatility of methyl mercaptan falls in between that of normal butane and isopentane, and that of ethyl mercaptan

666

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

gas recovery

gas to treating and selling stabilized, sweetened and mercaptanes free oil

gas reinjection dehydration and degasolination

gas to reinjection system

flash gas compressor

washing water

I stage separator reservoir fluid II stage separator

acid oily water treatment

Fig. 12. Three-phase separation with stabilization, sweetening and demercaptanization.

between isopentane and normal pentane. As a consequence, these compounds are concentrated in a very light and relatively restricted cut within the crude. Thus, upstream of neutralization, a crude oil pretreatment is necessary; this involves fractionating the light cut on which the neutralization mentioned above is to be carried out. This fractionation column (light gasoline splitter) is inserted downstream a stabilizer, and is therefore fed by the bottom product of the latter. The two columns are thus thermally integrated to reduce energy consumption to a minimum (Fig. 12). The splitter is a fractionation column with purity requirements to be respected on the bottom product, obviously referring to the content of the two contaminants (mercaptans and H2S). In a mediumlight crude, the top cut to be sent for treatment does not exceed 12-15% by volume, whereas for heavier crudes it may be much lower. To obtain a feed to the plant which is as small and light as possible, the column must have a considerable reflux at the top. Lightness is essential for the liquid-liquid extraction

on which this process is based. This solution, other operating parameters being equal, makes it easier to obtain the bottom specification whilst keeping the temperature of the reboiler at a lower level. Another parameter which heavily influences the size of the splitter is the working pressure. For the reason described above, this must be as low as possible. Obviously the light top product has a bubble pressure value which is far higher than atmospheric pressure at the temperature obtainable, for example, with an air condenser. In the example under consideration, the main parameters are as follows: working pressure of the condenser 2.5 absolute bar; working pressure of the reboiler 3 absolute bar; temperature on exit from the total condenser outlet 38C (bubble point of the top cut); temperature at the bottom of the column 210C. As can be seen from the simplified diagram (see again Fig. 12), the tray column (splitter) is fed in the middle section with the partially evaporated feed. The number of theoretical stages needed for a good fractionation is four above the feed and six below the

stabilizer

side separator

light gasoline splitter

caustic washing

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

667

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

Fig. 13. Demercaptanization of non-extractive

light gasoline (Merox process, UOP). product filter mercaptanes free gasoline prewash with H2S removal liquid-liquid extractor exhaust air regenerator oxidizer air air and disulfide separator gasoline

feed, giving a total of ten; to these we should add the top condenser and the bottom reboiler. In many cases, the reference point used to determine the degree of removal/neutralization of mercaptans is the value that can be guaranteed with a caustic wash of the light gasoline stream (expected value: 10 ppm; guaranteed value: 15-20), adapting the splitter to this specification. In this way a final value three or four times lower than that allowed by the highly restrictive GOST can be obtained. It should also be remembered that washing with a sodium hydroxide solution represents a complete purification treatment for the stabilized product. The latter, after treatment, will have a negligible H2S content. Taking into consideration the combined effect of the two contaminants (mercaptans and H2S) and the elimination of the most aggressive and lethal of these (H2S), it is unnecessary to overload the treatment excessively, reducing the residual methyl and ethyl mercaptan content in the bottom of the splitter to 10 ppm. By relaxing this specification to 60 ppm, a further reduction in the whole fractionation system is obtained, with a lowering of the bottom temperature; a lighter and smaller feed to the caustic wash is also obtained. By combining this specification with the optimization of column pressure drops, we can achieve bottom temperature levels which make it far simpler and more reliable to use a classic reboiler (of the once through type), also used in the stabilizer, with a slightly higher pressure level for the heating steam. The product at the bottom of the splitter is then exchanged with the bottom (partial) reboiler of the stabilizer, and subsequently with its side reboiler (see again Fig. 12).

The transformation of volatile mercaptans into disulphides is based on a two-stage oxidation process: RSHNaOH NaSRH2 O 2RSHO22R'SSR 2H2 O Both reactions involve the production of water, which tends to dilute the washing solution. The first reaction occurs in the washing column with the caustic solution, and the second refers to the oxidation by air of the light mercaptans to form disulphides; these simultaneously regenerate the sodium hydroxide, making it possible to separate out the disulphides produced. It is obvious that in regeneration the oxidization and separation of the disulphide produced shift the equilibrium of the first reaction on the opposite side to that which occurs in the washing column. Fig. 13 shows a diagram of this process. The top product from the splitter is sent to a liquid-liquid washing column by its reflux pump (see again Fig. 12). This column must operate at a pressure of at least 1.5 bar above the bubble point of the light cut to be treated at the highest point, where the treated stream is extracted. For this reason, the light gasoline, which also contains LPG, must be fed into the unit at fairly high pressure (taking into consideration that the washing column is about 30 m high). The optimal working temperature for washing is about 38C, and in this case is identical to that of the splitters top condenser. The column creates a contact between the two streams (caustic solution and light gasoline): the lighter stream from below, the heavier washing solution (sodium hydroxide at about 13% in weight) from above. Both the washing reaction with the sodium hydroxide and that of

668

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

mercaptan neutralization are slightly exothermic, but the rise in temperature in the two sections is limited by the considerable dilution with water of the reagents and the heat capacity of the solution. On the bottom of the extractor the rich solution is collected; at the top the treated hydrocarbons are separated. The solution is regenerated by oxidization with air in the presence of the catalyst in the liquid phase. The disulphides must be separated out in a purpose-built separator included in the oxidizer; the latter also separates out the exhaust air, which must be disposed of with an incinerator. The separated disulphides can be extracted, but in the case under examination they are remixed with the oil together with the treated hydrocarbon stream. The gasoline treated in the Merox is a light gasoline and LPG cut desulphurized to commercial specifications; thus, it can be commercialized separately at little cost, by fractionating the top product of the extractor. As far as the choice of materials is concerned, the same considerations outlined above for sweetening apply, taking into account that in this case corrosiveness is especially accentuated in the regeneration of the caustic solution with air.

5.3.5 Process equipment


Separators

The main function of this equipment was analysed both in the discussion of multiflash stabilization, where gas is separated from oil, and in the separation of water, where the properties of the most commonly used internals were noted. When the first stage separator receives extremely large quantities of gas, and the two-phase flow of the gathering system causes the condition of liquid slug flow followed by gas cushions, it is essential to dampen the kinetic energy of the liquid propelled by the gas by inserting a suitable momentum absorber in front of the fluid intake nozzle. This device may take a variety of forms, and in addition to absorbing the kinetic energy of the incoming fluid must also produce, as far as possible, an initial rough separation, causing the liquid phase to fall towards the bottom of the separator. This separator is designed in such a way as not only to guarantee the mean holdup time required to separate the phases, but also dampen the discontinuities created by the variable flow described above. In this case the vessel contains only a vortex breaker on the liquid outflow, and a demister on the

gas outflow. Often there is a system of nozzles on the bottom for the washing and removal of sediments (desander). Fig. 6 shows a typical three-phase separator for the intermediate stage which, in addition to straightening vanes (one or more sections depending on the greater or lesser difficulty of separation and coalescence), also contains a sieve tray for distribution. If the velocity inside the separator is very low, in the order of 1m/min, the distribution of flow may be insufficient. The use of one or more dividers (such as that shown in Fig. 6), is highly recommended for separators installed on floating production platforms (FPSO, Floating Production Storage & Offloading), subject to pitching and rolling which destabilize the levels and lead to mixing rather than separation. To discharge water, the system shows the use of a separation barrier which makes it possible to keep the outflows for both oil and water on the bottom. For sizing and the holdup time of the liquid required to separate the phases, see the example given earlier; for the gas phase see chapter 5.4. In oil separators, with the exception of some limited cases, gas rarely influences sizing, not even in the first stage of separation when the flow rate may be relatively high. Gravity separation influences particles with dimensions in the order of 100 mm; in the gas phase the coalescer installed on the outflow lowers this to a size of 10 mm or even 3 mm, with small crosssectional areas of flow and therefore modest dimensions. The velocity of the gas phase in the separator may be extremely high, and the crosssectional area (usually 50% of the total section) is therefore sufficient, and does not require specific evaluation. Vertical separators are most commonly used for the final atmospheric stage. If it is not necessary to separate out water, which can be done more effectively with horizontal separators, this equipment has the sole purpose of completely degassing the oil, avoiding the formation of vapours inside the stock tank. When the latter is of the floating roof type, this solution is mandatory for the integrity of the tank. The vertical separator (gas boot) is installed on an independent structure next to the tank, in order to fill it by gravity avoiding the use of pumps. The twophase stream from the preceding separator is sent into the top of the gas boot, where the gas is separated out. A distribution tray feeds the section beneath, which consists of a random packing of suitable height (1-1.5 m). From the bottom of the gas boot a line feeds the bottom of the stock tank, allowing it to be filled (Fig. 14).

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

669

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

Heat exchangers

Traditional shell and tubes exchangers do not differ in any way from those used in refineries. The thermodynamic calculation methods are obviously identical, and are available for process checks in all the most commonly used process simulation programmes. Due to their ease of sizing, these types of exchangers are used very often for field processes, even when it would be better to use different typologies; a plate heat exchanger is more difficult for non-specialist personnel to evaluate, and for this reason shell and tube exchangers only are often chosen in the initial phases of a project. When space and weight represent fundamental constraints, as in offshore processes, it is essential to use more compact solutions. Since the treated fluid has a very high fouling factor (0.0004 hm2 C/kal), floating head exchangers must be used; these become very large therefore the tube bundle must be removed. Alternative solutions, which are often much more expensive, reduce all of these problems definitively. When the oil is extremely viscous, the exchange surface of a shell and tube exchanger is very large. In these cases, the use of a plate exchanger is ideal. The latter is not subject to the limitation of crossover on outlet temperatures (the outlet temperatures on both sides of the exchanger are identical), and it can thus be used in pure counterflow.
Columns and their internals

By far the most commonly used equipment for the treatment of oil is the separator. However, the use of columns to treat crude oil is increasingly frequent. Tray columns are identical inside to those used in refineries. The trays most commonly used are valve type, due to their enormous operational flexibility
atmospheric gas

(turn down ratio up to 22-25% without compromising efficiency). Since they must guarantee mass transfer with large flow rate of liquid with respect to the vapours concerned, the trays used are always almost of the double split flow type; for very large columns fourpasses flow trays are also used. The use of random packing is rare in field treatment. The increasing use of FPSOs for the development of offshore fields has led to an almost exclusive use of separators and often of electrostatic desalters instead of columns. A tray column cannot work even with modest oscillations of the vessel (magnitudes in the order of 1o). By contrast, structured packing instead of trays is extremely useful due to its adaptability to oscillations. There are numerous applications for this type of internals, which have become more frequently used for offshore gas field treatments than for those in oil fields. With respect to the conventional valve tray, packing in general, and especially the structured type, offers the advantage of being more compact (reduction in both diameter and height for an equal number of equilibrium stages). This typology is not very flexible; it gives a far more modest turn down ratio than valve trays, and requires an extremely efficient liquid and gas distribution system, which is usually more sophisticated and delicate than a comparable system in a tray column. Structured packing in particular needs to be mounted with great precision; since this must be done in situ, it requires trained personnel and appropriate controls. Almost all columns have a demister element in their upper part, above the top tray and the distributors. In a tray column, the most frequently used system is the classic full-section wire mesh pad. A demisting unit with vanes (see again Fig. 6)

gas boot

M PDM positive displacement meter floating roof storage tank

oil to pipeline

oil from intermediate separator

transfer pump

booster pump
Fig. 14. Atmospheric separation, storage, pumping and measuring.

670

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

can also give excellent results and, since it requires less maintenance, is preferable for all offshore applications. As far as materials are concerned, all the internals described above must be made of stainless steel, even in the absence of corrosive components. Using this material allows for very reduced thicknesses, and thus a reduction in weight. Note that, in contrast to other sectors of the chemical and petrochemical industry, internals in plastic are not used.
Furnaces

In desalting and stabilization, furnaces are often used as direct heaters for the fluid to be treated, and as reboilers. When a furnace is used as a heater or reboiler, it is preferable to use the typology known as a pipe still or horizontal box, with burners on the bottom (Fig. 15). The burners on the bottom, unsuitable when we wish to use liquid fuel, are suited to the use of gas, since they facilitate a good distribution of the heat flow. The forced draft horizontal box is preferable to the vertical pipe still, which is far more widespread for other applications and takes up less room. The main reason for this choice, alongside the more homogeneous distribution of heat flow, is the frequent need to mechanically clean the coils, easier with horizontal pipes. The formation water emulsified in the oil causes significant scale in contact with the radiant wall of the furnace. For mechanical cleaning operations, there are flanged heads or fast-closing heads outside the combustion chamber. The density of heat flow, even in furnaces used as reboilers, must be heavily reduced in order to lower the wall temperature to the minimum. For particularly viscous oils, direct heating should be avoided and in any case a mean flow density of 8,000-9,000 Btu/hft2 or 21,700-24,400 kcal/hm2 should not be exceeded. Taking into account that the surface receiving the radiant heat is just over 50% of the total, and that the exchange coefficient over the inner wall is extremely low, it is obvious that the temperature of the latter may be excessively high even if the temperature of the heated fluid is low. To reduce this phenomenon, the fluid velocities should be kept extremely high, accepting the consequent loss of pressure. For furnaces with low flow rates of oil and low capacity, a single coil should be used to avoid problems linked to poor flow distribution, especially when working with low flow rates. Where it is necessary to use two or more coils in parallel, their distribution must be perfect. The burner should therefore not be fed with a two-phase vapourliquid produced in the heat exchanger sometimes found upstream the furnace. The furnace and the

exchangers must be fed by a pump, and the intake pressure must be above the bubble point of the crude at intake temperature. Inside the coil, pressure drop and heating may cause evaporation to take place, easing exchange by keeping the wall temperature low. When the furnace is used as a reboiler, circulation is always maintained with pumps; this resolves the problem of distribution in the coils. However, maintaining a high velocity (2-2.5 m/s on entry) may be helpful to avoid problems of high temperatures at the wall resulting from low flow rates. The percentage vaporization of the feed does not significantly exceed 10% in weight, and therefore does not cause problems, contributing to a significant improvement in the internal exchange coefficient by keeping the wall temperature relatively low. In this case, higher heat flow densities (10,000 Btu/hft2) are acceptable. Taking into consideration that this application (using a furnace as a reboiler) is linked to stabilization, in other words with fluid temperatures of 150-160C, it is easy to understand why the wall temperature must be kept as low as possible in this case too.

burners

A-A section B-B section

convection

radiant A A B B

burner vertical pipe still


Fig. 15. Types of furnaces.

burner horizontal box

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

671

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

In the case of two-phase flow, in addition to being easier to clean internally, horizontal flow causes fewer vibrations and problems than vertical flow. In many cases it is considered more appropriate to use all convection furnaces. The advantage of this solution is a greater uniformity in the distribution of heat over the exchange surface. For reasons of compactness, smaller coils are used in these furnaces, causing greater problems with the distribution of flow and greater difficulty in guaranteeing mechanical cleaning. Furthermore, the temperature at which hot gases enter the convection package is very high (above 700C); as a result this solution is far from ideal. The need to reduce the exchange surface, and consequently its dimensions and temperature control, leads to the use of large fans to recirculate fumes, needed to control their temperature. Taking this into consideration, this solution does not present significant advantages over the conventional radiant furnace provided with a convective zone (crossover). The latter has a critical point which must be carefully controlled with temperature measurements at the surface of each coil (skin points); this is the crossover, where there is a transition from radiant to convective heat flow. Under this condition, the two heat flows, with the convective heat under conditions of maximum fume temperature, are cumulative. A similar situation also occurs in the first rows of coils in an all convection-type furnace (excessive flow density, linked to the combined effect of high velocity and the high temperature of the fumes).
Heat recovery and steam generation units

burners, the associated refractors and fans to recycle flue gasses; furthermore they operate at lower temperatures (below 500C), yet still sufficiently high to ensure considerable compactness. If they are used on offshore platforms, recovery units for fumes resolve the problem of safety distances since no postcombustion units are needed. Finally, it is worth noting that starting up a furnace, even a small one like those used in field treatment, requires a relatively long time compared to that needed for a heat recovery unit. When the process temperature is relatively low, pressurized water can be used as an intermediate fluid; it is glycolated if the plant is used in cold zones. With this system, a fluid with a temperature of 165C can be obtained, more than enough to meet most of the heating needs described above. When the treatment scheme is unusually sophisticated, such as one which includes the demercaptanization of the crude, even the use of steam and especially back pressure steam is advisable. This type of energy recovery is extremely widespread in refining, but very limited in field treatments. It is very unusual to couple a steam boiler with a superheater to produce electrical energy with a back pressure turbine and to use the steam at medium and/or low pressure for heating, for example in desalting and stabilization. The main reason for this is that steam, and still more the boiler feed water relevant to its generation, is considered something to be avoided. However, today the need to reduce environmental damage is leading to the adoption of solutions with lower energy consumption.
Pumps

When dealing with large quantities of hydrocarbons, which were cheap in the recent past, it is easy to understand why special attention has not been paid to energy saving in production activities, and consequently to the recovery of energy, and heat in particular. Even today, in many non-optimal logistical situations more weight is given to the simplicity of plants than to energy recovery. The steam generator or hot oil boiler which produces the required heat can often be installed to recover heat from the flue gasses of the gas turbines. These are used as motors for the compression of associated gas and are needed to produce the electric energy required for both treatments and the infrastructure of an oil field. Since these are usually small machines compared to those in a mixed cycle electrical power station, they are specially suited to fast field installation. They are certainly more compact and lighter than furnaces of all convection type because they do not have a combustion chamber,

The pumps used in the oil industry are API standard, and therefore very expensive, although they are not made from particularly valuable materials. Due to its filming effect, crude oil normally protects the materials with which it comes into contact from corrosion, and for this reason the pumps used for oil are always made of normal carbon steel. Except under unusual circumstances, these are almost exclusively centrifuge pumps (single stage for low differential heads, multistage for high heads). Often the fluid to be pumped is at boiling point, and it is therefore important to use pumps with a low NPSH (Net Positive Suction Head), and to install them with special care. In many cases, when we need to pump from atmospheric stock tanks, vertical (barrel) single stage pumps are used. If the head required to deliver the oil is high, the pump described above (booster pump) is followed by a multistage transfer pump (see again Fig. 14).

672

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

Measurement systems

The treated oil is transported by tanker or oil pipeline. In both cases, it is common practice to place a fiscal measurement system on the loading line to the tanker or the onshore pipeline, to quantify production as accurately as possible. In order to do this, suitable instruments must be chosen and installed with care. The best installation point is downstream the booster pump, in other words at medium-low pressure (5-7 bar), but sufficient to ensure that the measurement is single-phase (see again Fig. 14). The measurement is also corrected for temperature. The meter should not be installed on a direct connection to the oil pipeline following a multistage pump; since this is expensive equipment, based on several lines operating in parallel to guarantee maximum precision at low flow rates (arrangeability), it is not economically viable to operate at high pressure. Installation between the booster pump and the transfer pump is therefore preferable. The most commonly used instrument is the PDM (Positive Displacement Meter), which, unlike a calibrated flange, measures the total flow rate of the product directly. In order to calibrate it correctly, and check its reliability on a periodic basis, installation is completed with a suitable meter prover.
Storage

Downstream the production process, the crude must be stored. There is a difference between the atmospheric tank installed at the treatment centre, and the tank farm at the crude oil loading terminal when, as is often the case, the oil is transported by sea. The difference lies basically in storage capacity. In the first case, it is generally equivalent to a days total production, whereas in the loading terminal the holdup is much longer in order to adapt to the movements of tankers. In the latter case, therefore, the average
Fig. 16. Filling losses of light products 0.5

storage time is over a week, and this entails the choice of tanks of suitable typology and size. In the past, these two types of tank were of different typology, not just as far as size was concerned, but also due to their different purpose. Assume that we have six oil fields, each having 50,000 bbl/d capacity of stabilized crude in different places. Their respective flow rates are sent to a single gathering and loading terminal on the coast, with a total reception capacity of 300,000 bbl/d. It will therefore be able to fill a million-barrel oil tanker roughly every three days. The minimum storage capacity of this terminal must be above this theoretical value in order to plan the arrival of tankers in an optimal way. If the terminal is designed for a capacity equivalent to ten days production, or three million barrels, the storage facility will consist of six large floating roof tanks, each with a capacity of about 80,000 m3. Every individual flow station with a capacity of 50,000 bpd located near the field, however, will have only a single tank with a capacity of 8,000 m3, or ten times less. The latter finds itself in different conditions from those in the terminal, since it is directly linked to the production facilites and subsequent treatment. Until a few years ago, this was improperly considered the final stage of oil treatment, as is still the case in numerous older oil fields. As a consequence the construction typology was different: cone roof tanks able to withstand a slight internal positive pressure (5 mbar). In some cases these tanks were the last stage of atmospheric degassing of the crude and separation of water. Their construction methodology therefore often allowed the tank to withstand greater internal positive pressures (50 mbar) with a rounded roof. Stock tanks are no longer used as the final degassing stage, and today even these small tanks are built with the floating roof technique. This solution allows the empty space between the surface of the liquid and the

in fixed roof tanks (GPSA).


0.4

working pressure, psig (atmospheric pressure) locus of maximum losses

filling loss (% of liquid pumped in)

0.3
1

0.2

2.5 5

0.1
20

10 15 25

0.0

10 11 12 13 14 15

vapour pressure at liquid temperature (psia)

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

673

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

roof to be reduced to a minimum; by minimizing this volume, losses due to breathing and the filling of the tank are also reduced. It is obvious that these losses alter the properties of the crude, since they consist of the lightest compounds. The diagram provided (Fig. 16) shows a simple correlation between losses due to the filling of the tank, and the true vapour pressure of the stored oil. The working pressures given are considerable, because this diagram refers to a cone roof tank. With a floating roof tank, losses are far smaller, but not nil. So-called breathing losses, on the other hand, are due to the temperature difference between day and night, and the variations in insolation to which both the tank and especially the layer of vapour between the roof and the liquid free surface are subjected. During the night the tank breathes, in other words it draws air inside which becomes saturated in hydrocarbon vapours; during the day these are released as the tank warms up, and therefore the vapour volume increases. Since the tank cannot withstand even modest pressures, either negative or positive, it is provided with a system of calibrated relief valves, placed so as to ensure that pressure is kept under control at negligeable figures. The decanting of water and the residual water content (maximum 0.5% in volume) was described above. Given the time the crude is held in the storage tank, this may facilitate some decanting of the residual water content. This does not represent a problem, and is generally resolved by placing a series of drainage valves on the bottom rim, helping to discharge the decanted water. In fact, during the storage of the crude, and especially in the final stock tanks where holdup times are very long, the main objective is not to facilitate decantation, but, by contrast, to ensure the homogeneity of the product. During stabilization with multiflash or with a stabilization column, operating temperatures are such that, except in extraordinary cases, the oil never reaches its cloud point; as a result, in process equipment working conditions do not cause the separation of paraffin. Taking into consideration the fact that during treatment the oil contains dissolved gas, the solubility of the paraffins under these conditions is higher, and it is therefore more probable that we have a non-Newtonian fluid in a stock tank than in a separator. When this happens, the solid paraffins which separate spontaneously from the crude do not cause problems; although these components have a high molecular weight, they have much lower densities than the mean density of oil, and therefore tend to remain in suspension rather than settling. However, if the oil contains emulsified water, since paraffins are strongly hygroscopic, emulsions known as waxy sludges form,

mainly composed of water and thus of high density; these do settle. A layer of these products therefore forms on the bottom of the tank, causing considerable problems in storage and loading operations. Even in the presence of crudes which are not particularly corrosive, significant corrosion phenomena have been noted at the interface between the paraffin-water emulsion at the bottom, and the overlying oil phase. The stored oil, due to the phenomena described above, comes into contact with the oxygen in the air, and is absorbed differently by the two phases; on the bottom of the tank an electrochemical corrosion process takes place, partly linked to the differentiated aeration of the two phases, and their different electric conductivity. In order to avoid the accumulation of waxy sludges on the bottom, various solutions can be adopted: the simplest is to use agitators inserted on the bottom rim of the tanks to accentuate the natural convective motions present in the stored mass, helping to reduce separation and consequent settling. These systems generally use high velocity jets injected into the oil mass. They are more effective if the crude injected is also heated. In many cases a genuine treatment process is carried out, involving the heating, melting of the wax and separation of water from the emulsions extracted from the bottom, and, obviously, the treatment of waste waters. The material used to build stock tanks, for the reasons outlined above, is exclusively carbon steel, regardless of their size and construction typology. Since water may settle on the bottom of the tanks, with an accumulation of dissolved corrosive components (both acid gases and fatty acids linked to the oxidation of the crude oil), it is advisable to use protective epoxy linings.
Vapour recovery plants

The vapours released into the atmosphere during the normal transportation of gasolines shall be recovered according to standards using suitable plants, based on refrigeration and/or absorption by the gasoline itself. Active carbon filters are also used as a final purification stage. A similar procedure should also be used during the transportation of crude oil, for the following reasons: the vapour pressure of oil is high (often higher than that of gasoline) and the concentration of evaporated hydrocarbons released into the atmosphere is higher; the compounds are similar, and like all non-combusted hydrocarbons are harmful to human health; since gasoline is a component of oil, the quantities involved in the latter are greater, and far more concentrated in restricted locations (for example loading and offloading terminals for oil tankers). It is important to stress that the use of vapour recovery systems during the transportation of gasoline

674

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

has an economic return which for large storage facilities and refineries compensates for investments. It is therefore to be hoped that this solution will also be adopted in the transportation of crude oil, as is already the case in some Northern European countries, especially Norway.

5.3.6 Secondary recovery systems


Water injection

Water injection is carried out in oil fields, and is used as a secondary recovery technique to maintain high levels of production, and to increase the final recovery of oil from the reservoir. This practice is carried out by pumping water into productive levels, following the indications obtained from reservoir studies, both as concerns the wells to be used for injection (number and location), and as concerns the flow rates and pressures to be adopted. This section will examine the treatments to which the water must be subjected before injection, to ensure its compatibility with both the reservoir fluids and reservoir rock. Any substance suspended in the injection water or having formed subsequently (due to the precipitation of salts, the flocculation of dispersed clays and the formation of gelatinous masses made up of colonies of bacteria) would obstruct the pores of the productive level, causing considerable damage to the reservoir. The water to be injected may come from the reservoir (formation water) or from other aquifer levels above or below the productive one, from rivers or from the sea if the reservoir is offshore or near the coast. The first and most important treatment which the injection water must undergo is filtration. Water almost always contains solids in suspension, made up of particles, even small ones, which must be removed. The degree of filtration required may vary. The most widespread standard is based on the use of dual media filters, suited to the removal of suspended particles larger than 3 mm. To protect the filters, depending on individual circumstances, other degrees of coarser filtration may be used to remove suspended solids, which depend on the origin of the water to be injected. If the injection water is withdrawn from subsurface levels at shallow depth (water with low salinity and free of oxygen), it does not require deoxygenation and filtration is limited to the final stage. By contrast, when surface water is to be injected, an extreme deoxygenation of the water must be carried out downstream the filtration described above. To avoid damaging the reservoirs equilibrium and encouraging the formation of bacteria, the water injected must be completely devoid of oxygen. The

residual oxygen content, after degassing or other types of deoxygenation, must be neutralized with appropriate compounds (oxygen scavengers) such as hydrazine and others. A very deep deareation, however, reduces the amount of oxygen scavenger used. The optimization of treatment thus depends on a good compromise between the investment needed for degassing, and the resulting reduction in operating costs for the use of chemicals. Vacuum degassing is the solution most commonly used today. The water is sent into the top of a column operating under vacuum, and distributed adequately over a bed which must create a large surface area per unit volume inside the column itself. Due to the vacuum, the oxygen content which at saturation and low temperature is 5 ppm wt., is reduced proportionally to the reduction of pressure, maintained inside the column by using an appropriate system. The vacuum most commonly used for this type of application is in the order of 40 tor (40 mm of Hg or 0.054 absolute bar). To produce the vacuum required, various systems can be used, including a liquid ring pump or a rotating screw compressor. Also frequently used is a combined system consisting of a liquid ring pump and an ejector which uses the air itself as a driving fluid to produce the first stage with a very low suction pressure. Sometimes classic vacuum units with multistage steam ejectors can be used. Electrical energy consumption for the system using a liquid ring pump or compressor is extremely modest since, even for extremely large amounts of water, the amount of air to be extracted is very small; the consumption of steam if the latter is used as driving fluid in a group of ejectors is also very low. The most frequently used installation for this purpose is a two-stage ejector with intermediate condenser. The flow rate of injection water is usually in the same order of magnitude as the oil production. The example provided below shows the orders of magnitude of the main operating parameters concerned. Assume an oil production of 200,000 bbl/d, and a corresponding water injection maximum flow rate of about 400,000 bbl/d, or 2,650 m3/h. The oxygen to be extracted is therefore in the order of 13.2 kg/h, to which we should also add nitrogen (about double), equal to 26.4 kg/h. Alongside the noncondensible gases, saturation water vapour is also extracted under suction pressure and temperature conditions, which in this case is assumed to be 20C. From thermodynamic tables for water, we can obtain its vapour pressure, which is 0.023 absolute bar. The concentration of water vapour in the total gas in equilibrium is proportional to its partial pressure (Daltons law), and therefore 0.023/0.05443.3% mol. The gas extracted

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

675

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

from the column therefore contains 13.2/3226.4/281.355 kmol/h of noncondensible gas; the total gas extracted is thus 1.355/(10.433) 2.39 kmol/h, or 58.23 kg/h. The amount of air entering the system due to defective seals should be added to this value. The concentration of oxygen in the vapours extracted is 0.412/2.390.172, not very different from that found in air. According to Henrys law, a reduction in pressure thus causes a proportional reduction in the residual oxygen concentration in the water which at equilibrium is about 5% of its initial value, in other words 250 ppb. Actually, the water is usually withdrawn at depth (20-30 m below the surface), and thus has a lower content than that assumed above. By reducing the pressure to 0.034 bar and repeating the simplified calculation above, it can be seen that the vapour pressure of the water acquires a predominant importance, bringing the concentration of the latter in the gas mixture extracted to very high values (68%). The removal of oxygen and the amount of gas extracted increase more than proportionally to the reduction of pressure. The residual dissolved oxygen concentration becomes 85 ppb. The example shown for a temperature of 20C makes it clear that this parameter has a significant impact on the performance of a vacuum column. If the temperature is 30C, the vapour pressure rises to 0.042 absolute bar; with a working pressure in the column of 0.054 absolute bar, the concentration of saturation water is 77.8% of that of the air extracted; the corresponding oxygen concentration in the vapours decreases to 6.8% and the concentration of oxygen in equilibrium in the water is about 90 ppb. A total of 125 kg/h of gas is extracted. This analysis demonstrates that the vacuum to be used is strictly connected to the temperature of the water. When the water is hotter, a less intense vacuum must be used. With these equilibrium values (85 and 90 ppb), a final oxygen content of 100 ppb can be obtained, and the amount of hydrazine or other oxygen scavenger needed for complete neutralization is negligible. The hydrazine or other reducing agent is injected in a diluted solution through the water itself. This is done by mixing the product in the recycling line between the booster pump and the bottom of the vacuum column from which it takes suction. In this way the reducing agent has a holdup time of a few minutes, needed to complete the reduction reaction. Whereas the vacuum unit is not the most important, and especially most bulky part, the reverse is true of columns. The large flow rate of water, even using appropriate internals, requires several large lines running in parallel. This is also true for pumps and

filters. With the flow rates described above, several pumps working in parallel are needed, driven by gas turbines. Usually the injection pressure is below the values used for gas, but is still frequently in the order of 120-150 bar. The injection system uses multistage centrifuge pumps of large capacity and enormous power, which are not suited to suctioning directly from a vessel operated at 0.05 absolute bar. As a consequence, booster pumps (centrifuge, single-stage, barrel type) must be inserted before the injection pumps to ensure that the latter have the correct suction pressure (4-5 absolute bar). Assuming 120 bar as the required injection pressure, the total power installed is slightly above 12 MW; at least four injection lines are therefore needed, each with a flow rate of 700 m3/h, and consisting of a booster pump with a power of 150 kW and an injection pump of about 3 MW. The booster pumps are driven by electrical motors, whereas the multistage pumps are coupled directly to a gas turbine. The water pumping system, like that for the injection of gas described in the following section, does not have spare machines. The subdivision over several parallel lines guarantees the partial availability of the system in the event of one machine failing. An alternative, more compact but more expensive, deoxygenation system than vacuum degassing is based on hydrogenation with a platinum catalyst, allowing all the dissolved oxygen to be transformed into water. In many offshore platforms, to protect the water intake systems and their pumps, electrolytic cells are used for the production of chlorine by the hydrolysis of sea water. In this process, hydrogen is also produced; this led to the idea of treating water containing dissolved oxygen in a catalytic reactor allowing the inverse reaction to the hydrolysis described above to take place. On the basis of the data provided in the preceding example, the stoichiometric flow rate of hydrogen is relatively modest (1.65 Kg/h). Despite this, catalytic deoxygenation is still more expensive than traditional vacuum deoxygenation, although it does have a considerable advantage linked to its greater compactness. Finally, it is important to remember that water analyses may sometimes show a possible supersaturation in salts, such as sulphates, under reservoir conditions; in this case other additives must be injected to prevent their deposition. Units may also be installed for the sulphate removal of the water based on the use of semipermeable membranes. Any formation of gelatinous masses caused by bacteria is prevented with the injection of a biocide, identical to those used for water intakes. The materials most frequently used for the injection of water in offshore platforms are: carbon steel protected with an epoxy lining for all vessels

676

ENCYCLOPAEDIA OF HYDROCARBONS

TREATMENT PLANTS FOR OIL PRODUCTION

TEG contactor

I stage gas compressor associated gas

II stage gas compressor

to the injection well

TEG reg.
Fig. 17. Gas injection with TEG dehydration.

under low pressure preceding degassing, including filters and columns, and duplex steel for booster and injection pumps. When ground water is injected, it is sufficient to use carbon steel.
Gas injection

The injection of gas is an alternative secondary recovery technique to the injection of water described above. The decision to use one or other of these techniques, or both simultaneously, may be dictated by the configuration of the reservoir and its petrophysical properties, as well as by the type of oil it contains. Often this choice is linked to the availability of the fluid to be injected. The decision to inject gas is often taken in oilproducing fields when there is a large availability of gas associated with the production of oil. The gas to be injected into the reservoir may come from the reservoir itself, either, as in the case mentioned above, as gas associated with the oil, or as gas produced separately from a level above or below, but not in hydraulic communication with the productive oil level. The gas used for injection may also come from other reservoirs near the production reservoir, either as associated gas or otherwise. Injection is carried out into the top of the oil reservoir through injection wells, whose location, number, flow rate and injection pressures are established on the basis of the relevant reservoir studies. The gas to be injected, recovered as associated gas, is usually conditioned. The reinjection pressure is a fundamental parameter in determining the compression and treatment scheme linked to it. Often the operating pressures range from 250 to 400 bar. In some cases, pressures of 800 bar may be reached. In all cases, however, the distribution network at the wells operates in the dense phase. Under these conditions, the fluid

injected, though described as a gas, has physical properties such as density which make it far more similar to a liquid. It makes little sense to speak of phase separation for a mixture of hydrocarbons at these pressure levels and at ambient temperature; rather, it is important to avoid the separation of a water phase which under these conditions could cause hydrates to form, even at relatively high temperatures. If associated gas is used, this is collected at the pressure of the first separation stage, and subsequently compressed through a series of compression stages, with intermediate cooling and separation (Fig. 17). Assume a pressure of 20 absolute bar and a temperature of 30C as suction conditions, and a pressure of 250 absolute bar as the reinjection condition. As is known, the maximum compression ratio per stage is linked to the delivery temperature, which must not exceed values which might cause mechanical problems, or problems with seals and lubrication. A very conservative value is 150C. Currently, many vendors offer centrifuge compressors designed to operate at 180C for delivery, and this allows for a considerable reduction in the number of stages required. The heating of the gas during compression depends largely on its composition, and specifically on the parameter Kcp /cv where cp is the specific heat at constant pressure, and cv at constant volume. There are numerous nomograms allowing the simple calculation of the requisite power and delivery temperature for a compression stage as a function of the type of gas, that is K, of the compression ratio P 2 /P 1 where P2 and P1 are the absolute delivery and suction pressures respectively, and of the suction temperature T1. In the case under consideration, the required compression involves two stages in series. The

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

677

DEVELOPMENT PHASE OF HYDROCARBON FIELDS

compression ratio is thus about 3.55, and the intermediate pressure is 71 absolute bar, whereas the final pressure, considering a pressure drop of 0.5 bar in the final cooling, is 250.5 absolute bar. After the first stage the gas is cooled to 30C using a water exchanger followed by a vertical separator (KOD, KnockOut Drum), which removes any condensed liquids. It is then delivered to the second stage and compressed to the required pressure. Usually, since the efficiency of compression is linked to the effective volumetric flow rate, at suction into the second stage the latter is about four times lower, and efficiency is lower as a consequence. Although there is an extensive literature on this subject, it is preferable to consult the vendor to obtain the actual efficiency and determine the type of compressor most suited to obtaining the desired result. Assuming a flow rate of gas to be injected of 1.8 million standard cubic metres/day, with a density relative to air of 0.66, the power needed for each individual stage is about 4,000 kW, and the delivery temperature is slightly below 150C. It is therefore sufficient to use a machine with 9 MW of actual power to drive both compression stages simultaneously. This can be obtained by using a gas turbine with a slightly higher nominal power (10 MW). Both stages of compression are fitted directly to the shaft of the gas turbine. The latter, being a machine with a variable speed, allows for continuous control of the compressor by varying the r.p.m. (rotations per minute). In injection systems, spare machines are not generally used. Partial availability, in the event of the failure of a single machine, is thus guaranteed by distributing the injection system over two or more lines. If space is not a determining factor, as instead is the case for offshore platforms, the injection system is distributed over two compression lines, which remain two-stage, with nominal power of 5 MW turbines as drivers. A relatively low temperature has been assumed for interstage cooling, which can be obtained offshore, for example, where cooling water at an appropriate temperature is available. For second stage cooling this is not necessary, and the final temperature may be higher. A value of 60-80C is sufficient; consequently the second cooling to be carried out at high pressure is far more modest than the intermediate cooling. This can be done with air, or with water at a higher temperature than that needed for the first stage. In order to ensure that injection functions well, the gas must be dehydrated. If it is not particularly corrosive, the dew point needed for injection is that which guarantees a margin of 5C with respect to the worst transportation conditions in the distribution network to the injection wells. Both offshore and

80,000 60,000 40,000 20,000


correction

1.0 0.9 0.8

correction for gas gravity 50F 150 100 250 2 00 30 0

lb of water/106 ft3of wet gas (at 60F and 14.7 psia)

10,000 8,000 6,000 4,000 2,000 1,000 800 600 400 200 100 80 60 40 20 10 8 6 4 2 1

gas gravity 0.7 1.2 1.4 1.6 1,8 0.6 0.8 1.0 20 25 30 35 40 45 50 molecular weight

14

.7 ps 25 ia

50 100

hydrate formation line

Dashed lines are meta-stable equilibrium. Actual equilibrium is lower water content. Angle is a function of composition

60 40

40

80

120 160 200

temperature (F)

Fig. 18. Saturation water content in natural gas.

onshore it is sufficient to reach a dew point of 0C at 250 bar. In most cases, the most suitable point at which to insert the dehydration treatment is between the two compression stages, after cooling. The diagram showing the saturation in water of natural gas (Fig. 18) indicates that the saturation at 71 absolute bar and 30C is relatively low, about 40 lb/MMSCF (Million Standard Cubic Feet), whereas at 20 absolute bar at the suction into the first stage it is 110 lb/MMSCF, and thus much higher. The water content, which corresponds to a temperature of gas saturation of 0C at 250 absolute bar is 3.5 lb/MMSCF. The amount of water to be removed is therefore extremely modest, as is the efficiency of dehydration. This level of dehydration can be obtained with a very simple TEG (triethylene glycol) dehydration unit. As far as saturation in hydrocarbons is concerned, the gas can be condensate removed or reinjected as it is. Any condensate removal may allow for some additional recovery of light hydrocarbons which can

678

ENCYCLOPAEDIA OF HYDROCARBONS

2 30 00 500 0 400 1,5 800 3,0 00 00 2,0 00 5 10 ,000 ,00 0

280

TREATMENT PLANTS FOR OIL PRODUCTION

be recycled during the stabilization of the oil. The ideal pressure for this field treatment is 45-60 absolute bar, and thus slightly below the intermediate pressure in the example considered.

Bibliography
API (American Petroleum Institute) (1982) Guide for pressurerelieving and depressurizing systems, API Recommended Practice 521. Benedict M. et al. (1951) An empirical equation for thermodynamic properties of light hydrocarbons and their mixtures. Constants for twelve hydrocarbons, Chemical Engineering Progress, 47, 419-422. Brown G.G., Souders M. (1932) Fondamental design of absorbing and stripping columns for complex vapours, Industrial and Engineering Chemistry, 24, 519. Kern D.Q. (1950) Process heat transfer, New York, McGrawHill. Kremser A. (1930) Theoretical analysis of absorption process, National Petroleum News, 22, 43.

Lockart R.W., Martinelli R.C. (1949) Proposed correlation of data for isothermal two-phases, two-components flow in pipes, Chemical Engineering Progress, 45, 1. McKetta J.J., Wehe A.H. (1958) Use this chart for water content of natural gases, Hydrocarbon Processing, 37, 8. Maddox R.N. (1977) Gas and liquid sweetening, Norman (OK), Campbell Petroleum. Perry R.H., Chillton C.H. (editorial direction) (1973) Chemical engineers handbook, New York, McGraw-Hill. Reid R.D. et al. (1977) The properties of gases and liquids, New York-London, McGraw-Hill.

References
Katz D.L. et al. (1959) Handbook of natural gas engineering, New York, McGraw-Hill.

Romano Bianco
Scientific Consultant

VOLUME I / EXPLORATION, PRODUCTION AND TRANSPORT

679

Potrebbero piacerti anche