You are on page 1of 178

G19RA

Reservoir Engineering

Heriot-Watt University

Edinburgh EH14 4AS, United Kingdom


2

Produced by Heriot-Watt University, 2014

Copyright © 2014 Heriot-Watt University

All rights reserved. No part of this publication may be reproduced,


stored in a retrieval system or transmitted in any form or by any means
without express permission from the publisher.

This material is prepared to support the degree programmes in


Chemical and Petroleum Engineering.

Distributed by Heriot-Watt University


Reservoir Engineering A

©HERIOT-W ATT UNIVERSITY June 2014 v1


3

Acknowledgements
Thanks are due to the members of Heriot-Watt, School of Energy Geoscience
Infrastructure and Society who planned and generated this material.

We would like to acknowledge the assistance and contributions from colleagues


across the University and students in preparing this and support material.

©HERIOT-W ATT UNIVERSITY June 2014 v1


4

Topic 1 Introduction to Reservoir Engineering


1 INTRODUCTION
1.1 Reserves Estimation
1.2 Development Planning
1.3 Production Operations Optimisation

2 RESERVOIR ENGINEERING TECHNIQUES

3 RESERVE ESTIMATING
3.1 Definitions
3.2 Proven Reserves
3.2.1 Exercises – Reserve Definitions
3.3 Unproved Reserves
3.3.1 Probable Reserves
3.3.2 Possible Reserves
3.4 Reserve Status Categories
3.4.1 Developed:
3.4.1.1 Producing
3.4.1.2 Non-producing:
3.4.2 Undeveloped Reserves:

4 PROBABILISTIC REPRESENTATION OF RESERVES

5 VOLUME IN – PLACE CALCULATIONS


5.1 Volume of Oil and Gas in-Place
5.2 Recovery Factors

6 PHASES OF DEVELOPMENT

©HERIOT-W ATT UNIVERSITY June 2014 v1


5

LEARNING OUTCOMES

Having worked through this chapter the Student will be able to:

• Show using a block diagram the integration of reservoir engineering with other
petroleum engineering and other subjects.

• Sketch a diagram showing the probability versus recoverable reserves


indicating, proven, proven + probable and proven + probable + possible
reserves.

• Present a simple equation for volumes of oil and gas in-place & calculate the
reserves

• Draw a sketch showing the various phases of production from build up to


economic limit.

• Draw a sketch illustrating the various recovery scenarios from primary to


tertiary recovery.

©HERIOT-W ATT UNIVERSITY June 2014 v1


6

1 INTRODUCTION

With the petroleum industry’s desire to conserve and produce oil and gas more
efficiently a field of specialisation has developed called Petroleum Reservoir
Engineering. This new science which can be traced back only to the mid 1930’s has
been built up on a wealth of scientific and practical experience from field and
laboratory. In the text of Craft & Hawkins1 on Applied Reservoir Engineering it is
commented that “as early as 1928 petroleum engineers were giving serious
consideration to gas-energy relationships and recognised the need for more precise
information concerning physical conditions as they exist in wells and underground
reservoirs. Early progress in oil recovery methods made it obvious that computations
made from wellhead or surface data were generally misleading.” Dake 2, in his text
"The Practise of Reservoir Engineering", comments that “Reservoir Engineering
shares the distinction with geology in being one of the ‘underground sciences’ of the
oil industry, attempting to describe what occurs in the wide open spaces of the
reservoir between the sparse points of observation – the wells”

The reservoir engineer in the multi-disciplinary perspective of modern oil and gas field
management is located at the heart of many of the activities acting as a central co-
ordinating role in relation to receiving information processing it and passing it on to
others. This perspective presented by Dake2 is shown in the Figure below.

2
Figure 1 Reservoir Engineering in Relation to Other Activities (adapted Dake ).

Dake2 has usefully specified the distinct technical responsibilities of reservoir engineers as:

• Contributing, with the geologists and petrophysicists, to the estimation of

©HERIOT-W ATT UNIVERSITY June 2014 v1


7

hydrocarbons in place.

• Determining the fraction of discovered hydrocarbons that can be recovered.

• Attaching a time scale to the recovery.

• Day-to-day operational reservoir engineering throughout the project lifetime.

The responsibility of the first is shared with other disciplines whereas the second is
primarily the responsibility of the reservoir engineer. Attaching a time scale to
recovery is the development of a production profile and again is not an exclusive
activity. The day-to-day operational role is on going through the duration of the
project.

A project can be conveniently divided into two stages and within these the above
activities take place, the appraisal stage and the development phase. The appraisal
phase is essentially a data collection and processing phase with the one objective of
determining the ‘viability’ of a project. The development phase covers the remaining
period if the project is considered viable from the time continuous production
commences to the time the field is abandoned. Reservoir engineering activity in
various forms takes place during both of these stages.

The activities of reservoir engineering fall into the following three general categories:

(i) Reserves Estimation


(ii) Development Planning
(iii) Production Operations Optimisation

1.1 Reserves Estimation


The underground reserves of oil and gas form the oil company’s main assets.
Quantifying such reserves forms therefore a very important objective of the practising
reservoir engineer but it is also a very complex problem, for the basic data is usually
subject to widely varying interpretations and on top of that, reserves may be affected
significantly by the field development plan and operating practice. It is an on-going
activity during exploration development planning and during production. It is clearly
a key task of the appraisal phase for it is at the heart of determining project viability.

©HERIOT-W ATT UNIVERSITY June 2014 v1


8

Before any production has been obtained, the so-called ‘volumetric estimate of
reserves’ is usually made. Geological and geophysical data are combined to obtain a
range of contour maps that are integrated to allow the hydrocarbon bearing rock
volumes to be estimated. From well log petrophysical analysis, estimates of an average
porosity and water saturation can be made and when applied to the hydrocarbon rock
volume yield an estimate of stock tank oil in place (STOIIP). Since it is well known
that only a fraction of this oil may in fact be ‘recoverable’, laboratory tests on cores
may be carried out to estimate movable oil. The reserve estimate finally arrived at is
little more than an educated guess but a very important one for it determines company
policy.

The Society of Petroleum Engineers in collaboration with the World Petroleum


Congress published definitions with respect to reserves and these are now accepted
world-wide3. These definitions have been used in the summary of reserve definitions
which follow.

1.2 Development Planning


Oilfield development, particularly in the offshore environment, is a ‘front loaded’
investment. Finance has to be committed far in advance not only of income guaranteed
by the investment, but frequently also of good definitive data on the character of the
field. Much of the responsibility for this type of activity falls on the reservoir
engineers because of their appreciation for the complex character of sub-surface fluid
behaviour under various proposed development schemes.

1.3 Production Operations Optimisation


Producing fields will seldom behave as anticipated and, of course, by the very nature
of this sort of activity, the balance of forces in the reservoir rock gets severely upset by
oil and gas production. The reservoir engineer is frequently called upon to ‘explain’ a
certain aspect of well performance, such as increasing gas-oil ratio, sand and/or water
production and more importantly will be asked to propose a remedy. The actual
performance of the reservoir as compared to the various model predictions is another
ongoing perspective during this phase.

2 RESERVOIR ENGINEERING TECHNIQUES

For many problems encountered by reservoir engineers today, the advance of


computing capability is enabling reservoir engineering modelling methods
(‘simulations’) to be carried out at the engineer's desk, previously considered
impossible. These methods are still complemented by simpler spreadsheet based

©HERIOT-W ATT UNIVERSITY June 2014 v1


9

techniques and these are often used to give the engineer confidence in the output of the
simulation models.

The basis of the development of the 'model' of the reservoir are the various sources of
data. As the appraisal develops the uncertainty reduces in relation to the quality of the
forecasts predicted by the model. Building up this ‘geological’ model of the reservoir
progresses from the early interpretation of the geophysical surveys, through various
well derived data sets, which include drilling information, indirect wireline
measurements, recovered core data, recovered fluid analysis, pressure depth surveys,
to information generated during production.

3 RESERVE ESTIMATING

The Society of Petroleum Engineers SPE and World Petroleum Congress WPO agreed
classification of reserves3 provides a valuable standard by which to define reserves,
the section below is based on this classification document.

3.1 Definitions
Reserves are those quantities of petroleum which are anticipated to be
commercially recovered from known accumulations from a given date forward.

All reserve estimates involve some degree of uncertainty. The uncertainty depends
chiefly on the amount of reliable geologic and engineering data available at the time of
the estimate and the interpretation of these data. The relative degree of uncertainty
may be conveyed by placing reserves into one of two principal classifications, either
proved or unproved.

Unproved reserves are less certain to be recovered than proved reserves and may be
further sub-classified as probable and possible reserves to denote progressively
increasing uncertainty in their recoverability. Estimation of reserves is carried out
under conditions of uncertainty. The method of estimation is called deterministic if a
single best estimate of reserves is made based on known geological, engineering, and
economic data. The method of estimation is called probabilistic when the known
geological, engineering, and economic data are used to generate a range of estimates
and their associated probabilities. Identifying reserves as proved, probable, and
possible has been the most frequent classification method and gives an indication of
the probability of recovery. Because of potential differences in uncertainty, caution
should be exercised when aggregating reserves of different classifications.

©HERIOT-W ATT UNIVERSITY June 2014 v1


10

Reserves estimates will generally be revised as additional geologic or engineering data


becomes available or as economic conditions change.

Reserves may be attributed to either natural energy or improved recovery methods.


Improved recovery methods include all methods for supplementing natural energy or
altering natural forces in the reservoir to increase ultimate recovery. Examples of such
methods are pressure maintenance, gas cycling, waterflooding, thermal methods,
chemical flooding, and the use of miscible and immiscible displacement fluids. Other
improved recovery methods may be developed in the future as petroleum technology
continues to evolve.

3.2 Proven Reserves


Proven reserves are those quantities of petroleum which, by analysis of geological
and engineering data, can be estimated with reasonable certainty to be
commercially recoverable, from a given date forward, from known reservoirs
and under current economic conditions, operating methods, and government
regulations.

Proved reserves can be categorised as developed or undeveloped.

If deterministic methods are used, the term reasonable certainty is intended to express
a high degree of confidence that the quantities will be recovered. If probabilistic
methods are used, there should be at least a 90% probability that the quantities actually
recovered will equal or exceed the estimate.

Establishment of current economic conditions should include relevant historical


petroleum prices and associated costs and may involve an averaging period that is
consistent with the purpose of the reserve estimate, appropriate contract obligations,
corporate procedures, and government regulations involved in reporting these reserves.
In general, reserves are considered proved if the commercial producibility of the
reservoir is supported by actual production or formation tests. In this context, the term
proved refers to the actual quantities of petroleum reserves and not just the
productivity of the well or reservoir. In certain cases, proved reserves may be assigned
on the basis of well logs and/or core analysis that indicate the subject reservoir is
hydrocarbon bearing and is analogous to reservoirs in the same area that are producing
or have demonstrated the ability to produce on formation tests.

©HERIOT-W ATT UNIVERSITY June 2014 v1


11

The area of the reservoir considered as proved includes (1) the area delineated by
drilling and defined by fluid contacts, if any, and (2) the undrilled portions of the
reservoir that can reasonably be judged as commercially productive on the basis of
available geological and engineering data. In the absence of data on fluid contacts, the
lowest known occurrence of hydrocarbons controls the proved limit unless otherwise
indicated by definitive geological, engineering or performance data. Reserves may be
classified as proved if facilities to process and transport those reserves to market are
operational at the time of the estimate or there is a reasonable expectation that such
facilities will be installed. Reserves in undeveloped locations may be classified as
proved undeveloped provided (1) the locations are direct offsets to wells that have
indicated commercial production in the objective formation, (2) it is reasonably certain
such locations are within the known proved productive limits of the objective
formation, (3) the locations conform to existing well spacing regulations where
applicable, and (4) it is reasonably certain the locations will be developed. Reserves
from other locations are categorised as proved undeveloped only where interpretations
of geological and engineering data from wells indicate with reasonable certainty that
the objective formation is laterally continuous and contains commercially recoverable
petroleum at locations beyond direct offsets.

Reserves are those quantities of petroleum which are anticipated to be commercially


recovered from known accumulations from a given date forward.
i.e. Reserves refer to what can be produced in the future.

Figure 2 gives a schematic of reserves showing the progression with time.

Figure 2 Variations of Reserves During Field Life.

©HERIOT-W ATT UNIVERSITY June 2014 v1


12

What are the amounts termed that are not recoverable? The quantity of hydrocarbons
that remains in the reservoir are called remaining hydrocarbons in place, NOT
remaining reserves!

Reserves which are to be produced through the application of established improved


recovery methods are included in the proved classification when :

1 Successful testing by a pilot project or favourable response of an installed


program in the same or an analogous reservoir with similar rock and fluid
properties provides support for the analysis on which the project was based, and,

2 It is reasonably certain that the project will proceed. Reserves to be recovered


by improved recovery methods that have yet to be established through
commercially successful applications are included in the proved classification
only:

3 After a favourable production response from the subject reservoir from either;

a A representative pilot or

b An installed program where the response provides support for the analysis on
which the project is based and

4 It is reasonably certain the project will proceed.

3.3 Unproved Reserves


Unproved reserves are based on geologic and/or engineering data similar to that
used in estimates of proved reserves; but technical, contractual, economic, or
regulatory uncertainties preclude such reserves being classified as proved.

Unproved reserves may be further classified as probable reserves and possible


reserves. Unproved reserves may be estimated assuming future economic conditions
different from those prevailing at the time of the estimate. The effect of possible future
improvements in economic conditions and technological developments can be
expressed by allocating appropriate quantities of reserves to the probable and possible
classifications.

©HERIOT-W ATT UNIVERSITY June 2014 v1


13

3.3.1 Probable Reserves


Probable reserves are those unproved reserves which analysis of geological and
engineering data suggests are more likely than not to be recoverable. In this
context, when probabilistic methods are used, there should be at least a 50%
probability that the quantities actually recovered will equal or exceed the sum of
estimated proved plus probable reserves. In general, probable reserves may include :

1 Reserves anticipated to be proved by normal step-out drilling where subsurface


control is inadequate to classify these reserves as proved,

2 Reserves in formations that appear to be productive based on well log


characteristics but lack core data or definitive tests and which are not analogous
to producing or proved reservoirs in the area,

3 Incremental reserves attributable to infill drilling that could have been


classified as proved if closer statutory spacing had been approved at the time of
the estimate,

4 Reserves attributable to improved recovery methods that have been established


by repeated commercially successful applications when;

a a project or pilot is planned but not in operation and


b rock, fluid, and reservoir characteristics appear favourable for commercial
application,

5 Reserves in an area of the formation that appears to be separated from the


proved area by faulting and the geologic interpretation indicates the subject area
is structurally higher than the proved area,

6 Reserves attributable to a future workover, treatment, re-treatment, change of


equipment, or other mechanical procedures, where such procedure has not been
proved successful in wells which exhibit similar behaviour in analogous
reservoirs, and

7 Incremental reserves in proved reservoirs where an alternative interpretation of


performance or volumetric data indicates more reserves than can be classified as
proved.

©HERIOT-W ATT UNIVERSITY June 2014 v1


14

3.3.2 Possible Reserves


Possible reserves are those unproved reserves which analysis of geological and
engineering data suggests are less likely to be recoverable than probable reserves.

In this context, when probabilistic methods are used, there should be at least a 10%
probability that the quantities actually recovered will equal or exceed the sum of
estimated proved plus probable plus possible reserves. In general, possible reserves
may include:

1 Reserves which, based on geological interpretations, could possibly exist


beyond areas classified as probable,

2 Reserves in formations that appear to be petroleum bearing based on log and


core analysis but may not be productive at commercial rates,

3 Incremental reserves attributed to infill drilling that are subject to technical


uncertainty,

4 Reserves attributed to improved recovery methods when;

a a project or pilot is planned but not in operation and


b rock, fluid, and reservoir characteristics are such that a reasonable doubt
exists that the project will be commercial, and

5 Reserves in an area of the formation that appears to be separated from the


proved area by faulting and geological interpretation indicates the subject area is
structurally lower than the proved area.

3.4 Reserve Status Categories


Reserve status categories define the development and producing status of wells and
reservoirs.

3.4.1 Developed:
Developed reserves are expected to be recovered from existing wells including
reserves behind pipe (i.e. the well has not been perforated In these sections and this Is
termed behind pipe). Improved recovery reserves are considered developed only after
the necessary equipment has been installed, or when the costs to do so are relatively
minor. Developed reserves may be sub-categorised as producing or non-producing.

3.4.1.1 Producing:

©HERIOT-W ATT UNIVERSITY June 2014 v1


15

Reserves subcategorised as producing are expected to be recovered from completion


intervals which are open and producing at the time of the estimate. Improved recovery
reserves are considered producing only after the improved recovery project is in
operation.

3.4.1.2 Non-producing:
Reserves subcategorised as non-producing include shut-in and behind-pipe reserves.
Shut-in reserves are expected to be recovered from (1) completion intervals which are
open at the time of the estimate but which have not started producing, (2) wells which
were shut-in for market conditions or pipeline connections, or (3) wells not capable of
production for mechanical reasons. Behind-pipe reserves are expected to be recovered
from zones in existing wells, which will require additional completion work or future
recompletion prior to the start of production.

3.4.2 Undeveloped Reserves:


Undeveloped reserves are expected to be recovered:

1 From new wells on undrilled acreage,

2 From deepening existing wells to a different reservoir, or

3 Where a relatively large expenditure is required to;

a Recomplete an existing well or

b Install production or transportation facilities for primary or improved


recovery projects.

4 PROBABILISTIC REPRESENTATION OF RESERVES

Whereas in the deterministic approach the volumes are determined by the calculation
of values determined for the various parameters, with the probalistic statistical analysis
is used, using tools like Monte Carlo methods. The curve as shown in the Figure 3
below presents the probability that the reserves will have a volume greater or equal to
the chosen value.

©HERIOT-W ATT UNIVERSITY June 2014 v1


16

Figure 3 Probabilistic Representation of Recoverable Reserves.

On this curve:

The proven reserves represent the reserves volume corresponding to 90% probability
on the distribution curve.

The probable reserves represent the reserves volume corresponding to the difference
between 50 and 90% probability on the distribution curve.

The possible reserves represent the reserves volume corresponding to the difference
between 10 and 50% probability on the distribution curve.

As with the deterministic approach there is also some measure of subjectivity in the
probalistic approach. For each of the elements in the following equation, there is a
probability function expression in low, medium and high probabilities for the
particular values. A schematic of a possible distribution scenario for each of the
elements and the final result is given below in the Figure 4.

©HERIOT-W ATT UNIVERSITY June 2014 v1


17

Figure 4 Probablistic Reserve Estimates.

The resulting calculations produce a probability function for a field as shown in the
Figure 5 below, where the values for the three elements are shown:

Proven = 500 MM stb the P90 Figure.

Probable = 240 MM stb which together with the proven makes up the P50 Figure of
740MMstb

Possible = 120 MM stb which together with the proven and probable makes up the P10
value of 860MMstb

Figure 5 Reserves Cumulative Probability Distribution.

As a field is developed and the fluids are produced, the shape of the probability curve
changes. Probability Figures for reserves are gradually converted into recovery leaving
less uncertainty with respect to the reserves. This is illustrated in Figure 6.

©HERIOT-W ATT UNIVERSITY June 2014 v1


18

Figure 6 Ultimate Recovery and Reserves Distribution For a Mature Field.

5 VOLUME IN-PLACE CALCULATIONS

5.1 The volume of oil and gas in-place depends on a number of parameters:
The aerial coverage of the reservoir, A
The thickness of the reservoir rock contributing to the hydrocarbon volume, hn
The pore volume, as expressed by the porosity, , the reservoir quality rock.
The proportion of pore space occupied by the hydrocarbon (the saturation),
1-Sw

The simple equation used in calculation of the volume of fluids in the reservoir, V, is

V=Ahn(1-Sw) (1)

where:
A= average area
hn = nett thickness. nett thickness = gross thickness x nett: gross ratio
 = average porosity
Sw = average water saturation.

When expressed as stock tank (or standard gas) volumes, the equation above is
divided by the formation volume factor Bo (or Bg for gas.)

©HERIOT-W ATT UNIVERSITY June 2014 v1


19

(2)

To convert volumes at reservoir conditions to stock tank conditions formation volume


factors are required where Bo and Bg are the oil and gas formation volume factors.
These are defined in subsequent chapters. The expression of original oil in place is
termed the stock tank oil Initially in place (STOIIP).

The recovery factor, RF indicates the proportion of the in-place hydrocarbons expected
to be recovered. To convert in place volumes to reserves we need to multiply the
STOIIP by the recovery factor so that:

Reserves = STOIIP x RF (3)

The line over the various terms indicates the average value for these spatial
parameters.

The reservoir area A, will vary according to the category; proven, probable or
possible, that is being used to define the reserves.

5.2 The Recovery Factor, ER


The proportion of hydrocarbons recovered is called the recovery factor. This factor is
influenced by a whole range of factors including the rock and fluid properties and the
drive mechanisms. The variability of the formation characteristics, the heterogeneity
can have a large influence on recovery. The development process being implemented
and the geometries and location of wells again will also have a large influence.
Calculating recovery therefore in the early stages is not feasible and many assumptions
have to be included in such calculations. It is in this area that reservoir simulation can
give indications but the quality of the calculated figure is limited by the sparse amount
of quality data on which the simulation is based.

The American Petroleum Institute6 has analysed the recoveries of different fields and
correlations have been presented for different reservoir types and drive mechanisms.
The API presents correlations for recoveries, ER, as follows.

For sandstone and carbonate reservoirs with solution gas drive

(6)

©HERIOT-W ATT UNIVERSITY June 2014 v1


20

For sandstone reservoirs with water drive

(7)

b refers to bubble point conditions, i is the initial condition and a, refers to


abandonment pressure.

6 Phases of Development
During the development there are a number of phases. Not all of these phases may be
part of the plan. There is the initial production build up to the capacity of the facility as
wells are brought on stream. There is the plateau phase where the reservoir is
produced at a capacity limited by the associated production and processing facilities.
Different companies work with different lengths of the plateau phase and each project
will have its own duration. There comes a point when the reservoir is no longer able to
deliver fluids at this capacity and the reservoir goes into the decline phase. The
decline phase can be delayed by assisting the reservoir to produce the fluids by the use
of for example ‘lifting’ techniques such as down-hole pumps and gas lift. The decline
phase is often a difficult period to model and yet it can represent a significant amount
of the reserves. These phases are illustrated in Figure 7

Figure 7 Phases of Production.

©HERIOT-W ATT UNIVERSITY June 2014 v1


21

The challenge facing the industry is the issue of the proportion of hydrocarbons left
behind. The ability to extract a greater proportion of the in-place fluids is obviously a
target to be aimed at and over recent years recoveries have increased through the
application of innovative technology. Historically there have been three phases of
recovery considered. Primary recovery, which is that recovery obtained through the
natural energy of the reservoir.

Secondary recovery is considered when the energy is supplemented by injection of


fluids, for example gas or water, to maintain the pressure or partially maintain the
pressure. The injected fluid also acts as a displacing fluid sweeping the oil to the
producing wells. After sweeping the reservoir with water or gas there will still be
remaining oil: oil at a high saturation where the water for a range of reasons, for
example, well spacing, viscosity, reservoir characteristics to name just a few, has by-
passed the oil. The oil which has been contacted by the injected fluid will not be
completely displaced from the porous media. Because of characteristics of the rock
and the fluids a residual saturation of fluid is held within the rock. Both of these
unrecovered amounts, the by-passed oil and the residual oil are a target for enhanced
oil recovery methods, EOR.
Much effort was put into enhanced oil recovery (EOR) research up until the mid
seventies. Sometimes it is termed tertiary recovery. When the oil price has dropped the
economics of many of the proposed methods are not viable. Many are based on the
injection of chemicals which are often oil based. The subject of EOR has not been
forgotten and innovative methods are being investigated within the more volatile oil
price arena. Figure 8 gives a schematic representation of the various phases of
development and includes the various improved recovery methods. More recently a
new term has been introduced called Improved Oil Recovery (IOR). IOR is more
loosely defined and covers all approaches which might be used to improve the
recovery of hydrocarbons in place. Clearly it is not as specific as EOR but provides
more of an achievable target than perhaps some of the more sophisticated EOR
methods.

©HERIOT-W ATT UNIVERSITY June 2014 v1


22

Figure 8 Oil Recovery Mechanisms.

REFERENCES

1 Craft, B.C. and Hawkins, M.F. Applied Reservoir Engineering, Prentice-Hall


Inc.
2 Dake, L.P., The Practise of Reservoir Engineering. Elsevier
3 Society Of Petroleum Engineers. Reserves Definitions.
4 Chierici,G.L. Principles of Petroleum Reservoir Engineering. Vol 1 Springer
Verlag
5 Hollois, A.P. Some petroleum engineering considerations in the change over of
the Rough Gas field to the storage mode. Paper EUR 295 Proc Europec.
6 API. A Statistical Study of the Recovery Efficiency. American Petroleum
Institute. Bull D14, 1st Edition
7 Archer, J.S. and Wall,C.G. Petroleum Engineering Principles and Practise,
Graham and Trotman.

©HERIOT-W ATT UNIVERSITY June 2014 v1


23

TOPIC 2 RESERVOIR PRESSURES AND


TEMPERATURES

1 INTRODUCTION

2 ABNORMAL PRESSURES

3 FLUID PRESSURES IN HYDROCARBON SYSTEMS

4 PRESSURE GRADIENTS AROUND WATER-OIL CONTACT

5 TECHNIQUES FOR PRESSURE MEASUREMENT

6 RESERVOIR TEMPERATURE

LEARNING OUTCOMES

Having worked through this chapter the Student will be able to:

• Define the terms overburden pressure, hydrostatic pressure and hydrodynamic


pressure.

• Draw the normal hydrostatic pressure gradient for water systems.

• Define normal pressured reservoirs, overpressured reservoirs and


underpressured reservoirs

• Describe briefly and sketch the pressure gradients associated with overpressured
and underpressured reservoirs.

• Describe briefly, sketch and present equations for the pressures in a water
supported oil and gas bearing formation and identify the oil water and gas oil
contacts.

©HERIOT-W ATT UNIVERSITY June 2014 v1


24

• Illustrate how a downhole formation pressure device can be used to discriminate


permeability layers after production has commenced.

1. INTRODUCTION

Determining the magnitude and variation of pressures in a reservoir is an important


aspect in understanding various aspects of the reservoir, both during the exploration
phase but also once production has commenced.

Oil and gas accumulations are found at a range of sub-surface depths. At these depths
pressure exists as a result of the depositional process and from the fluids contained
within the porous media. These pressures are called overburden pressures and fluid
pressures. These pressures are illustrated in Figure 1.

The overburden pressure is caused by the pressure of rock which is transmitted


through the sub-surface by grain-to grain contacts. This overburden pressure is of the
order of 1 psi/ft. If we use this overburden pressure gradient of 1 psi/ft. then the
overburden pressure Pov, in psig at a depth of D feet is

Pov = 1.0D (1)

The overburden pressure is balanced in part by the pressure of the fluid within the pore
space, the pore pressure, and also by the grains of rock under compaction. In
unconsolidated sands, loose sands, the overburden pressure is totally supported by the
fluid and the fluid pressure Pf is equal to the overburden pressure Pov. In deposited
formations like reservoir rocks the fluid pressure is not supporting the rocks above but
arises from the continuity of the aqueous phase from the surface to the depth D in the
reservoir. This fluid pressure is called the hydrostatic pressure. The hydrostatic
pressure is imposed by a column of fluid at rest. Its value depends on the density of the
water w, which is affected by salinity. In a sedimentary basin, where sediment has
settled in a region of water and hydrocarbons have been generated and trapped, we can
expect a hydrostatic pressure. For a column of fresh water the hydrostatic pressure is
0.433 psi/ft. For water with 55,000 ppm of dissolved salts the gradient is 0.45 psi/ft;
for 88,000 ppm of dissolved salts the gradient is about 0.465 psi/ft.

Its variation with depth is given by the equation.

Pf = wDg (2)

©HERIOT-W ATT UNIVERSITY June 2014 v1


25

where g is the acceleration due to gravity.

There is another fluid pressure which arises as a result of fluid movement and that is
called the hydrodynamic pressure. This is the fluid potential pressure gradient which
is caused by fluid flow. This however does not contribute to in-situ pressures at rest.

Figure 1 Gives the relationship between the overburden pressure and the hydrostatic pressure.

Fluid pressure in hydrocarbon accumulations are dictated by the prevailing water


pressure in the vicinity of the reservoir. In a normal situation the water pressure at any
depth is:

(3)

where dP/dD is the hydrostatic pressure gradient

This equation assumes continuity of water pressure from the surface and constant
salinity. In most cases even though the water bearing sands are divided between
impermeable shales, any break of such sealing systems will lead to hydrostatic
pressure continuity, but the salinity can vary with depth.

Reservoirs whose water pressure gradient when extrapolated to zero depth give an
absolute pressure equivalent to atmospheric pressure are called normal pressured
reservoirs.

©HERIOT-W ATT UNIVERSITY June 2014 v1


26

Subsurface Pressures

Hydrostatic pressure is defined as the pressure exerted by water at any given point
within the fluid at rest, i.e. generated by the density of the fluid and vertical height of
the fluid column.

P = wgh

in metric units:

N kg m
= x xm
m 2 m 3 s2

F
m=a

N N 1 m
 2 =m x 3 x 2 xm
m m s
s2

N N
2 = 2
m m

i.e. the item g is a Force/volume term, or

force N Pa
area x depth m x m m )
( 2 =

In imperial units, the g term is usually quoted in lb/gal. To convert this to


psi/ft requires the following:

1 ft3 = 7.48 gal.

1 ft2 = 144 in2

psi lb gal 1 ft 2
1  1 x 7.48 3 x 2
ft gal ft 144 in

1 psi/ft
i.e. = 0.0519
1 lb/ gal

e.g., fresh water with a density of 8.34 lb/gal exerts a pressure of 8.34 x
0.0519 = 0.433 psi/ft

©HERIOT-W ATT UNIVERSITY June 2014 v1


27

i.e., in a vertical column of this fluid, the pressure at any depth is 0.433 psi
greater for each 1 ft increase in depth.

Overburden Pressure at any point in the formation is the pressure exerted by


the total weight of the overlying formations. If the thickness and density of the
formations is know (e.g. a density log is available in a well), then the
overburden is the sum of each layer thickness and its density:
z

s   (z)dz
0

s is the overburden pressure, z is the depth interval,  is the density.

For imperial / metric units in common use:


z

s(psi)  0.433 (z)dz


0

 in g/cm3, z in feet. [ 1g/cm3 = 0.433 psi/ft]

The following table shows some grain densities of sedimentary rocks.

Lithology Matrix Lithology Matrix Lithology Matrix


Density Density Density
g/cm3 g/cm3 g/cm3

Sandstone 2.65 Anhydrate 2.98 Clay 2.70 - 2.80


Limestone 2.71 Halite 2.03
Dolomite 2.87 Gypsum 2.35

The average bulk density (overburden) for sedimentary basins is approximately


2.3 g/cc or 1 psi/ft.

For comparison, the density of common reservoir fluids is rather different:

Fluid Density
g/cm3

Fresh Water 1.00


Salt Water 1.15
Oil 0.80

Fluid Gradients

©HERIOT-W ATT UNIVERSITY June 2014 v1


28

Consider the following diagram of a normally pressured reservoir sequence. The depth
to the oil water contact is x; the depth to the gas oil contact is z. The depth of interest
is denoted by h (i.e. h can vary through the reservoir section where you want to
calculate the pressure)

surface
h=0

h x y z
gas gas

oil
water
water

Assumption: Pores of all sediments are filled with water in the crust. Pores are all
interconnected, transition zones between fluids, oil/water, gas/oil, gas/water are
relatively small.

The change in pressure, dp = gdh

dp
is the hydrostati c pressure gradient.
dh

If p1 is the pressure at h1, p2 is the pressure at h2 and  does not change much with
depth, p2-p1 = g(h2-h1). Therefore the following formulae for initial hydrostatic
pressures in the pore spaces are applicable.

Water filled pores

pw -psurface = wgh

pw = pressure at depth h

psurface = pressure at depth 0 (surface), mean atmospheric pressure

The density of formation water,w does vary with depth because of different water
salinities. It can be fixed by measuring pressure at 2 depths.

Oil filled pores

©HERIOT-W ATT UNIVERSITY June 2014 v1


29

po(h) - po(x) = og(h - x)

po(x) is the pressure at the oil/water contact, Po(h) is the pressure at any depth h.

Gas filled pores

pg(h) - pg(y) = gg(h-y)

pg(y) is the pressure at the gas/oil contact, pg(h) is the pressure at any depth h.

If there is no difference in pressure at the oil/water contact and the oil/gas


contact then
pw(x) = po(x); po(y) = pg(y)

po(h) = psurface+ wgx + og(h - x)

= psurface + (w - o)gx + ogh. i.e. po(h) = ogh + constant


_____________________

and
pg(h) = psurface + wgx + o(y - x)g + g(h - y)g

= psurface + (w - o)gx + (o - g)gy + ggh i.e. pg(h) = ggh+ constant
_________________________________

Similarly for a GWC:

pg(h) = psurface + (w - g)gz + ggh


__________________________

For example, an OWC is found at 10,000 ft TVD (true vertical depth). Water
gradient = 0.45 psi/ft, oil gradient = 0.35 psi/ft. What is the pressure at
9,000 ft?

po = psurface + (w - o)gx + ogh : x = 10,000 ft, h = 9,000 ft.

= 14.7 + (0.45 - 0.35) x 10,000 + 0.35 x 9,000

= 4164.7 psi
OR:
pw at 10,000 ft = 0.45 x 10,000 + 14.7 = 4514.7 psi

©HERIOT-W ATT UNIVERSITY June 2014 v1


30

at OWC, pw = po

po = 4514.7 psi
= 0.35 x 10,000 + constant

constant = 4514.7 - 0.35 x 10,000 = 1014.7

po at 9,000 ft = 0.35 x 9,000 + constant


= 0.35 x 9,000 + 1014.7
= 4164.7 psi

These equations can be used to determine pressures within gas, oil and water
fluid regimes.

2. ABNORMAL PRESSURE

Under certain conditions, fluid pressures may depart substantially from the normal
pressure. Overpressured reservoirs are those where the hydrostatic pressure is
greater than the normal pressure and underpressured reservoirs are below normal
pressure, Figure 1. They are called abnormal pressured reservoirs and can be
defined by the equation:

(4)

where C is a constant, being positive for overpressured and negative for an


underpressured system.

For abnormally pressured reservoirs, the sand is sealed off from the surrounding strata
so that there is no hydrostatic pressure continuity to the surface.

Conditions which cause abnormal fluid pressure in water bearing sands have been
identified by Bradley 2 and include (Figure 2):

©HERIOT-W ATT UNIVERSITY June 2014 v1


31

Figure 2 Causes of overpressurring

• Thermal effects, causing expansion or contraction of water which is unable to


escape; an increase in temperature of 1˚F can cause an increase of 125 psi in a
sealed fresh water system.

• Rapid burial of sediments consisting of layers of sand and clay. Speed of burial
does not allow fluids to escape from pore space.

• Geological changes such as uplifting of the reservoir, or surface erosion both of


which result in the water pressure being too high for the depth of burial. The
opposite occurs in a down thrown reservoir.

• Osmosis between waters having different salinity, the sealing shale acting as a
semi-permeable membrane. If the water within the seal is more saline than the
surrounding water, the osmosis will cause a high pressure and vice versa.

With abnormally pressured reservoirs a permeability barrier must exist, which inhibits
pressure release. These may be lithological or structural. Common lithological
barriers are evaporates and shales.

©HERIOT-W ATT UNIVERSITY June 2014 v1


32

If reservoirs are all normal pressured systems then the pressure gradient for these
reservoirs would be virtually all the same, other than from the influence of salinity.
The figure below shows the water pressure gradients for a number of reservoirs in the
North Sea and indicates the significant overpressuring in this region.

3
Figure 3 Examples of overpressured reservoirs in the North Sea

3. FLUID PRESSURES IN HYDROCARBON SYSTEMS

Pressure gradients in hydrocarbon systems are different from those of water systems
and are determined by the oil and gas phase in-situ densities, o and g of each fluid.

The pressure gradients are a function of gas and oil composition but typically are:

©HERIOT-W ATT UNIVERSITY June 2014 v1


33

(5)

(6)

(7)

For a reservoir containing both oil and a free gas cap a pressure distribution results, as
in the Figure 4. As can be seen, the composition of the respective fluids gives rise to
different pressure gradients indicated above. These gradients will be determined by the
density of the fluids which result from the specific composition of the fluids.

Figure 4 Pressure distribution for an oil reservoir with a gas cap and an oil-water contact.

The nature of the pressure regime and the position and recognition of fluid contacts are
very important to the reservoir engineer in evaluating reserves, and determining
depletion policy.

The data used for these fluid contacts comes from:

©HERIOT-W ATT UNIVERSITY June 2014 v1


34

(i) Pressure surveys


(ii) Equilibrium pressures from well tests
(iii) Flow of fluid from particular minimum and maximum depth
(iv) Fluid densities from reservoir samples
(v) Saturation data from wireline logs
(vi) Capillary pressure data from cores
(vii) Fluid saturation from cores

4. PRESSURE GRADIENTS AROUND THE WATER-OIL CONTACT

Water is always present in reservoir rocks and the pressure in the water phase Pw and
the pressure in the hydrocarbon phase Po are different. If P is the pressure at the
oil/water contact where the water saturation is 100%, then the pressure above this
contact for the hydrocarbon and water are:

Po = P - ogh (8)

Pw = P - wgh (9)

The difference between these two pressures is the capillary pressure Pc. In a
homogenous water-wet reservoir with an oil-water contact the variation of saturation
and phase pressure from the water zone through the capillary transition zone into the
oil is shown in Figure 5.

Figure 5 Pressure Gradients around the Water-Oil Contact

©HERIOT-W ATT UNIVERSITY June 2014 v1


35

In the transition zone the phase pressure difference is given by the capillary pressure
which is a function of the wetting phase saturation.
Pc = Po - Pw (10)

at hydrostatic equilibrium
Pc(Sw) = ∆gh
∆ = w-o
h = height above free water level

The free water level, FWL, is not coincident with the oil-water contact OWC. The
water contact corresponds to the depth at which the oil saturation starts to increase
from water zone. The free water level is the depth at which the capillary pressure is
zero.

The difference in depth between the oil-water contact and the free water level depends
on the capillary pressure which in turn is a function of permeability, grain size etc.

Providing the phase is continuous the pressures in the respective phases are:

Po = PFWL - ogh (11)


Pw = PFWL - wgh (12)

On the depth-pressure diagram the intersection of the continuous phase pressure line
occurs at the free water level.

5. TECHNIQUES FOR PRESSURE MEASUREMENT

Earlier tests for vertical pressure logging have been replaced by open-hole testing
devices that measure the vertical pressure distribution in the well, and recover
formation samples.

One such device which was introduced in the mid seventies which has established
itself in reservoir evaluation is the repeat formation tester RFT or more recently,
Modular dynamic tool MDT (Schlumberger trade name). It was initially developed as
a device to take samples. Over the years however its main application is to provide
pressure -depth profiles over reservoir intervals. The device places a probe through the
well mud cake and allows small volumes of fluid to be taken and pressure
measurements to be made (Figure 6). It can only be operated therefore in an open hole
environment. The unit can be set at different locations in the well and the pressure
gradient thereby obtained.

©HERIOT-W ATT UNIVERSITY June 2014 v1


36

Figure 6 Original Schematic of the RFT Tool

By comparing current pressure information with those obtained prior to production,


important reservoir description can be obtained which will aid reservoir depletion,
completion decisions and reservoir simulation.

In the Montrose Field in the North Sea, figure 7 shows the pressure depth survey in a
well. Only the top 45ft of the 75ft oil column had been perforated. The initial pressure
gradient indicates the oil and water gradients at the condition of hydrostatic
equilibrium (i.e. this is the first exploration well at a time when there had been no
production from the reservoir). The second survey shows a survey after 2 years of
high production rate, and reveals the reservoir behaviour under dynamic conditions.
The various changes in slope in the pressure profile reveal the partial restricted flow in
certain layers. Similar surveys in each new development wells (Figure 8) show the
similar profiles and enable the detailed layered structure of the reservoir to be
characterised which is important for reservoir simulation purposes. Note that the actual
gradients can vary slightly from the original well to the subsequently drilled wells, and
this variation will reflect the unique nature of the reservoir at the location of each well.

©HERIOT-W ATT UNIVERSITY June 2014 v1


37

Figure 7 RFT Pressure Survey in Development Well of Montrose Field 3.

Time since start


of production
when well was
drilled
Initial
1 year
2 years
3 years
4 years
5 years

Figure 8 RFT Pressure Surveys on a number of Montrose Wells3

©HERIOT-W ATT UNIVERSITY June 2014 v1


38

6. RESERVOIR TEMPERATURE
The temperature of the earth increases from the surface to centre. The heat flow
outwards through the Earth’s crust generates a geothermal gradient, . This
temperature variation conforms to both a local and regional geothermal gradient,
resulting from the thermal characteristics of the lithology and more massive
phenomenon associated with the thickness of the earth’s crust along ridges, rifts and
plate boundaries.

In most petroleum basins the geothermal gradient is of the order of 1.6˚F/100 ft. (0.029
K/m) The thermal characteristics of the reservoir rock and overburden give rise to
large thermal capacity and with a large surface area in the porous reservoir one can
assume that flow processes in a reservoir occur at constant reservoir temperature.
The local geothermal gradient will be influenced by associated geological features like
volcanic intrusions etc. The local geothermal gradient can be deduced from wellbore
temperature surveys . However they have to be made under stabilised conditions since
they can be influenced by transient cooling effects of circulating and injected fluids.

During drilling the local thermal gradient can be disturbed and by analysis of the
variation of temperature with time using a bottom hole temperature (BHT) gauge the
local undisturbed temperature can be obtained.

Without temperature surveys the temperature at a vertical depth can be estimated using
a surface temperature of 15 oC (60 oF) at a depth D.

T(D) = 288.2 + D (K)

REFERENCES

1. Dake,L.P. Fundamentals of Reservoir Engineering. Elsevier

2. Bradley,J.S. Abnormal Formation Pressure. The American Association of


Petroleum Geologists Bulletin. Vol 59, No6

3. Bishlawi,M and Moore,RL: Montrose Field Reservoir Management. SPE


Europec Conference, London,(EUR166)

©HERIOT-W ATT UNIVERSITY June 2014 v1


39

TOPIC 3 RESERVOIR FLUIDS COMPOSITION

1 INTRODUCTION

2 HYDROCARBONS
2.1 Chemistry of Hydrocarbons
2.2 Alkanes or Paraffinic Hydrocarbons
2.3 Isomerism
2.4 Unsaturated Hydrocarbons
2.5 Napthene Series
2.6 Aromatics
2.7 Asphalts

3 NON-HYDROCARBON COMPOUNDS

4 COMPOSITIONAL DESCRIPTION FOR RESERVOIR ENGINEERING


4.1 Definitions of Composition in Reservoir Engineering

5 GENERAL ANALYSIS
5.1 Surface Condition Characterisation

LEARNING OUTCOMES

Having worked through this chapter the Student will be able to:

• Describe briefly the origin, nature and appearance of petroleum fluids.


• Be aware that the principal components of petroleum fluids to be hydrocarbons.
• Draw a diagram illustrating the classification of hydrocarbons and to identify
paraffin’s (alkanes), aromatics and cyclic aliphatics (napthas).
• List the non-hydrocarbon compounds which might be present in small quantities
in reservoir fluids.
• Define the black oil model description of the composition of a reservoir fluid.
• Define the compositional model description of a reservoir fluid.
• Describe briefly the concept of pseudo components in fluid composition
characterization.
• Be aware of the American Petroleum Institute descriptor for petroleum fluids
°API.
• Be able to calculate the °API gravity given the specific gravity.

©HERIOT-W ATT UNIVERSITY June 2014 v1


40

1 INTRODUCTION

Petroleum deposits vary widely in chemical composition and depending on location


have entirely different physical and chemical properties. The very complex
characteristics are evident from the many products which can be produced from oil and
gas.

What is petroleum? Petroleum is a mixture of naturally occurring hydrocarbons which


may exist in the solid, liquid or gaseous states, depending on the conditions of
temperature and pressure to which it is subjected.1

Petroleum deposits occurring as a gaseous state are termed natural gas, in the liquid
state as petroleum oil or crude oil and in the solid state as tars, asphalts and waxes.

For a mixture with small molecules it will be a gas at normal temperature and pressure
(NTP). Mixtures containing larger molecules will be a liquid at NTP and larger
molecules as a solid state, for example, tars and asphalts.

The exact origin of these deposits is not clear but is considered to be from plant,
animal and marine life through thermal and bacterial breakdown.

The composition of crude oil consists mainly of organic compounds, principally


hydrocarbons with small percentages of inorganic non-hydrocarbon compounds. Such
as carbon dioxide, sulphur, nitrogen and metal compounds. The hydrocarbons may
include the lightest (C1 methane ) to napthenes and polycyclics with high molecular
weights.

The appearance varies from gases, through very clear liquids, yellow liquids to a dark,
often black, highly viscous material, the variety obviously being a function of
composition. Although the principal elements are carbon (84-87%), and hydrogen (11-
14%), crude oil can vary from a very light brown liquid with a viscosity similar to
water to a very viscous tar like material .

Water is always present in the pore space of a reservoir, since the original depositional
environment for the rocks was water. This water has subsequently been displaced by
the influx of hydrocarbons but not totally since surface tension forces acting in the
rock pore space cause some of the water to be retained.

©HERIOT-W ATT UNIVERSITY June 2014 v1


41

For reservoir engineering purposes the description of the composition is an important


characterisation parameter for the determination of a range of physical parameters
important in various reservoir volumetric and flow calculations. It is not the concern
of the reservoir engineer to determine the composition with respect to understanding
the potential to separate the material to a range of saleable products. For this reason
therefore simplistic characterisation approaches are used.

The two compositional characterisation approaches used are the compositional model
and the black oil model. The basis of the compositional model is a multicomponent
description in terms of hydrocarbons and the black oil model is a two component
description in terms of produced oil, stock tank oil and produced gas, solution gas.

2 HYDROCARBONS

2.1 Chemistry of Hydrocarbons


The compositional model uses hydrocarbons as the descriptor since hydrocarbons
represent the largest proportion in petroleum fluids. It is important to review briefly
the chemistry of hydrocarbons.

The hydrocarbon series is represented in figure 1 below

Figure 1 Classification of Hydrocarbon.

The hydrocarbons divide into two groupings with respect to the arrangement of the
carbon molecules and the bonds between the carbon molecules. The arrangement of
the molecules are open chain or cyclic and the bonds between the carbon are saturated
(single) bonds or unsaturated (multiple) bonds.

2.2 Alkanes or Paraffinic Hydrocarbons

©HERIOT-W ATT UNIVERSITY June 2014 v1


42

The largest series is the alkanes or paraffins which are open chain molecules with
saturated bonds. Carbon has a valance of four and therefore the formula for these
compounds is CnH2n+2. These saturated hydrocarbons include all the paraffins in
which the valence of the carbon atoms is satisfied by single covalent bonds. This type
of structure is very stable. Unsaturated hydrocarbons are those where the valence of
some of the carbon atoms is not satisfied with single covalent bonds so they are
connected by two or more bonds which make them less stable and more prone to
chemical change.

The paraffin series begins with methane (CH4), a gas. Pentane to pentadecane are
liquids and the chief constituents of uncracked gasoline. Its higher members are waxy
solids. In a given bore hole the wax may clog the pore space next to the hole as gas
expands and cools.

The paraffins are the largest constituent of crude oil and are characterised by their
chemical inertness. Clearly they would not have remained as they are if this were not so.

2.3 Isomerism
From methane to propane there is only one way to arrange the branched chains
however above propane there are alternative arrangements and these are called
isomers.

Structural formulae do not represent the actual structure of the molecules. Isomers are
substances of the same composition that have different molecular structure and
therefore different properties, for example, normal butane and isobutane.

normal butane CH3CH2CH2CH3 - B.Pt. 31.1˚F

isobutane CH3CHCH3 - B.Pt. 10.9˚F

CH3

Pentane has three structures (isomers). Clearly the number of isomers increase as the
number of carbon atoms increases. Hexane has 5 isomers and heptane 9.

Table 1 below gives some of the basic physical properties of the more common
hydrocarbons of the paraffin series and Table 2 lists the state of the various pure
components demonstrating that components which might be solid on their own
contribute to liquid states when part of a mixture. Figure 2 gives some structural
formula for three paraffin compounds.

©HERIOT-W ATT UNIVERSITY June 2014 v1


43

Table 1 Physical properties of common hydrocarbons.

Table 2 Alkanes or Paraffin Hydrocarbons Cn H 2n + 2.

©HERIOT-W ATT UNIVERSITY June 2014 v1


44

Figure 2 Some structural formulae for saturated hydrocarbons.

2.4 Unsaturated Hydrocarbons


These are hydrocarbons which have double or triple bonds between carbon atoms.
They have the potential to add more hydrogen or other elements and are therefore
termed unsaturated. There are termed the olefins, and there are two types, alkenes, for
example ethylene, CH2=CH2, which have a carbon-carbon double bond and alkynes,
for example acetylene,CH=CH which have a carbon carbon triple bond. Both
compound types being unsaturated are generally very reactive and hence are not
found in reservoir fluids.

2.5 Napthene Series


The napthene series (CnH2n) sometimes called cycloparaffins or alicyclic hydrocarbons
are identified by having single covalent bonds but the carbon chain is closed and is
saturated. They are very stable and are important constituents of crude oil. Their
chemical properties are similar to those of the paraffins. A crude oil with a high
napthene content is referred to as an napthenic based crude oil. An example is
cyclohexane C6H12. Figure 3 gives the structural formula for two napthenic
compounds.

Figure 3 Structural formula for two naphenic compounds.

2.6 Aromatics
The aromatic series (CnH2n-6) is an unsaturated closed-ring series, based on the
benzene compound and the compounds are characterised by a strong aromatic odour.
Various aromatic compounds are found in crude oils. The closed ring structure gives
them a greater stability than open compounds where double or triple bonds occur.
Figure 4 gives the structural formula for two aromatic compounds.

©HERIOT-W ATT UNIVERSITY June 2014 v1


45

Figure 4 Structural formula for two aromatic compounds.

The aromatic-napthene based crudes are usually associated with limestone and
dolomite reservoirs such as those found in Iran, the Persian Gulf and Borneo.

Some crude oils used to be described, more from a refining perspective, according to
the relative amount of these non paraffin compounds. Crude oils would be called
paraffinic, napthenic or aromatic. It is not a classification of value in reservoir
engineering.

©HERIOT-W ATT UNIVERSITY June 2014 v1


46

Table 3 Physical properties of some common petroleum reservoir fluid constituents.

2.7 Asphalts
Asphalt is not a series by itself. Asphalts are highly viscous to semi-solid, brown-
black hydrocarbons of high molecular weight usually containing a lot of sulphur and
nitrogen, which are undesirable components, and oxygen. Asphalts are closely related
to the napthene series and because of their high nitrogen and oxygen content they may
be considered juvenile oil, not fully developed.

3 NON-HYDROCARBON COMPOUNDS

Although small in volume, generally less than 1%, non-hydrocarbon compounds have
a significant influence on the nature of the produced fluids with respect to processing
and the quality of the products.

The more common non-hydrocarbon constituents which may occur are:


sulphur, oxygen, nitrogen compounds, carbon dioxide and water.

Sulphur and its associated compounds represent 0.04% - 5% by weight. These


corrosive compounds include sulphur, hydrogen sulphide (H2S ),which is very toxic,
and mercaptans of low molecular weight ( these are produced during distillation and
require special metals to avoid corrosion). Non-corrosive sulphur materials include
sulphides. Sulphur compounds have a bad smell and both the corrosive and non-
corrosive forms are undesirable. On combustion these products produce S02 and S03
which are undesirable from an environmental perspective.

Oxygen compounds, up to 0.5% wt., are present in some crudes and decompose to
form napthenic acids on distillation, which may be very corrosive.

Nitrogen content is generally less than 0.1% wt., but can be as much as 2%. Nitrogen
compounds are complex. Gaseous nitrogen reduces the thermal quality of natural gas
and needs to be blended with high quality natural gas if present at the higher levels.

Carbon Dioxide is a very common constituent of reservoir fluids, especially in gases


and gas condensates. Like oxygen it is a source of corrosion. It reacts with water to
form carbonic acid and iron to form iron carbonate. Carbon dioxide like methane has a
significant impact on the physical properties of the reservoir fluids.

©HERIOT-W ATT UNIVERSITY June 2014 v1


47

Other compounds. Metals may be found in crude oils at low concentration and are of
little significance. Metals such as copper, iron, nickel, vanadium and zinc may be
present. Produced natural gas may contain helium, hydrogen and mercury.

Inorganic compounds. The non-oil produced fluids like water will clearly contain
compounds arising from the minerals present in the rock, their concentration will
therefore vary according to the reservoir. Their composition however can have a very
significant effect on the reservoir behaviour with respect to their compatibility with
injected fluids. The precipitation of salts, scale, is a serious issue in reservoir
management.

Many of these salts need to be removed on refining as some generate HC1 when heated with
water.

4. COMPOSITIONAL DESCRIPTION FOR RESERVOIR


ENGINEERING

4.1 Definitions of Composition in Reservoir Engineering


In petroleum engineering, and specifically in reservoir engineering, the main issue is
one of the physical behaviour and characteristics of the petroleum fluids. The
composition of the fluid clearly has a significant impact on the behaviour and
properties. In petroleum engineering therefore the description of the composition is a
key to determine the physical properties and behaviour.

Two models are used in this industry to describe the composition for physical property
prediction purposes, the black-oil model and the compositional model.

The black-oil model is a 2 component description of the fluid where the two
components are, the fluids produced at surface, stock tank oil and solution gas.
Associated with this model are black-oil parameters like solution gas-oil ratio and the
oil formation volume factor. These parameters are discussed in the chapter on liquid
properties.

The compositional model is a compositional description based on the paraffin series


CnH2n+2. The fluid is described with individual compositions of normal paraffins up to
a limiting C number. Historically, C6, more common now to go up to C9, or even
higher. Components greater than the limiting C number are lumped together and
defined as a C+ component.

©HERIOT-W ATT UNIVERSITY June 2014 v1


48

Isomers, normal and iso are usually identified up to pentane. Non paraffinic
compounds are assigned to the next higher paraffin according to its volatility. The
material representing all compounds above the limiting carbon number are called the
C+ fraction, so C7+ for a limiting value of C6 and C10+ for a limiting value of C9.

The physical properties of paraffins up to the limiting C number are well known and
documented. The C+ component is however unique to the fluid and therefore two
properties are used to characterise it, apparent molecular weight and specific
gravity.

Figure 5 illustrates the compositional model and its application as reservoir fluids are
produced to surface. Although the individual components contribute to a single liquid
reservoir phase for an oil, when the fluids are produced to surface they produce a gas
phase, solution gas, and a liquid phase, stock tank oil. The distribution characteristics
of the individual components is complex and not just a function of temperature and
pressure. For reservoir fluids the composition is also an influence on the distribution.
This makes it a difficult task to predict this distribution perspective since reservoir
fluid compositions are unique. Improved methods of chemical analysis make it
possible to describe the oil up to a C value of C29. Although such definitions provide a
very accurate description, the associated computer effort in using such a
comprehensive description does lead to the use of pseudo components. Pseudo
components are obtained by grouping the various C number compositions, thereby
reducing the description to 4 or 5 "pseudo components". A number of methods exist to
group the various C values and other components.

©HERIOT-W ATT UNIVERSITY June 2014 v1


49

Figure 5 Compositional model

5. GENERAL ANALYSIS

5.1 Surface condition characterisation


Reservoirs as well as having unique compositions also exist at specific pressures and
temperatures. It is important therefore to provide a common basis for describing the
quantities of fluids in the reservoir and throughout the production process.

The basis chosen is the fluids at surface conditions, the surface conditions being 14.7
psia or 101.3 kPa and 60ºF or 298K. These conditions are called standard conditions.
For gas therefore this yields standard cubic feet SCF or standard cubic meters. It is
useful to consider this expression not as volume but as mass, the volume of which will
vary according to density. For liquids we express surface conditions as stock tank
volumes either stock tank barrels STB or stock tank cubic meters STM3. The relative
amount of gas to oil is expressed by the gas-oil ratio GOR SCF/STB.

©HERIOT-W ATT UNIVERSITY June 2014 v1


50

Since there are so many types of oil, each with a wide range of specific gravity, an
arbitrary non-linear relationship was developed by the American Petroleum Institute
(API) to classify crude oils by weight on a linear-scaled hydrometer. The observed
readings are always corrected for temperature to 60ºF, by using a prepared table of
standard values.

141.5
Degrees API = -131.5
Sp.Gr.at 60ºF (1)

Sp.Gr = specific gravity relative to water at 60ºF.

The API gravity of water is 10º. A light crude oil would have an API gravity of 40º,
while a heavy crude would have an API gravity of less than 20º. In the field, the API
gravity is readily measured using a calibrated hydrometer.

There are no definitions for categorising reservoir fluids, but the following Table 4
indicates typical GOR, API and gas and oil gravities for the five main types. The
compositions show that the dry gases contain mostly paraffins, with the fraction of
longer chain components increasing as the GOR and API gravity of the fluids
decrease.

©HERIOT-W ATT UNIVERSITY June 2014 v1


51

Table 4 Typical values for different reservoir fluids.

REFERENCES.

1. Amyx, J.W., Bass, D.M., and Whiting, R.L."Petroleum Reservoir Engineering",


McGraw-Hill Book Company, New York

©HERIOT-W ATT UNIVERSITY June 2014 v1


52

TOPIC 4 PHASE BEHAVIOUR OF


HYDROCARBON SYSTEMS

1 INTRODUCTION AND DEFINITIONS

2 PHASE BEHAVIOUR OF PURE SUBSTANCES


2.1 The Phase Diagram

3 USE OF PHASE DIAGRAMS


3.1 Pressure - Temperature Diagrams (PT)
3.2 Pressure Volume Diagram (PV)

4 TWO COMPONENT SYSTEMS


4.1 Pressure Volume Diagram
4.2 Pressure Temperature Diagram
4.3 Critical Point
4.4 Retrograde Condensation

5 MULTI-COMPONENT HYDROCARBON
5.1 Oil Systems (Black Oils and Volatile Oils)
5.2 Retrograde Condensate Gas
5.3 Wet Gas
5.4 Dry Gas

6 COMPARISON OF THE PHASE DIAGRAMS OF RESERVOIR FLUIDS

7 RESERVOIRS WITH A GAS CAP

©HERIOT-W ATT UNIVERSITY June 2014 v1


53

LEARNING OUTCOMES

Having worked through this chapter the Student will be able to:

General
• Define; system, components, phases, equilibrium, intensive and extensive
properties.

Pure Components
• Sketch a pressure-temperature (PT) diagram for a pure component and illustrate
on it; the vapour-pressure line, critical point, triple point, sublimation-pressure
line, the melting point line, the liquid, gas and solid phase zones.
• Define the critical pressure and critical temperature for a pure component.
• Describe briefly with the aid of a PT diagram the behavior of a pure component
system below( left ) and above ( right) of the critical point.
• Sketch the pressure- volume (PV) diagram for a pure component illustrating the
behavior above the bubble point, between the bubble and dewpoint and below
the dewpoint.
• Sketch a series of PV lines for a pure component with a temperature below, at
and above the critical temperature.
• Sketch the three dimensional phase diagram for pure component systems.

Two Components
• Plot a PV diagram for a 2 component system and identify key parameters.
• Plot a PV diagram for a 2 component system and identify key parameters and
the relationship to the vapour pressure lines for the two pure components.
• Sketch the critical point loci for a series of binary mixtures including methane
and indicate how a mixture a mixture of methane and another component can
exist as 2 phases at pressures much greater than the 2 phase limit for the two
contributing components.
• Draw a PT diagram for a two component system, to illustrate the
cricondentherm, cricondenbar and the region of retrograde condensation.
• Define the terms cricondentherm .
• Explain briefly what retrograde condensation is.

Multicomponent Systems
• Sketch a PT and PV diagrams to illustrate the behaviour at constant temperature
for a fluid in a PVT cell. Identify key features.
• Draw a PT diagram for a heavy oil, volatile oil, retrograde condensate gas, wet
gas and dry gas. Illustrate and explain the behaviour of depletion from the
undersaturated condition to the condition within the phase diagram.

©HERIOT-W ATT UNIVERSITY June 2014 v1


54

• Describe briefly with the aid of a sketch, the reasons for and the process of gas
cycling, for retrograde gas condensate reservoirs.
• Plot a PT diagram for a reservoir with a gas cap to illustrate the gas at dew point
and oil at bubble point.

Miscellaneous
• With the aid of sketch explain the process of critical point drying.

INTRODUCTION

Oil and gas reservoir fluids are mixtures of a large number of components which when
subjected to different pressure and temperature environments may exist in different
forms, which we call phases. Phase behaviour is a key aspect in understanding the
nature and behaviour of these fluids both in relation to their state in the reservoir and
the changes which they experience during various aspects of the production process. In
this chapter we will review the qualitative aspects of the behaviour of reservoir fluids
when subjected to changes in pressure and temperature.

1 DEFINITIONS

Before we consider the effect of temperature and pressure on hydrocarbon systems we


will define some terms:

• System - amount of substances within given boundaries under specific


conditions composed of a number of components. Everything within these
boundaries are part of the system and that existing outside of the boundaries are
not part of the system. If anything moves across these boundaries then the
system will have changed.

• Components - those pure substances which produce the system under all
conditions.

For example, in the context of reservoir engineering, methane, ethane, carbon dioxide
and water are examples of pure components.

• Phases - This term describes separate, physically homogenous parts which are
separated by definite boundaries.1 Examples in the context of water are the three
phases, ice, liquid water and water vapour.

©HERIOT-W ATT UNIVERSITY June 2014 v1


55

• Equilibrium - When a system is in equilibrium then no changes take place


with respect to time in the measurable physical properties of the separate
phases.

• Intensive and extensive properties - physical properties are termed either


intensive or extensive. Intensive properties are independent of the quantity of
material present. For example density, specific volume and compressibility
factor are intensive properties whereas properties such as volume and mass are
termed extensive properties; their values being determined by the total quantity
of matter present.

The physical behaviour of hydrocarbons when pressure and temperature changes can
be explained in relation to the behaviour of the individual molecules making up the
system. Temperature, pressure and intermolecular forces are important aspects of
physical behaviour.

The temperature is an indication of the kinetic energy of the molecules. It is a


physical measure of the average kinetic energy of the molecules. The kinetic energy
increases as heat is added. This increase in kinetic energy causes an increase in the
motion of the molecules which also results in the molecules moving further apart.

The pressure reflects the frequency of the collision of the molecules on the walls of
its container. As more molecules are forced closer together the pressure increases.

Intermolecular forces are the attractive and repulsive forces between molecules. They
are affected by the distance between the molecules. The attractive forces increases as
the distance between the molecules decreases until however the electronic field of the
molecules overlap and then further decrease in distance causes a repulsive force,
which increases as the molecules are forced closer together.

The molecules in gases are widely spaced and attractive forces exist between the
molecules whereas for liquids where the molecules are closer together there is a
repelling force which causes the liquid to resist further compression.

The hydrocarbon fluids of interest in reservoir systems are composed of many


components however in understanding the phase behaviour of these systems it is
convenient to reflect on the behaviour of single and two component systems.

©HERIOT-W ATT UNIVERSITY June 2014 v1


56

2 PHASE BEHAVIOUR OF PURE SUBSTANCES

2.1 The Phase Diagram


It is beneficial to study the behaviour of a pure hydrocarbon under varying pressure
and temperature to gain an insight into the behaviour of more complex hydrocarbon
systems.

Phase diagrams are useful ways of presenting the behaviour of systems. They are
generally plots of pressure versus temperature and show the phases that exist under
these varying conditions.

Figure 1 gives a pressure - temperature phase diagram for a single-component system


on a pressure temperature diagram and the following points are to be noted.

Figure 1 Pressure temperature diagram for a single component system

Vapour Pressure Line


The vapour pressure line divides regions where the substance is a liquid, 2, from
regions where it is a gas, 3. Above the line indicates conditions for which a substance
is a liquid, whereas below the line represent conditions under which it is a gas.
Conditions on the line indicate where both liquid and gas phases coexist.

©HERIOT-W ATT UNIVERSITY June 2014 v1


57

Critical Point
The critical point C. is the limit of the vapour pressure line and defines the critical
temperature, Tc and critical pressure, Pc of the pure substance. For a pure
substance the critical temperature and critical pressure represents the limiting state for
liquid and gas to coexist. A more general definition of the critical point which is both
applicable to multi component as well as single component systems is: the critical
point is the point at which all the intensive properties of the gas and liquid are equal.

Triple Point
The triple point represents the pressure and temperature at which solid, liquid and
vapour co-exist under equilibrium conditions. Petroleum engineers seldom deal with
hydrocarbons in the solid state, however, more recently solid state issues are a concern
with respect to wax, asphaltenes and hydrates.

Sublimitation-Pressure Line
The extension of the vapour-pressure line below the triple point represents the
conditions which divides the area where solid exists from the area where vapour exists
and is also called the sublimation - pressure line.

Melting Point Line


The melting line divides solid from liquid. For pure hydrocarbons the melting point
generally increases with pressure so the slope of the line is positive. (Water is
exceptional in that its melting point decreases with pressure).

3 USE OF PHASE DIAGRAMS

3.1 Pressure -Temperature Diagrams (PT)


Consider the behaviour of a cell charged with a pure substance and the volume varied
by the frictionless displacement of a piston as shown in Figure 2, below. This
represents the actual cells used in laboratories to measure the variation in pressure
temperature and volume. These PVT measurements represent the nature of the
changes in pressure temperature and volume of the fluids as they move from the
reservoir up the tubing string, into the separator and finally to the stock tank.

©HERIOT-W ATT UNIVERSITY June 2014 v1


58

Figure 2 Phase Changes With Pressure at Constant Temperature

For example, following the path 1 - 2 in Figure 3 on the pressure-temperature diagram,


ie holding temperature constant and varying pressure by expansion of the cylinder.

Figure 3 Pressure-Temperature Diagram for a Single-Component System

©HERIOT-W ATT UNIVERSITY June 2014 v1


59

As the pressure is reduced, the pressure falls rapidly until a pressure is reached lying
on the vapour pressure line. A gas phase will begin to form and molecules leave the
liquid. At further attempts to reduce the pressure the volume of gas phase increases,
while liquid phase volume decreases but the pressure remains constant. Once the
liquid phase disappears further attempts to reduce pressure will be successful as the
gas expands.

Above the critical temperature, following the path 3 - 4, a decrease in pressure will
cause a steady change in the physical properties, for example a decrease in density but
there will not be an abrupt density change as the vapour pressure line is not crossed.
No phase change takes place.

Consider the behaviour of the system around the critical point. If we go from point A
to point B, by increasing the temperature, we go though a distinctive phase change on
the vapour pressure line where two phases, liquid and gas co-exist. If we now go a
different route to B, starting with the liquid state at ‘A’ increase the pressure
isothermally (constant temperature ) to a value greater than Pc at E. Then keeping the
pressure constant increase the temperature to a value greater than T c at point F. Now
decrease the pressure to its original value at G. Finally, decrease the temperature
keeping the pressure constant until B is reached. The system is now in the vapour
state and this state has been achieved without an abrupt phase change. The vapour
states are only meaningful in the two phase regions. In areas far removed from the
two phase region particularly where pressure and temperature are above the critical
values, definition of the liquid or gaseous state is impossible and the system is best
described as in the fluid state.

3.2 Pressure Volume Diagram (PV)


The process just described in 3.1 can also be represented on a pressure-volume
diagram at constant temperature (Figure 5). As the pressure is reduced from 1, a large
change in pressure occurs with small change in volume due to the relatively low
compressibility of the liquid. When the vapour pressure is reached gas begins to form.
This point is called the bubble point, ie the point at which the first few molecules leave
the liquid and form small bubbles of gas. As the system expands more liquid is
vaporised at constant pressure. The point at which only a minute drop of liquid
remains is called the dew point. Sharp breaks in the line denote the bubble point and
dew point.

©HERIOT-W ATT UNIVERSITY June 2014 v1


60

Figure 5 Pressure-Volume diagram for a Single-Component System

For a pure substance vapour pressures at bubble point and dew point are equal to the
vapour pressure of the substance at that temperature. Above the critical point, ie
3 - 4, the PV behaviour line shows no abrupt change and simply shows an expansion
of the substance and no phase change. This fluid is called a super critical fluid.

A series of expansions can be performed at various constant temperatures and a


pressure volume diagram built up and the locus of the bubble point and dew point
values gives the bubble point and dew point lines which meet at the critical point.
Conditions under the bubble point and dew point lines represent the conditions where
two phases coexist whereas those above these curves represent the conditions where
only one phase exists. At the critical temperature the P,T curve goes through the
critical point, Figure 6

©HERIOT-W ATT UNIVERSITY June 2014 v1


61

Figure 6 Series of PV lines for a pure component

The pressure volume curve for pure component ethane is given in Figure 7

The locus of the bubble points and dew points form a three-dimensional diagram and
when projected in to a P-T diagram give the vapour pressure line (Figure 8).

©HERIOT-W ATT UNIVERSITY June 2014 v1


62

Figure 7 Pressure-Volume Diagram of Ethane

Figure 8 Three Dimensional Phase Diagram for a Pure Component System

4 TWO COMPONENT SYSTEMS


Reservoir fluids contain many components but we will first consider a system
containing two components, such a system is called a binary.

4.1 Pressure Volume Diagram


The behaviour of a mixture of two components is not as simple as for a pure
substance. Figure 9 shows the P-V diagram of a two-component mixture for a
constant temperature system.

©HERIOT-W ATT UNIVERSITY June 2014 v1


63

Figure 9 Pressure-Volume Line for a Two-Component System at Constant Temperature

The isotherm is very similar to the pure component but the pressure increases as the
system passes from the dew point to the bubble point. This is because the composition
of the liquid and vapour changes as it passes through the two-phase region. At the
bubble point the composition of the liquid is essentially equal to the composition of
the mixture but the infinitesimal amount of gas is richer in the more volatile
component. At the dew point the composition of vapour is essentially the mixture
composition whereas the infinitesimal amount of liquid is richer in the less volatile
component. Breaks in the line are not as sharp as for pure substances.

The pressure-volume diagram for a specific n-pentane and n-heptane mixture is given
in Figure 10. Clearly a different composition of the two components would result in a
different shape of the diagram.

©HERIOT-W ATT UNIVERSITY June 2014 v1


64

Figure 10 Pressure-Volume Diagram for N-Pentane and N-Heptane (52.4 mole % Heptane)
ref. 4

4.2 Pressure Temperature Diagram


Compared to the single line representing the vapour pressure curve for pure substances
there is a broad region in which the two phases co-exist. The two-phase region of the
diagram is bounded by the bubble point line and the dew point line, and the two
lines meet at the critical point. Points within a loop represent two-phase systems (Figure
11).

Consider the constant temperature expansion of a particular mixture composition. At 1


the substance is liquid and as pressure is reduced liquid expands until the bubble point
is reached. The pressure at which the first bubbles of gas appear is termed the bubble
point pressure. As pressure is decreased liquid and gas co-exist until a minute amount
of liquid remains at the dew point pressure. Further reduction of pressure causes
expansion of the gas.

By carrying out a series of constant temperature expansions the phase envelope is


defined and within the envelope contours of liquid to gas ratios obtained. These are
called quality lines and describe the pressure and temperature conditions for equal
volumes of liquid. The quality lines converge at the critical point.

©HERIOT-W ATT UNIVERSITY June 2014 v1


65

4.3 Critical Point


In the same way as pure components, when more than one component is present
liquid and gases cannot coexist, at pressures and temperatures higher than the critical
point. The critical point for a more than one component mixture is defined as a point
at which the bubble point line and dew point line join, ie. it is also the point at which
all the intensive properties of the liquid are identical. This aspect is a very severe test
for physical property prediction methods.

If the vapour pressure lines for the pure components are drawn on the P-T diagram
then the two-phase region for the mixture lies between the vapour pressure lines. In
the Figure 11 the critical temperature of the mixture TcAB lies between TcA and TcB
whereas the critical pressure PcAB lies above PcA and PcB. It is important to note that
the PcAB and TcAB of the mixture does not necessarily lie between the Pc & Tc of the
two pure components.

Figure 11 Pressure-Temperature Diagram for a Two Component System

©HERIOT-W ATT UNIVERSITY June 2014 v1


66

A specific mixture composition will give a specific phase envelope lying between the
vapour pressure lines. A mixture with different proportions of the same components
will give a different phase diagram. The locus of the critical points of different mixture
compositions is shown in Figure 12 for the ethane and n-heptane system, and in Figure
13 for a series of binary hydrocarbon mixtures. Figure 13 demonstrates that for binary
mixture e.g. Methane and n-decane two phases can coexist at conditions of pressure
considerably greater than the two phase limit, (critical conditions) for the separate pure
components. Methane is a significant component of reservoir fluids.

Figure 12 Pressure-Temperature Diagram for the Ethane-Heptane System 2

©HERIOT-W ATT UNIVERSITY June 2014 v1


67

Figure 13 Critical Point Loci for a Series of Binary Hydrocarbon Mixtures 2

4.4 Retrograde Condensation


Within the two phase region our two component system there can be temperatures and
pressures higher than the critical temperature where two phases exist and similarly
pressures. These limiting temperatures and pressures are the cricondentherm and
cricondenbar . The cricondentherm can be defined as the temperature above which
liquid cannot be formed regardless of pressure, or expressed differently, as the
maximum temperature at which two phases can exist in equilibrium. The cricondenbar
can be defined as the pressure above which no gas can be formed regardless of
temperature or as the maximum pressure at which two phases can exist in equilibrium.
(Figure 14).

These limits are of particular significance in relation to the shape of the diagram in
Figure 14.

©HERIOT-W ATT UNIVERSITY June 2014 v1


68

Consider a single isotherm on Figure 14. For a pure substance a decrease in pressure
causes a change of phase from liquid to gas. For a two-component system below Tc a
decrease in pressure causes a change from liquid to gas.

We now consider the constant temperature decrease in pressure, 1-2-3 , in Figure 14


at a temperature between the critical temperature and the cricondentherm. As pressure
is decreased from 1 the dew point is reached and liquid forms, i.e., at 2 the system is
such that 5% liquid and 95% vapour exists, i.e. a decrease in pressure has caused a
change from gas to liquid, opposite to the behaviour one would expect. The
phenomena is termed Retrograde Condensation. From 2 - 3, the amount of liquid
decreases and vaporisation occurs and the dew point is again reached where the system
is gas. Retrograde condensation occurs at temperatures between the critical
temperature and cricondentherm. The retrograde region is shown shaded in the Figure.

Figure 14 Phase Diagram Showing Conditions for Retrograde Considerations

5. MULTI-COMPONENT HYDROCARBON

Using two component systems we have examined various aspects of phase behaviour.
Reservoir fluids contain hundreds of components and therefore are multicomponent
systems. The phase behaviour of multicomponent hydrocarbon systems in the liquid-
vapour region however is very similar to that of binary systems however the

©HERIOT-W ATT UNIVERSITY June 2014 v1


69

mathematical and experimental analysis of the phase behaviour is more complex.


Figure 15 gives a schematic PT & PV diagram for a reservoir fluid system. Systems
which include crude oils also contain appreciable amounts of relatively non-volatile
constituents such that dew points are practically unattainable.

Figure 15 Phase Diagrams for Multicomponent Systems

We will consider the behaviour of several examples of typical crude oils and natural
gases:
 Low-shrinkage oil (heavy oil - black oil)
 High-shrinkage oil (volatile oil)
 Retrograde condensate gas
 Wet gas
 Dry Gas

Figure 16 is a useful diagram to illustrate the behaviour of the respective fluid types
above. However it should be emphasised that for each fluid type there will be different
scales. The vertical lines help to distinguish the different reservoir fluid types.

Isothermal behaviour below the critical point designates the behaviour of oil systems
and the fluid is liquid in the reservoir, whereas behaviour to the right of the critical
point illustrates the behaviour of systems which are gas in the reservoir.

©HERIOT-W ATT UNIVERSITY June 2014 v1


70

Figure 16 Phase diagram for reservoir fluids

5.1 Oil Systems ( Black Oils and Volatile Oils)


Figures 17&18 illustrate the PT phase diagrams for black and volatile oils.

The two-phase region covers a wide range of pressure and temperature. Tc is higher
than the reservoir temperature. In Figure 17 the line 1-2-3 represents the constant
reservoir temperature pressure reduction that occurs in the reservoir as crude oil is
produced for a black oil. These oils are a common oil type. The dotted line shows the
conditions encountered as the fluid leaves the reservoir and flows through the tubing to
the separator.

If the initial reservoir pressure and temperature are at 2, the oil is at its reservoir
bubble point and is said to be saturated, that is, the oil contains as much dissolved
gas as it can and a further reduction in pressure will cause formation of gas. If the
initial reservoir pressure and temperature are at 1, the oil is said to be undersaturated,
i.e. the pressure in the reservoir can be reduced to Pb before gas is released into the
formation. For an oil system the saturation pressure is the bubble point pressure.

©HERIOT-W ATT UNIVERSITY June 2014 v1


71

Figure 17 Phase Diagram for a Black Oil

As the pressure is dropped from the initial condition as a result of production of fluids,
the fluids remain in single phase in the reservoir until the bubble point pressure
corresponding to the reservoir temperature is reached. At this point the first bubbles of
gas are released and their composition will be different from the oil being more
concentrated in the lighter ( more volatile) components. When the fluids are brought
to the surface they come into the separator and as shown on the diagram, the separator
conditions lie well within the two phase region and therefore the fluid presents itself as
both liquid and gas.

Volatile oil contains a much higher proportion of lighter and intermediate


hydocarbons than heavier black oil and therefore they liberate relatively large volumes
of gas leaving smaller amounts of liquid compared to black oils. For this reason they
used to be called high shrinkage oils. The diagram in Figure 18 shows similar
behaviour to the black oil except that the lines of constant liquid to gas are more
closely spaced.

Points 1 and 2 have the same meaning as for the black oil. As the pressure is reduced
below 2 a large amount of gas is produced such that at 3 the reservoir contains 40%
liquid and 60% gas.

At separator conditions 65% of the fluid is liquid, i.e. less than previous mixture.

©HERIOT-W ATT UNIVERSITY June 2014 v1


72

Figure 18 Phase Diagram for a Volatile Oil

For both of these fluids types one can prevent the reservoir fluid going two phase by
maintaining the reservoir pressure above its saturation pressure by injecting fluids into
the reservoir. The most common practise is the use of water as a pressure maintenance
fluid.

5.2 Retrograde Condensate Gas


If the reservoir temperature lies between the critical point and the cricondentherm a
retrograde gas condensate field exists and Figure 19 gives the PT diagram for such a
fluid. Above the phase envelope a single phase fluid exists. As the pressure declines
to 2 a dew point occurs and liquid begins to form in the reservoir. The liquid is richer
in heavier components than the associated gas. As the pressure reduces to 3 the
amount of liquid increases. Further pressure reduction causes the reduction of liquid
in the reservoir by re-vaporisation. It is important to recognise that the phase diagram
below for a retrograde condensate fluid represents the diagram for a constant
composition system.

Before production the fluid in the reservoir exists as a single phase and is generally
called a gas. It is probably more accurate to call it a dense phase fluid. If the
reservoir drops below the saturation pressure the dew point, then retrograde
condensation occurs within the formation. The nature of this condensing fluid is only
in recent years being understood. It was previously considered that the condensing

©HERIOT-W ATT UNIVERSITY June 2014 v1


73

fluid would be immobile since its maximum proportion was below the value for it to
have mobility. It was considered therefore that such valuable condensed fluids would
be lost to production and the viability of the project would be that from the ‘wet’ gas.

Figure 19 Phase Diagram for a Retrograde Condensate Gas

One of the development options for such a field therefore is to set in place a pressure
maintenance procedure whereby the reservoir pressure does not fall below the
saturation pressure. Water could be used as for oils but gas might be trapped behind
the water as the water advances through the reservoir. Gas injection, called gas
cycling ( Figure 20 ), is the preferred yet very expensive option. In this process the
produced fluids are separated at the surface and the liquid condensates, high value
product relative to heavy oil, are sent for export, in an offshore situation probably by
tanker. The ‘dry’ gas is then compressed and reinjected into the reservoir to maintain
the pressure above the dew point.

©HERIOT-W ATT UNIVERSITY June 2014 v1


74

Figure 20 Gas cycling process

Looking at the PT phase diagram one might consider that "blowing the reservoir
down" quickly might be an option and as a result vaporise the condensed liquids in the
formation. This is not a serious option since once the reservoir pressure falls below
the dew point the impact of the increasing liquid proportion remaining in the reservoir
causes the phase diagram to move to the right relative to reservoir conditions, and any
vaporising will be of the lightest components which are likely to be in good supply and
therefore not of significant value.

5.3 Wet Gas


The phase diagram for a mixture containing smaller hydrocarbon molecules lies well
below the reservoir temperature, Figure 21. The reservoir conditions always remain
outside the two-phase envelope going from 1 to 2 and therefore the fluid exists as a
gas throughout the reduction in reservoir pressure. For a wet gas system, the separator
conditions lie within the two-phase region, therefore at surface heavy components
present in the reservoir fluid condense under separator conditions and this liquid is
normally called condensate.

©HERIOT-W ATT UNIVERSITY June 2014 v1


75

Figure 21 Phase Diagram for a Wet Gas

The reference wet gas, clearly does not refer to the system being wet due to the
presence of water but due to the production condensate liquids.

5.4 Dry Gas


The phase envelope of the dry gas, which contains a smaller fraction of the C2-C6
components, is similar to the wet gas system but with the distinction that the separator
also lies outside the envelope in the gas region (Figure 22). The term dry indicates
therefore that the fluid does not contain enough heavier HC’s to form a liquid at
surface conditions.

©HERIOT-W ATT UNIVERSITY June 2014 v1


76

Figure 22 Phase Diagram for a Dry Gas

6 RESERVOIRS WITH A GAS CAP

Figure 23 illustrates a simplification of the phase diagrams associated with an oil


reservoir with a gas cap. The phase diagram for the gas cap fluid, the oil reservoir
fluid and for a fluid representing the combination fluid of a mixture of gas and liquid
in the same proportions as they exist in the reservoir are presented.

©HERIOT-W ATT UNIVERSITY June 2014 v1


77

Figure 23 Phase Diagram for an Oil Reservoir with a Gas Cap

The diagram illustrates that at the gas-oil contact the gas is at its dew pressure, the oil
is at its bubble point pressure and the combination fluid lies on the constant proportion
quality line representing the ratio of the gas and oil as they exist in the reservoir
system. The gas cap may be dry, wet or condensate depending on the composition and
phase diagram of the gas.

REFERENCES

1. Figure 1 Daniels, F Farrington: “Outlines of Physical Chemistry,” John Wiley &


Sons,Inc New York,

2. Figure 2 Brown,GG et al. “ Natural Gasoline and Volatile Hydrocarbons,”


Natural Gasoline Association of America, Tulsa, Okl.
Figure 10 Sage, S.G.,Lacy,W.N. Volumetric and Phase Behaviour of
Hydrocarbons, Gulf Publishing Co.Houston

©HERIOT-W ATT UNIVERSITY June 2014 v1


78

TOPIC 5 BEHAVIOUR OF GASES

1 IDEAL GASES
1.1 Boyle's Law
1.2 Charles' Law
1.3 Avogadro's Law
1.4 The Equation of State For an Ideal Gas
1.5 The Density of an Ideal Gas
1.6 Standard Conditions
1.7 Mixtures of Ideal Gases
1.7.1 Dalton's Law of Partial Pressures
1.7.2 Amagat's Law
1.8 Apparent Molecular Weight
1.9 Specific Gravity of a Gas

2 BEHAVIOUR OF REAL GASES


2.1 Compressibility Factor For Natural Gases
2.2 Law of Corresponding States
2.3 Pseudocritical Properties of Natural Gases
2.4 Impact of Nonhydrocarbon Components on z Value
2.5 Standard Conditions For Real Reservoir Gases

3 GAS FORMATION VOLUME FACTOR

4 COEFFICIENT OF ISOTHERMAL COMPRESSIBILITY OF GASES

5 VISCOSITY OF GASES

LEARNING OUTCOMES

Having worked through this chapter the Student will be able to:

• Present the ideal equation of state, PV=nRT.


• Calculate the mass of an ideal gas given PV &T values.
• Derive an equation to calculate the density of an ideal gas.
• Convert a mixture composition between weight and mole fraction.
• Present an equation and calculate the apparent molecular weight of a mixture.
• Define and calculate the specific gravity of a gas.
• Present the equation of state, EOS, for a ‘real gas’ and explain what ‘Z’ is,

©HERIOT-W ATT UNIVERSITY June 2014 v1


79

PV=ZnRT.
• Define the pseudocritical pressure and psuedocritical temperature and be
able to use them to determine the ‘Z’ value for a gas mixture.
• Express and calculate reservoir gas volumes in terms of standard cubic
volumes.
• Define the gas formation volume factor and derive an equation for it using
the EOS.
• Calculate the volume of gas in a reservoir in terms of standard cubic volumes
given prerequisite data.

INTRODUCTION

A gas is a homogenous fluid that has no definite volume but fills completely the vessel
in which it is placed. The system behaviour of gases is vital to petroleum engineers
and the laws governing their behaviour should be understood. For simple gases these
laws are straightforward but the behaviour of actual hydrocarbon gases particularly at
the conditions occurring in the reservoir are more complicated.

We will review the laws that relate to the pressure, volume and temperatures of gases
and the associated equations. These relationships were previously termed gas laws; it
is now more common to describe them as equations of state.

1 IDEAL GASES

The laws relating to gases are straightforward in that the relationships of pressure,
temperature and pressure are covered by one equation. First consider an ideal gas. An
ideal gas is one where the following assumptions hold:

• Volume of the molecules is insignificant with respect to the total volume of the gas.
• There are no attractive or repulsive forces between molecules or between
molecules and container walls.
• There is no internal energy loss when molecules collide.

Out of these assumptions come the following equations.

1.1 Boyle’s Law


At constant temperature the pressure of a given weight of a gas is inversely
proportional to the volume of a gas.

©HERIOT-W ATT UNIVERSITY June 2014 v1


80

i.e.
(1)

P = pressure, V = volume, T = temperature.

1.2 Charles’ Law


At constant pressure, the volume of a given weight of gas varies directly with the
temperature:

i.e.
(2)

The pressure and temperature in both laws are in absolute units.

1.3 Avogadro’s Law


Avogadro’s Law can be stated as: under the same conditions of temperature and
pressure equal volumes of all ideal gases contain the same number of molecules. That
is, one molecular weight of any ideal gas occupies the same volume as the molecular
weight of another ideal gas at a given temperature and pressure.

Specifically, these are:

(i) 2.73 x 1026 molecules/lb mole of ideal gas.


(ii) One molecular weight (in lbs) of any ideal gas at 60˚F and 14.7 psia
occupies a volume of 379.4 cu ft.

One mole of a material is a quantity of that material whose mass in the unit system
selected is numerically equal to the molecular weight.

eg. one lb mole of methane CH4 = 16 lb


one kg mole of methane CH4 = 16kg

1.4 The Equation of State for an Ideal Gas


By combining the above laws an equation of state relating pressure, temperature and
volume of a gas is obtained.

©HERIOT-W ATT UNIVERSITY June 2014 v1


81

PV
= constant
T (3)

R is the constant when the quantity of gas is equal to one mole.

It is termed the Universal Gas Constant and has different values depending on the
unit system used, so that;

R in oilfield units =
cu ft psia
10.732
lb mole R

Table 1 gives the values for different unit systems.

Table 1 Values of R for different unit systems.

For n moles the equation becomes:

PV = nRT (4)

T= absolute temperature oK or oR where


ºK=273 +oC and oR=460 +oF

To find the volume occupied by a quantity of gas when the conditions of temperature
and pressure are changed from state 1 to state 2 we note that:

PV PV PV
n = is a constant so that 1 1 = 2 2
RT T1 T2

©HERIOT-W ATT UNIVERSITY June 2014 v1


82

1.5 The Density of an Ideal Gas


Since density is defined as the mass per unit volume, the ideal gas law can be used to
calculate densities. r r
m
 g  mass / volume 
V

where g is the gas density


For 1 mole m = MW MW = Molecular weight

(5)

1.6 Standard Conditions


Oil and gas at reservoir conditions clearly occur under a whole range of temperatures
and pressures.

It is common practice to relate volumes to conditions at surface, ie 14.7 psia and 60˚F.
ie

Pres Vres P V
= sc sc
Tres Tsc (6)

sc - standard conditions res - reservoir conditions

This relationship assumes that reservoir properties behave as ideal. This is NOT the
case as will be discussed later.

1.7 Mixtures of Ideal Gases


Petroleum engineering is concerned not with single component gases but mixtures of a
number of gases.

Laws established over early years governing ideal gas mixtures include Dalton’s Law
and Amagat’s Law.

©HERIOT-W ATT UNIVERSITY June 2014 v1


83

1.7.1 Dalton’s Law of Partial Pressures


The total pressure exerted by a mixture of gases is equal to the sum of the pressures
exerted by its components. The partial pressure is the contribution to pressure of the
individual component.

Consider a gas made up of components A, B, C etc


The total pressure of the system is the sum of the partial pressures

ie

P = PA + PB + PC + ..... (7)

where A, B and C are components.

therefore

(8)

where yj = mole fraction of jth component.

The pressure contribution of a component, its partial pressure, is the total pressure
times the mole fraction.

1.7.2 Amagat’s Law


Amagat’s Law states that the volume occupied by an ideal gas mixture is equal to the
sum of the volumes that the pure components would occupy at the same temperature
and pressure. Sometimes called the law of additive volumes.

i.e.

©HERIOT-W ATT UNIVERSITY June 2014 v1


84

V = VA + VB + VC (9)

(10)

i.e, for an ideal gas the volume fraction is equal to the mole fraction.

It is conventional to describe the compositions of hydrocarbon fluids in mole terms.


This is because of the above laws. In some circumstances however weight
compositions might be used as the basis and it is straight forward to convert between
the two.

1.8 Apparent Molecular Weight


A mixture does not have a molecular weight although it behaves as though it had a
molecular weight. This is called the apparent molecular weight. AMW

If yj represents the mole fraction of the jth component:

AMW for air = 28.97, a value of 29.0 is usually sufficiently accurate.

1.9 Specific Gravity of a Gas


The specific gravity of a gas, g is the ratio of the density of the gas relative to that of
dry air at the same conditions.

(11)

©HERIOT-W ATT UNIVERSITY June 2014 v1


85

Assuming that the gases and air are ideal.

Mg = AMW of mixture, Mair = AMW of air.

2 BEHAVIOUR OF REAL GASES

The equations so far listed apply basically to ideal systems. In reality, however,
particularly at high pressures and low temperatures the volume of the molecules are no
longer negligible and attractive forces on the molecules are significant.

The ideal gas law, therefore, is not too applicable to light hydrocarbons and their
associated fluids and it is necessary to use a more refined equation.

There are two general methods of correcting the ideal gas law equation:

(1) By using a correction factor in the equation PV = nRT


(2) By using another equation-of-state

2.1 Compressibility Factor for Natural Gases


The correction factor ‘z’ which is a function of the gas composition, pressure and
temperature is used to modify the ideal gas law to:

PV = znRT (12)

where the factor ‘z’ is known as the compressibility factor and the equation is known
as the compressibility equation-of-state or the compressibility equation.

The compressibility factor is not a constant but varies with changes in gas
composition, temperature and pressure and must be determined experimentally (Figure
1).

To compare two states the law now takes the form:

P1V1 PV
= 2 2
z1T1 z 2 T2 (13)

©HERIOT-W ATT UNIVERSITY June 2014 v1


86

z is an expression of the actual volume to what the ideal volume would be.

i.e.

Vactual
z =
Videal (14)

Figure 1 Typical plot of the compressibility factor as a function of pressure at constant


temperature.

Although all gases have similar shapes with respect to z the actual values are
component specific. However through the law of corresponding states all pure gases
are shown to have common values.

2.2 Law of Corresponding States


The law of corresponding states shows that the properties of many pure liquids and
gases have the same value at the same reduced temperature (T r) and pressure (Pr)
where:

T P
Tr = and Pr =
Tc Pc (15)

Where, Tc and Pc are the pure component critical temperature and pressure.

©HERIOT-W ATT UNIVERSITY June 2014 v1


87

The compressibility factor ‘z’ follows this law. It is usually presented vs T r and Pr.
Although in many cases pure gases follow the Law of Corresponding States, the gases
associated with hydrocarbon reservoirs do not. The Law has however been used to
apply to mixtures by defining parameters called pseudo critical temperature and
pseudocritical pressure.

For mixtures a pseudocritical temperature and pressure, Tpc and Ppc is used such that:

(16)

where y is the mole fraction of component j and Tcj and Pcj are the critical temperature
and pressure of component j.

It should be emphasised that these pseudo critical temperature and pseudocritical


pressures are not the same as the real critical temperature and pressure. By
definition the pseudo values must lie between the extreme critical values of the pure
components whereas the actual critical values for mixtures can be outside these limits,
as was observed in the Phase Behaviour chapter.

For mixtures the compressibility factor (z) has been generated with respect to natural
gases 1, where ‘z’ is plotted as a function of pseudo reduced temperature, T pr and
pseudo reduced pressure Ppr where

©HERIOT-W ATT UNIVERSITY June 2014 v1


88

Figure 2 Compressibility factors for natural gas1 (Standing & Katz, Trans AIME).

(17)
T P
Tpr = and Ppr =
Tpc Ppc

©HERIOT-W ATT UNIVERSITY June 2014 v1


89

The use of this chart, Figure 2, has become common practise to generate z values for
natural gases. Poettmann and Carpenter 2 have also converted the chart to a table.
Various equations have also been generated based on the tables.

2.3 Pseudocritical Properties of Natural Gases


The pseudocritical properties of gases can be computed from the basic composition
but can also be estimated from the gas gravity using the correlation presented in
Figure 3.

Figure 3 Pseudocritical properties of natural gases. 3

2.4 Impact of Nonhydrocarbon Components on z value.

©HERIOT-W ATT UNIVERSITY June 2014 v1


90

Components like hydrogen sulphide, and carbon dioxide have a significant impact on
the value of z. If the method previously applied is used large errors in z result.
Wichert and Aziz 4 have produced an equation which enables the impact of these two
gases to be calculated.

2.5 Standard Conditions for Real Reservoir Gases


It is convenient to describe the quantity of gas to a common basis and this is termed
the standard conditions, giving rise to the standard cubic foot and the standard cubic
metre. The petroleum engineer is primarily interested in volume calculations for
gaseous mixtures. Throughout the industry gas volumes are measured at a standard
temperature of 60˚F (15.6˚C) and at a pressure of 14.7 psia (one atmosphere). These
conditions are referred to as standard temperature and pressure STP. Standard
Cubic Feet, the unit of volume measured under these conditions is sometimes
abbreviated SCF or scf (SCM is Standard Cubic Metres). It is helpful to consider these
expressions not as volumes but as an alternate expression of the quantity of material.
For example a mass of gas can be expressed as so many standard cubic feet or metres.

3 GAS FORMATION VOLUME FACTOR

The petroleum industry expresses its reservoir quantities at a common basis of surface
conditions which for gases is standard cubic volumes. To convert reservoir volumes
to surface volumes the industry uses formation volume factors. For gases we have Bg,
the gas formation volume factor, which is the ratio of the volume occupied at
reservoir temperature and pressure by a certain weight of gas to the volume occupied
by the same weight of gas at standard conditions. The shape of Bg as a function of
pressure is shown in Figure 4.

volume occupied at reservoir temperature and pressure


Bg =
volume occupied at STP

The gas formation volume factor can be obtained from PVT measurements on a gas
sample or it may be calculated from the equations-of-state discussed previously.

One definition of the gas formation volume factor is: it is the volume in barrels that
one standard cubic foot of gas will occupy as free gas in the reservoir at the
prevailing reservoir pressure and temperature.

Depending on the definition the units will change and the units will be; rb reservoir
barrels free gas/scf gas or rm3 reservoir cubic meters free gas/scm gas

©HERIOT-W ATT UNIVERSITY June 2014 v1


91

Figure 4 Gas Formation Volume Factor, Bg.

For example Bg for a reservoir at condition 2 is;

V2 Psc T2 z2
Bg = =
Vsc P2 Tsc zsc (20)

‘sc’ refers to standard conditions. z at standard conditions is taken as 1.0

The reciprocal of Bg is often used to calculate volumes at surface so as to reduce the


possibility of misplacing the decimal point associated with the values of Bg being less
than 0.01, ie:

E is sometimes referred to as the expansion factor.

Usually the units of Bg are barrels of gas at reservoir conditions per standard cubic foot
of gas, ie bbl/SCF or cubic metres per standard cubic metre.

VR
Bg =
Vsc (21)

Subscripts R and sc are reservoir and standard conditions respectively.

©HERIOT-W ATT UNIVERSITY June 2014 v1


92

znRT
VR =
P (22)

T and P at reservoir conditions:

zsc nRTsc
Vsc =
Psc (23)

z = 1 for standard conditions

(24)

Since Tsc = 520˚Rm Psc = 14.7 psia for most cases

zT cu. ft
Bg = 0.0283
P SCF

(25)
zT cu. ft bbl
Bg = 0.0283 ´
P SCF 5.615 cu ft
or
zT res bbl
Bg = 0.00504
P SCF

4 COEFFICIENT OF ISOTHERMAL COMPRESSIBILITY OF GASES

The compressibility factor, z, must not be confused with the compressibility which is
defined as the change in volume per unit volume for a unit change in pressure, or

(26)

Vm is the specific volume or volume per mole.

©HERIOT-W ATT UNIVERSITY June 2014 v1


93

cg is not the same as z, the compressibility factor.

For an ideal gas:

PV = nRT or:

(27)

For real gases:

znRT
V =
P

(28)

z/P can be obtained from the slope of the z vs P curve.

The Law of Corresponding states can be used to express the above equation in another
form

©HERIOT-W ATT UNIVERSITY June 2014 v1


94

Combining this equation with eqn 28 above yields

(29)

Units of cg = P-1, and cgPc is dimensionless

cgPpc is called pseudo reduced compressibility, cpr

Since the pseudo reduced compressibility is a function of ‘z’ and pseudo reduced
pressure, the graph of Figure 2 can be used with Equation 29 to calculate values of cpr.

5 VISCOSITY OF GASES

Viscosity is a measure of the resistance to flow. It is given in units of centipoise. A


centipoise is a gm/100 sec.cm. The viscosity term is called dynamic viscosity.

©HERIOT-W ATT UNIVERSITY June 2014 v1


95

Gas viscosity reduces as the pressure is decreased. At low pressures an increase in


temperature increases gas viscosity whereas at high pressures gas viscosity decreases
as the temperature increases. Figure 5 gives the values for pure component ethane.

Figure 5 Viscosity of ethane.

REFERENCES

1. Standing MB and Katz DL Density of Natural Gases. Trans AIME, 146

2. Poettmann FH and Carpenter PG The Multiphase Flow of Gas and Water


through Vertical Flow Strings with Application to the Design of Gas Lift
Installations. API Drilling and Production Practise.

3. Brown GG et al. Natural Gasoline and Volatile Hydrocarbons” National


Gasoline Assoc. of America, Tulsa, Okl.

4. Wichert, E and Aziz,K “ Calculate Z’s for sour gases” Hyd Proc. 51

5. Katz, D.L., Handbook of Natural Gas Engineering, McGraw Hill, NY

6. Carr N et al. Viscosity of natural gases under pressure. Trans AIME 201, 264,

7. Lee et al “The viscosity of natural gases.” Trans AIME 237, 997-1000

©HERIOT-W ATT UNIVERSITY June 2014 v1


96

8. Pitzer K S et al The Volumetric and Thermodynamic Properties of Fluids II.


Compressibility Factor, Vapour Pressure and Entropy of Vaporisation.
J. Am.Chem. Soc. 77, No 13,3433-3440

9. Danesh, A PVT and Phase Behaviour of Petroleum Reservoir Fluids.


Elsevier ISBN:0 444 82196 1 p129-162

©HERIOT-W ATT UNIVERSITY June 2014 v1


97

TOPIC 6 PROPERTIES OF RESERVOIR


LIQUIDS

1 COMPOSITION BLACK OIL MODELS

2 GAS SOLUBILITY, RS

3 OIL FORMATION VOLUME FACTOR, BO

4 TOTAL FORMATION VOLUME FACTOR, BT

5 BELOW THE BUBBLE POINT

6 OIL COMPRESSIBILITY

7 FLUID DENSITY
7.1 Specific Gravity of a Liquid
7.2 Density Based on Ideal Solution Principles

8 FORMATION VOLUME FACTOR OF GAS CONDENSATE, BGC

9 VISCOSITY OF OIL

10 COMPARISON OF RESERVOIR FLUID MODELS

LEARNING OUTCOMES

Having worked through this chapter the Student will be able to:

• Define gas solubility, Rs and plot vs. P for a reservoir fluid.


• Define undersaturated and saturated oil.
• Explain briefly flash and differential liberation
• Define the oil formation volume factor Bo, and plot Bo vs. P for a reservoir
fluid.
• Define the Total Formation Volume factor Bt, and plot Bt vs. P alongside a
Bo vs. P plot.

©HERIOT-W ATT UNIVERSITY June 2014 v1


98

• Present an equation to express Bt in terms of Bo, Rs and Bg.


• Express oil compressibility in terms of oil formation volume factor.
• Calculate the density of a reservoir fluid mixture, using ideal solution
principles, at reservoir pressure and temperature, using density correction chart
for C1& C2 and other prerequisite data.
• Define the formation volume factor of a gas condensate
• Calculate the reserves and production of gas and condensate operating above
the dewpoint, given prerequisite data.
• List the comparisons of the black oil and compositional model in predicting
liquid properties

©HERIOT-W ATT UNIVERSITY June 2014 v1


99

1 COMPOSITION - BLACK OIL MODEL

As introduced in the chapter on Composition, petroleum engineers require a


compositional description tool to use as a basis for predicting reservoir and well fluid
behaviour. The two approaches that are commonly used are the multicomponent
compositional model described in the earlier chapter and the two component black
oil model. The latter simplistic approach has been used for many years to describe the
composition and behaviour of reservoir fluids. It is called the “Black Oil Model”.

The black oil model considers the fluid being made up of two components - gas
dissolved in oil and stock tank oil. The compositional changes in the gas when
changing pressure and temperature are ignored. To those appreciating
thermodynamics this simplistic two component model is difficult to cope with. The
Black Oil Model, illustrated in Figure 1, is at the core of many petroleum engineering
calculations, and associated procedures and reports.

Associated with the black oil model are Black Oil model definitions in relation to Gas
Solubility and Formation Volume Factors.

Figure 1 Black Oil Model

©HERIOT-W ATT UNIVERSITY June 2014 v1


100

2 GAS SOLUBILITY

Although the gas associated with oil and the oil itself are multicomponent mixtures it
is convenient to refer to the solubility of gas in crude oil as if we were dealing with a
two-component system.

The amount of gas forming molecules in the liquid phase is limited only by the
reservoir conditions of temperature and pressure and the quantity of light components
present.

The solubility is referred to some basis and it is customary to use the stock tank barrel.

Solubility = f (pressure, temperature, composition of gas, composition of crude oil)

For a fixed gas and crude, at constant T, the quantity of solution gas increases with p,
and at constant p, the quantity of solution gas decreases with T.

Rather than determine the amount of gas which will dissolve in a certain amount of oil
it is customary to measure the amount of gas which will come out of solution as the
pressure decreases. Figure 2 illustrates the behaviour of an oil operating outside the PT
phase diagram in its single phase state when the reservoir pressure is above its
reservoir bubble point at 1. Fluid behaviour in the reservoir is single phase and the oil
is said to be undersaturated . In this case a slight reduction of pressure causes the
fluid to remain single phase. If the oil was on the boundary bubble point pressure line
at 2 then a further reduction in pressure would cause two phases to be produced, gas
and liquid. This saturated fluid is one that upon a slight reduction of pressure
releases some gas. The concept of gas being produced or coming out of solution gives
rise to this gas solubility perspective. Clearly when the fluids are produced to the
surface as shown by the undersaturated oil in Figure 2 the surface conditions lie within
the two phase area and gas and oil are produced. The gas produced is termed solution
gas and the oil at surface conditions stock tank oil. These are the two components
making up the reservoir fluid, clearly a very simplistic concept.

The gas solubility Rs is defined as the number of cubic feet (cubic metre) of gas
measured at standard conditions, which will dissolve in one barrel (cubic metre)
of stock tank oil when subjected to reservoir pressure and temperature. At initial
reservoir conditions, the subscript i is added and becomes Rsi.

In metric units the volumes are expressed as cubic metre of gas at standard conditions
which will dissolve in one cubic metre of stock tank oil.

©HERIOT-W ATT UNIVERSITY June 2014 v1


101

Figure 2 Production of reservoir hydrocarbons above bubble point

Figure 3 gives a typical shape of gas solubility as a function of pressure for a reservoir
fluid at reservoir temperature. When the reservoir pressure is above the bubble point
pressure then the oil is undersaturated, i.e. capable of containing more gas. As the
reservoir pressure drops gas does not come out of solution until the bubble point is
reached, over this pressure range therefore the gas in solution is constant. At the
bubble point pressure, corresponding to the reservoir temperature, two phases are
produced, gas and oil. The gas remaining in solution therefore decreases.

The nature of the liberation of the gas is not straight forward. Within the reservoir
when gas is released then its transport and that of the liquid is influenced by the
relative permeability of the rock. The gas does not remain with its associated oil i.e.
the system changes. In the production tubing and in the separator it is considered that
the gas and associated liquid remain together i.e. the system is constant. The amount
of gas liberated from a sample of reservoir oil depends on the conditions of the
liberation. There are two basic liberation mechanisms:

©HERIOT-W ATT UNIVERSITY June 2014 v1


102

Figure 3 Solution Gas - Oil Ratio as a Function of Pressure.

Flash liberation - the gas is evolved during a definite reduction


in pressure and the gas is kept in contact with the
liquid until equilibrium has been established.

Differential liberation - the gas being evolved is being continuously


removed from contact with the liquid and the
liquid is in equilibrium with the gas being evolved
over a finite pressure range.

The two methods of liberation give different results for Rs.

Production of a crude oil at reservoir pressures below the bubble point pressure occurs
by a process which is neither flash nor differential vaporisation. Once enough gas is
present for the gas to move toward the wellbore the gas tends to move faster than the
oil. The gas formed in a particular pore tends to leave the liquid from which it was
formed thus approximating differential vaporisation, however, the gas is in contact
with liquid throughout the path through the reservoir. The gas will also migrate
vertically as a result of its lower density than the oil and could form a secondary gas
cap.

Fluid produced from reservoir to the surface is considered to undergo a flash process
where the system remains constant.

©HERIOT-W ATT UNIVERSITY June 2014 v1


103

3 OIL FORMATION VOLUME FACTOR, B O

The volume occupied by the oil between surface conditions and reservoir or other
operating changes is that of the total system; the ‘stock tank oil’ plus its associated or
dissolved ‘solution gas’. The effect of pressure on the complex stock tank liquid and
the solution gas is to induce solution of the gas in the liquid until equilibrium is
reached. A unit volume of stock tank oil brought to equilibrium with its associated gas
at reservoir pressure and temperature will occupy a volume greater than unity (unless
the oil has very little dissolved gas at very high pressure).

The relationship between the volume of the oil and its dissolved gas at reservoir
condition to the volume at stock tank conditions is called the Oil Formation Volume
Factor Bo. The shape of the Bo vs. pressure curve is shown in Figure 4. It shows that
above the bubble point pressure the reduction in pressure from the initial pressure
causes the fluid to expand as a result of its compressibility. This relates to the chapter
on Phase Behaviour where for an oil the PV diagram shows a large decline in pressure
for a small increase in volume, being again an indication of the compressibility of the
liquid. Below the bubble point pressure this expansion due to compressibility of the
liquid is small compared to the ‘shrinkage’ of the oil as gas is released from solution.

The oil formation volume factor, is the volume in barrels (cubic metres) occupied
in the reservoir, at the prevailing pressure and temperature, by one stock tank
barrel (one stock tank cubic metre) of oil plus its dissolved gas.

Figure 4 Oil formation volume factor

©HERIOT-W ATT UNIVERSITY June 2014 v1


104

These black oil parameters, Bo and Rs are illustrated in Figure 5 a,b,&c from Craft and
Hawkins 1 reservoir engineering text, where they present the Rs and Bo curve for the
Big Sandy field in the USA. The visual concept of the changes during pressure and
temperature decrease is also presented.

©HERIOT-W ATT UNIVERSITY June 2014 v1


105

Figure 5 Gas to oil ratio and oil formation volume factor for Big Sandy Field reservoir oil 1.

The formation factor Bo may be multiplied by the volume of stock tank oil to find the
volume of reservoir required to produce that volume of stock tank oil. The shrinkage
factor can be multiplied by the volume of reservoir oil to find the stock tank volume.

It is important to note that the method of processing the fluids will have an effect on
the amount of gas released and therefore on both the values of the solution gas-oil ratio
and the formation volume factor. A reservoir fluid does not have single Bo or Rs
values. Bo & Rs are dependant on the surface processing conditions.

4 TOTAL FORMATION VOLUME FACTOR, BT

In reservoir engineering it is sometimes convenient to know the volume occupied in


the reservoir by one stock tank barrel of oil plus the free gas that was originally
dissolved in it. A factor is used called the total formation-volume factor Bt, or the
two-phase volume-factor and is defined as the volume in barrels that 1.0 STB and
its initial complement of dissolved gas occupies at reservoir temperature and
pressure, i.e. it includes the volume of the gas which has evolved from the liquid and
is represented by:

Bg (Rsb - Rs)

i.e. Bt = Bo + Bg (Rsb - Rs) (1)


Rsb = the solution gas to oil ratio at the bubble point

©HERIOT-W ATT UNIVERSITY June 2014 v1


106

Figure 8a Total formation volume factor or two phase volume factor (Mercury was commonly
used to pressurise the fluids in a PVT cell as it was inert and negligible compressibility)

Its application comes from the Material Balance equation (Chapter 15) where it is
sometimes used to express the volume of oil and associated gas as a function of
pressure. It is important to note that Bt does not have volume significance in reservoir
terms since the assumption in Bt is that the system remains constant. As mentioned
earlier if the pressure drops below the bubble point in the reservoir then the gas
coming out of solution moves away from its associated oil because of its favorable
relative permeability characteristics.

Figure 8b gives a comparison of the total formation-volume factor with the oil
formation-volume factor. Clearly above Pb the two values are identical since no free
gas is released. Below Pb the difference between the values represents the volume
occupied by free gas.

Figure 8b Total and oil formation volume factor

©HERIOT-W ATT UNIVERSITY June 2014 v1


107

The value of Bt can be estimated by combining estimates of Bo and calculation of Bg


and known solubility values for the pressures concerned.

5 BELOW THE BUBBLE POINT

Figure 9 depicts the behaviour below the bubble point when produced gas at the
surface comes from two sources: the solution gas associated with the oil entering the
wellbore plus free gas which has come out of solution in the reservoir and migrated to
the wellbore. The total producing gas to oil ratio, R, is made up of the two
components: solution gas Rs and the free gas which is the difference (R-Rs). The
diagram illustrates the volumes occupied by these two in the reservoir, the solution gas
being part of Bo and the free gas volume through Bg.

Figure 9 Production of reservoir hydrocarbons below bubble point

6 OIL COMPRESSIBILITY

The volume changes of oil above the bubble point are very significant in the context of
recovery of undersaturated oil. The oil formation volume factor variations above the
bubble point reflect these changes but they are more fundamentally embodied in the
coefficient of compressibility of the oil, or oil compressibility.

©HERIOT-W ATT UNIVERSITY June 2014 v1


108

The equation for oil compressibility is

in terms of formation volume factors this equation yields

Assuming that the compressibility does not change with pressure the above equation
can be integrated to yield ;

V2
co ( P2 - P1 ) = - ln
V1
where P1 & P2, and V1 & V2 represent the pressure and volume at conditions 1 & 2.

7 FLUID DENSITY

Liquids have a much greater density and viscosity than gases, and the density is
affected much less by changes in temperature and pressure. For petroleum engineers
it is important that they are able to estimate the density of a reservoir liquid at
reservoir conditions.

7.1 Specific Gravity of a Liquid

(4)

The specific gravity of a liquid is the ratio of its density to that of water both at the
same T & P. It is sometimes given as 60˚/60˚, i.e. both liquid and water are measured
at 60˚ and 1 atmos.

The petroleum industry uses another term called ˚API gravity where

©HERIOT-W ATT UNIVERSITY June 2014 v1


109

(5)

where o is specific gravity at 60˚/60˚.

There are several methods of estimating the density of a petroleum liquid at reservoir
conditions. The methods used depend on the availability and nature of the data. When
there is compositional information on the reservoir fluid then the density can be
determined using the ideal solution principle. When the information we have is that
of the produced oil and gas then empirical methods can be used to calculate the density
of the reservoir fluid.

7.2 Density based on Ideal Solution Principles


Mixtures of liquid hydrocarbons at atmospheric conditions behave as ideal solutions.
An ideal solution is a hypothetical liquid where no change in the character of the
liquids is caused by mixing and the properties of the mixture are strictly additive.

Petroleum liquid mixtures are such that ideal-solution principles can be applied for the
calculation of densities and this enables the volume of a mixture from the composition
and the density of the individual components. The principle is illustrated using the
following exercise. Data for the specific components are given in the tables at the end
of the chapter

Liquids at their bubble point or saturation pressure contain large quantities of


dissolved gas which at surface conditions are gases and therefore some consideration
for these must be given in the additive volume technique. This physical limitation does
not impair the mathematical use of a pseudo liquid density for methane and ethane
since it is only a step in its application to determine a reservoir condition density. This
is achieved by obtaining apparent liquid densities for these gases and determining a
pseudoliquid density for the mixture at standard conditions which can then be adjusted
to reservoir conditions.

Standing & Katz 8 carried out experiments on mixtures containing methane plus other
compounds and ethane plus other compounds and from this were able to determine a
pseudo-liquid (fictitious) density for methane and ethane

Correlations have been obtained by experiment giving apparent liquid densities of


methane and ethane versus the pseudoliquid density (Figure 10).

©HERIOT-W ATT UNIVERSITY June 2014 v1


110

Figure 10 Variation of apparent density of methane and ethane with density of the system 8.

To use the correlations a trial and error technique is required whereby the density of
the system is assumed and the apparent liquid densities can be determined. These
liquid densities are then used to compute the density of the mixture by additive
volumes and the value checked against the initial assumption. The procedure continues
until the two values are the same.

When non hydrocarbons are present, the procedure is to add the mole fractions of the
nitrogen to methane, the mole fraction of carbon dioxide to ethane and the mole
fraction of hydrogen sulphide to propane.

This trial and error method is very tedious so Standing and Katz devised a chart which
removes the trial and error required in the calculation. The densities have been
converted into the density of the heavier components, C3+, and the weight percent of
the two light components, methane and ethane in the C1+ and C2+ mixtures. Figure11.

©HERIOT-W ATT UNIVERSITY June 2014 v1


111

Figure 11 Pseudo-liquid density of systems containing methane and ethane 10.

We shall examine through examples various ways of calculating downhole reservoir


fluids densities dependant on the data available.
The three considered are:

1. The composition of the reservoir fluid is known.

2. The gas solubility, the gas composition and the surface oil gravity is known

1. The composition of the reservoir fluid is known.

©HERIOT-W ATT UNIVERSITY June 2014 v1


112

The pseudodensity is converted to reservoir conditions firstly by taking the effect of


pressure and secondly accounting for the effect of temperature. The variation of
density with respect to pressure and temperature has been investigated and it has been
demonstrated that thermal expansion is not affected by pressure. Standing & Katz took
National Petroleum Standards data and with supplementary data produced correction
factors for pressure and temperature to convert atmospheric density to reservoir
density.

The compressibility and thermal expansion effects have been expressed graphically in
Figures 12 and 13.

Figure 12 Density correction for compressibility of liquids 10.

©HERIOT-W ATT UNIVERSITY June 2014 v1


113

Figure 13 Density correction for thermal expansion of liquids 10.

Full compositional data may not always be available and the characterisation of the
produced fluids will vary from full compositional analysis to a description of the fluids
in terms of gas and oil gravity. The procedure just described is for the situation where
the composition of the reservoir fluid is known. The procedures which follow cover
the situation where a less comprehensive analysis is available. These methods make
use of empirical correlations.

©HERIOT-W ATT UNIVERSITY June 2014 v1


114

2. Reservoir Density when the Gas Solubility , the gas composition and the
surface oil gravity are known

By considering surface liquid as a single component and knowing the composition of


the collected gas the techniques previously discussed can be used to determine
reservoir liquid density.

8 FORMATION VOLUME FACTOR OF GAS CONDENSATE

The situation for a wet gas or gas condensate is different for a conventional oil when
one is considering the volume changes taking place upon release to surface conditions.
For a wet gas or condensate system, liquid at surface is gas in the formation. The
comparison therefore with respect to conditions in the reservoir to those at the surface
is distinctly different from an oil system, where an oil in the reservoir produces gas
and liquids at the surface. For a wet gas or condensate, a gas in the reservoir produces
gas and liquids at the surface.

The formation-volume factor therefore for a condensate, bgc is defined as the


volume of gas in the reservoir required to produce 1.0 stb of condensate at the
surface. The units are generally barrels of gas at res. Conditions per barrel of stock
tank oil. There are a number of methods of estimating Bgc.

9 VISCOSITY OF OIL

The viscosity of oil at reservoir temperature and pressure is less than the viscosity of
the dead oil because of the dissolved gases and the higher temperature. Correlations
are available which enable the dissolved gas and pressure effect on the dead oil
viscosity to be determined. Danesh 2 has given a good review of many of the empirical
approaches. The favoured correlations are those of Beggs and Robinson 11, Egbogah
and Ng 12 ,Vazquez and Beggs13 , and Labedi14 . Figure 14 gives plots, presented by
McCain 17 , of the correlation of dead oil viscosity from Egbogah and Ng 12 , and
Figure 15 shows the impact of dissolved gas from the Beggs and Robinson 11
correlation.

©HERIOT-W ATT UNIVERSITY June 2014 v1


115

Figure 14 Dead oil viscosities 17.

Figure 15 Viscosities of saturated black oils 11.

©HERIOT-W ATT UNIVERSITY June 2014 v1


116

Beggs and Robinson 11 examined 600 oil samples over a wide range of pressure and
temperature and came up with the following correlation.

od = 10A - 1 (6)

where, log A = 3.0324 - 0.0202oAPI -1.163 log T


od is the dead oil viscosity in cp and T is in oF.

Egbogah and Ng 12, had a different expression for A


log A = 1.8653 - 0.025086oAPI -0.56441 log T

Examination of these correlations has shown that they are not very reliable with
errors of the order of 25% (DeGetto15)

Beggs and Robinson 11 gave a correlation to give the impact of dissolved gas.

ob = CodB (7)

where C = 10.715 (Rs + 100)-0.515


and B = 5.44 (Rs + 150)-0.338
ob is the saturated oil viscosity

Vazquez and Beggs13 presented an equation to take into account pressure on viscosity
above the saturation pressure.

o = ob (P/Pb)D (8)

where D = 2.6 P1.187 e-11.513-8.98 x10-5P

This is presented in Figure 19 from McCain 17.

©HERIOT-W ATT UNIVERSITY June 2014 v1


117

Figure 16 Viscosities of undersaturated black oils 17.

Labedi (ref 14) also produced an empirical correlation to determine viscosity at


pressures above the bubble point

o= ob + (P/Pb-1)(10-2.488ob0.9036 Pb0.6151 /100.0197oAPI ) (9)

Danesh 2 in his text compared the various correlations from a published experimental
viscosity value in a well known PVT report.

©HERIOT-W ATT UNIVERSITY June 2014 v1


118

10 COMPARISON OF RESERVOIR FLUID MODELS

It is useful to summarise the differences between the Black Oil Model approach
compared to the Compositional Model in predicted fluid properties.

The suitability of the two approaches is largely related to the nature of the fluid. For
heavier oils where there are low GOR’s as compared to volatile oils with high GOR’s,
black oil models are likely to be suitable. For the more volatile systems where there
are more significant compositional variations in a reservoir as pressure is depleted,
compositional models are considered more capable of predicting fluid behaviour.

The computational requirements of compositional models used to be a restriction


when carrying out large reservoir simulation. The continued development of
computing and associated equations of state modelling reduces these former
restrictions.

Companies are now focusing their attention on being capable of modelling the total
process from fluid extraction from the reservoir, through well production and facility
treatment. At present separate modelling occurs, and many of these models are not
compatible. The black oil approach is certainly considered by many to be from a former
era.

The list below gives a summary comparison of the two approaches.

Black Oil Models


• 2 components - solution gas and stock tank oil
• Bo, Rs, etc.
• Empirical correlations
• After the event description of fluid properties

Compositional Models
• N components based on paraffin series
• Equation of state based calculations
• Feed forward calculation of fluid properties

The black oil approach which has been a major theme of this chapter uses the
characteristics of the produced fluids to determine the composition and properties of
the feed reservoir mixture, i.e. a back calculation.

©HERIOT-W ATT UNIVERSITY June 2014 v1


119

Table 1 Physical properties of the paraffin hydrocarbon and miscellaneous compounds.

©HERIOT-W ATT UNIVERSITY June 2014 v1


120

REFERENCES

1. Craft,BC & Hawkins, MF. Applied Reservoir Engineering” Prentice Hall, NY


2. Danesh, A, PVT and Phase Behaviour of Petroleum Reservoir Fluids.
Elsevier. pp 66-77
3. Standing MB “A pressure-Volume-Temperature Correlation for Mixtures of
Californian Oils and Gases”, Drill & Prod, Proc.275-287
4. Lasater, J.A. “ Bubble Point Correlation “, Trans AIME, 213,379-381
5. Vasquez,M and Beggs,HD “Correlations for Fluid Physical Property Prediction
“ JPT,968-970,
6. Glaso, O “Generalised Pressure Volume Temperature Correlations” JPT,785
795
7. Marhoun,MA, “PVT Correlations for Middle East Crude Oils” JPT, 650-665
8. Standing, M.B. and Katz,D.L. “ Density of Crude Oils Saturated with Natural
Gas” Trans AIME 146, 159
9. Kessler, MG and Lee,BI,: “Improved Prediction of Enthalpy of Fractions,” Hyd
Proc. 55,153-158.
10. Standing,M “Volumetric and Phase Behaviour of Oil Field Hydrocarbon
Systems” SPE Dallas
11. Beggs,HD. and Robinson,JR: Estimating the Viscosity of Crude Oil Systems”
JPT,27,1140-1141
12. Egboghah,EO and Ng,JT: ‘An improved Temperature Viscosity Correlations for
Crude Oil Systems”, J.Pet Sci and Eng.,5,197-200
13. Vasquez,M. and Beggs,HD :” Correlations for Fluid Physical Property
Predictions”. JPT,968
14. Labedi,R: “Use of Production Data to Estimate Volume Factor, Density and
Compressibility of Reservoir Fluids”, J.of Pet.Sci and Eng. 4.375-90
15. DeGhetto,G.,Paone,F. and Villa,M.: “Reliability Analysis of PVT Correlations
“,SPE 28904, Proc of Euro.Pet Conf. Lndn, 375-393
16. Danesh,A.,Krinis,D.,Henderson G.D., and Peden,J>M> “Visual Investigation of
Retrograde Phenomena and Gas Condensate Flow in Porous Media” 5th
European Symposium on Improved Oil Recovery ,Budapest
17. McCain,WD., “The Properties of Petroleum Fluids” Pennwell Books ,Tulsa, Ok
ISBN 0-87814-335-1
18. Macleod, DB., “On a Relation Between Surface Tension and Density,” Trans.,
Faraday Soc. 19,38-42.
19. Katz,DL.,”Handbook of Natural Gas Engineering”, McGraw Hill Book Co Inc.,
New Yk
20. Weinaug,KG and Katz,DL,: “Surface Tension of Methane-Propane Mixtures”.
I&EC,239-246
21. Firoozabadi,A , Katz,D.L., Soroosh,H.M and Sajjadian,V.A.: “Surface Tension

©HERIOT-W ATT UNIVERSITY June 2014 v1


121

of Reservoir Crude-Oil/Gas Systems Recognising the Asphalt in the Heavy


Fraction,” SPE Res Eng, 3,No 1, 265-272.

©HERIOT-W ATT UNIVERSITY June 2014 v1


122

TOPIC 7 FUNDAMENTAL PROPERTIES OF


RESERVOIR ROCKS
INTRODUCTION

1. CHARACTERISTICS OF RESERVOIR ROCKS

2. PHYSICAL CHARACTERISTICS OF
RESERVOIR ROCKS

3. POROSITY
3.1 Range of Values
3.2 Factors Which Affect Porosity
3.2.1 Packing and Size of Grains
3.2.2 Particle Size Distribution
3.2.3 Particle Shape
3.2.4 Cement Material

4. PERMEABILITY
4.1 Darcy's Law
4.2 Factors Affecting Permeability
4.3 Generalised Form of Darcy's Law
4.4 Dimensions of Permeability
4.5 Assumptions For Use of Darcy's Law
4.6 Applications of Darcy's Law
4.7 Field Units
4.8 Klinkenberg Effect
4.9 Reactive Fluids
4.10 Average Reservoir Permeability

5. STRESS EFFECTS ON CORE MEASUREMENTS

6. POROSITY - PERMEABILITY RELATIONSHIPS

7. SURFACE KINETICS
7.1 Capillary Pressure Theory
7.2 Fluid Distribution in Reservoir Rocks
7.3 Impact of Layered Reservoirs

©HERIOT-W ATT UNIVERSITY June 2014 v1


123

8. EFFECTIVE PERMEABILITY
8.1 Definition
8.2 Water Displacement of Oil
8.2.1 Water - Oil Relative Permeability
8.31 Gas Displacement of Oil and Gas – Oil
Relative Permeability

LEARNING OUTCOMES

Having worked through this chapter the Student will be able to:

• Define porosity and express it as an equation in terms of pore, bulk and grain
volume.
• Explain the difference between total and effective porosity.
• Define permeability and present an equation, Darcy’s Law, relating flow rate to
permeability in porous media.
• List the assumptions for the applicability of Darcy’s Law.
• Derive an equation based on Darcy’s Law relating flow of gas in a core plug and
the upstream and downstream pressures.
• Derive an equation relating flow rate to permeability for a radial incompressible
system.
• Comment on the difference between gas and liquid permeability (Klinkenberg
effect).
• Sketch a Figure relating liquid permeability to gas permeabilities plotted as a
function of reciprocal mean pressure.
• Briefly describe the impact of reservoir stresses on permeability and porosity
• Express the capillary pressure Pc as two equations, one in terms of interfacial
tension, contact angle and pore radius, and the other in terms of height and
density of fluids.
• Define the free water level.
• Draw the Pc or height vs. saturation capillary pressure curve and identify
significant features.
• Sketch and explain the impact of saturation, history, density difference and
interfacial tension in capillary pressure curves.
• Sketch the impact of capillary pressure effects on the saturation distribution of
stratified formations.
• Define effective and relative permeability and plot typical shapes.
• Define imbibition and drainage in the context of capillary pressure and relative

©HERIOT-W ATT UNIVERSITY June 2014 v1


124

permeability curves.
• Sketch the pore doublet model and use it to explain the retention of trapped oil in
large pores and briefly relate it to the principle behind some enhanced oil
recovery methods.
• Define mobility ratio.
• Sketch a typical gas-oil relative permeability curve.

©HERIOT-W ATT UNIVERSITY June 2014 v1


125

INTRODUCTION
The properties of reservoir rocks with respect to the fluids they contain and with
respect to the fluids which will be injected into them are important when
characterising a reservoir in terms of its reserves and the mobility of the fluids. This
next section gives a brief over view of these properties, and is followed by chapters on
their measurement and variation.

The reservoir engineer is concerned with the quantities of fluids contained within the
rocks, the transmissivity of fluids through the rock and other related properties.

1. CHARACTERISTICS OF RESERVOIR ROCKS

The specifications of a reservoir rock are such that there must be a large enough
capacity to store economically viable amounts of hydrocarbon and the hydrocarbon
must flow at economical rates when penetrated by a well. The factors which may
affect the capacity and the flow properties are the porosity, permeability, capillary
pressure, compressibility and fluid saturation. In the case of a reservoir rock, these are
not standard characteristics determined before formation of the rock, but are closely
linked to the geological processes that brought the sediments together and deposited
them in the sequences and under the chemical and physical changes inherent in the
system.

In order to contain enough oil or gas to make production economically viable, a


reservoir rock must exceed: a minimum porosity, a minimum thickness, a minimum
permeability, and a minimum area.

In order to extract the fluids the rock must be permeable which requires that there be
sufficiently large, interconnecting pores.

Although a permeable rock must also be porous, a porous rock is not necessarily
permeable. Certain volcanic rocks are porous but not permeable because the voids are
not interconnecting; shale may be quite porous but impermeable because the pores are
extremely small, thereby preventing free movement of the fluids contained within.

2. PHYSICAL CHARACTERISTICS OF RESERVOIR ROCKS

Considering a common reservoir rock, sandstone, the grains making up this rock are
all irregular in shape. The degree of irregularity, or lack of roundness reflects the
source sediments and the physical and chemical processes to which they were
subsequently exposed. Violent crushing or grinding action between rocks causes

©HERIOT-W ATT UNIVERSITY June 2014 v1


126

grains to be very irregular and sharp-edged. The tumbling action of grains along the
bottom of streams or seabeds smoothes sand grains. Wind-blown sand, as occurs in
moving dunes in deserts, results in sand grains that are even more rounded. Sand
grains that make up sandstone beds and fragments of carbonate materials that make up
limestone beds do not fit together congruently: the void space between the grains
forms the porosity.

The pore spaces (or interstices) in reservoir rock provide the container for the
accumulation of oil and gas and these give the rock its characteristic ability to absorb
and hold fluids. Most commercial reservoirs of oil and gas occur in sandstone,
limestone or dolomite rocks, however, some reservoirs occur in fractured shale and
even in basement rocks such as in Vietnam. Knowledge of the physical characteristics
of the pore space and of the rock itself (which controls the characteristics of the pore
space) is of vital importance in understanding the nature of a given reservoir.

For the reservoir engineer, porosity is one of the most important rock properties as a
measure of the space available for accumulation of hydrocarbon fluids.

3. POROSITY

The first step in forming a sandstone, for example, is to have a source of material
which is eroded and transported to low lying depressions and basins such as would be
found off the coasts of a landmass. The material would consist of a mixture of
minerals, but for a sandstone, the majority would be made of quartz in the form of
grains. When these were deposited, they would be surrounded by sea water or brine,
and as the sediment thickness increased, the weight or the pressure produced by the
overlying sediments would force the grains together. Where they contacted each other
large stresses would be produced and a phenomenon called pressure solution would
occur which dissolved the quartz at the points of contact between the grains until the
stresses reduced to a level which was sustainable by the grains. The dissolved material
would be free to precipitate in other regions of the sediment. In this way the initially
loose material would be solidified with discrete connections between the grains.
Initially, if subsea, the pore spaces would be filled with brine, and as the lithification
process occurred, some pore spaces would be isolated with the brine trapped inside. If
the vast majority were interconnected then the initial pore fluid would be free to be
swept through the rock by other fluids such as hydrocarbons. In this way the geometry
of the grains produces an assembly of solids with voids in between them. The grains
vary in diameter but may be from a few microns to several hundred microns. The
geometry of the pore spaces is such that they have narrow entrances (pore throats)
where the edges of the grains touch each other and larger internal dimensions (between

©HERIOT-W ATT UNIVERSITY June 2014 v1


127

the grains). The complicated nature of these interconnected pores is illustrated in


Figure 1 which is a metal cast of the pores in a sandstone rock.

Figure 1 Metallic Cast of Pore Spaces in a Consolidated Sand

One method of classifying reservoir rocks, therefore, is based on whether pore spaces (
in which the oil and gas is found) originated when the formation was laid down or
whether they were formed through subsequent earth stresses or ground water action.

The first type of porosity is termed original porosity and the latter, secondary or
induced porosity. This is illustrated in Figure 2.

Figure 2 Effective, isolated and total porosity

©HERIOT-W ATT UNIVERSITY June 2014 v1


128

Secondary porosity in limestone beds occurred as a result of fracturing, jointing,


dissolution, recrystallisation or a combination of these processes. Where water is
present in a carbonate formation, there is a continuous process of solution and
deposition or recrystallization. If solution is greater than deposition in any zone,
porosity will be developed between the crystal grains. An important type of porosity
of this kind is found in dolomite zones which occur in conjunction with large
limestone deposits. Dolomite may be deposited originally as a sedimentary rock, or it
may be formed by replacing the calcium carbonate in limestone rock with magnesium.

The impact of isolated pore space clearly cannot contribute to recoverable reserves of
fluid nor contribute to permeable pore space as illustrated in Figure 3.

Figure 3 Total, effective, isolated permeable and dead end pore space
Porosity is defined as the ratio of the void space or pore space (V p) in a rock to the bulk volume
(Vb) of that rock and it is normally expressed as a percentage of total rock volume. The
porosity is usually given the symbol  The matrix volume is the volume of the solid grains,
Vm.

Void volume
Porosity = x 100
Bulk volume

Bulk volume - Grain volume


Porosity = x 100
Bulk volume

pore volume
Porosity = ´ 100
void volume+grain volume (1)

©HERIOT-W ATT UNIVERSITY June 2014 v1


129

Figure 4 Representation of bulk, grain and pore volumes

These components are illustrated in Figure 4 for monosize spheres.

Total porosity is defined as the ratio of the volumes of all the pores to the bulk of a
material, regardless of whether or not all of the pores are interconnected. Effective
porosity is defined as the ratio of the interconnected pore volume of a material.

If the grains are represented by spheres stacked together as in Figure 4, then the pore
space can be seen between the solid grains.

Total Void Space


Total Porosity =
Vb

Interconnected Void Space


Effective porosity =
Vb

Induced or Secondary Porosity = porosity from fractures or vugs (large chambers


formed in certain carbonates and limestones caused by groundwater flow and
dissolution).

3.1 Range Of Values


The maximum porosity of porous media can be considered in relation to an assembly
of spheres arranged as a cubic packing of spheres. If the sides of a cube are assumed to
be formed by the lines drawn from the centre of each sphere to the adjacent spheres,
the cube in Figure 5 would be produced.

©HERIOT-W ATT UNIVERSITY June 2014 v1


130

Figure 5 Cube defined by the centres of each adjacent sphere

The length of each side would be 2x radius, giving the bulk volume as:

Vb = (2r)3 = 8r3

The grain volume would be the equivalent of the volume of one sphere

and the porosity (given the symbol ) would be

If the spheres fit in the cusps generated by the lower layer then a porosity of 26%
occurs. For a size distribution of spheres the ultimate minimum porosity would be
zero which would be the case if sufficient grains were available to completely fill the
pore spaces as shown in Figure 6 for part filling of the void.

©HERIOT-W ATT UNIVERSITY June 2014 v1


131

Figure 6 Minimum porosity when all pore spaces are filled

Several factors may combine to affect the porosity of a rock, but the main distinction
to be made is as follows based on the amount of connected pore volume, and whether
the pore space has been altered by dissolution or by fracturing after deposition and
lithification.

3.2 Factors Which Affect Porosity


The porosity (and permeability) of sandstone depend upon many factors, among which
are the packing, size and shape of the grains, variations in size of grains, arrangement
in which grains were laid down and compacted, and amount of clay and other
materials which cement the sand grains together.

3.2.1 Packing And Size Of Grains


The absolute sizes of the sand grains which make up a rock do not influence the
amount of porosity occurring in the rock. However variations in the range of sand
grains sizes do influence considerably the porosity.

3.2.2 Particle Size Distribution


If spheres of varying sizes are packed together, porosity may be any amount from 48
per cent to a very small amount approaching 0 per cent as shown in Figure 7.

3.2.3 Particle Shape


If the sand grains are elongated or flat and are packed with their flat surfaces together,
porosity and permeability may both be low we will discuss further in the context of
permeability.

©HERIOT-W ATT UNIVERSITY June 2014 v1


132

Figure 7 Reduction in porosity due to a range of particle sizes

3.2.4 Cement Material


Sandstones are compacted and usually cemented together with clays and minerals.
The porosity and permeability of a sandstone are both influenced to a marked degree
by the amount of cementing material present in the pore space and the way this
material occupies the pore space between the sand grains. The cementing material
may be uniformly located along the pore channels to reduce both porosity and
permeability or the cementing material may be located at the pore throats which
reduces the ability of fluid to enter the pore, but may not reduce the overall porosity of
the rock by a significant amount.

Limestone formations may have intergranular porosity. However, the pore openings
are more often inter-crystalline, that is spaces between microscopic crystals. They
also may take the form of pits or vugs caused by solution and weathering, or by
shrinkage of the matrix. These forms of porosity are called secondary porosity.
Another type of secondary porosity is that caused by fracturing and is very important
in that it permits many limestone rocks of otherwise low porosity to become excellent
reservoirs. Porosity may range from 50% to 1.5% and actual average values are listed
below:

Recent sands (loosely packed) 35 - 45%


Sandstones (more consolidated) 20 - 35%
Tight/well cemented sandstones 15 - 20%
Limestones (e.g. Middle East) 5 - 20%
Dolomites (e.g. Middle East) 10 - 30%
Chalk (e.g. North Sea) 5 - 40%

©HERIOT-W ATT UNIVERSITY June 2014 v1


133

A point that needs to be emphasised is that the concept of ‘porosity’ is complex and
therefore difficult to define and determine. It may refer to spaces between sand grains
or it may refer to limestone caves: it may even exclude a fraction of the free water
(water not bound chemically) present in the rock. Sometimes good estimates, (i.e.
relevant to reservoir development problems) may be obtained from laboratory studies,
or core samples, and sometimes such measurements are irrelevant.

In summary, the amount of porosity is principally determined by shape and


arrangement of sand grains and the amount of cementing material present, whereas
permeability depends largely on the size of the pore openings and the degree and type
of cementation between the sand grains. Although many formations show a
correlation between porosity and permeability, the factors influencing these
characteristics may differ widely in effect, producing rock having no correlation
between porosity and permeability.

4 PERMEABILITY

4.1 Darcy's Law


The permeability of a rock is the description of the ease with which fluid can pass
through the pore structure.

At one extreme, the permeability of many rocks is so low as to be considered zero


even though they may be porous. Such rocks may constitute the cap rock above a
porous and permeable reservoir and they include in their members clays, shales, chalk,
anhydrite and some highly cemented sandstones.

In petroleum engineering the unit of permeability is the Darcy, derived from the
empirical equation known as Darcy’s Law named after a French scientist who
investigated the flow of water through filter beds in 1856. His work provided the basis
of the study of fluid flow through porous rock.

(2)

where:

Q = flow rate in cm3/sec


A = cross sectional area of sample in cm2
∆P = pressure different across sample, atm

©HERIOT-W ATT UNIVERSITY June 2014 v1


134

 = viscosity in centipoise
L = length of sample in cm
k = permeability in Darcy

Darcy’s law of fluid flow states that rate of flow through a given rock varies directly
with the pressure applied, the area open to flow and varies inversely with the viscosity
of the fluid flowing and the length of the porous rock. In terms of equating the
parameters, the constant of proportionality in the equation is termed the permeability.
The unit of permeability is the Darcy which is defined as the permeability which will
permit a fluid of one centipoise viscosity (= viscosity of water at 68°F) to flow at a
linear velocity of one centimetre per second under a pressure gradient of one
atmosphere per centimetre. Permeability has the units Darcys. Figure 8 illustrates the
concept and the units of permeability

Figure 8 Concept of permeable rocks

Darcy’s Law experiment consisted of a sandpack through which water flowed at a


constant rate (Figure 9).

©HERIOT-W ATT UNIVERSITY June 2014 v1


135

Figure 9 Schematic of Darcy’s experiment

His results showed that the flowrate was directly proportional to the area open to flow,
the difference in pressure and inversely proportionate to the length of the sandpack,
i.e.

where Q is the flow rate, A is the area of the end of the core, h1 and h2 are the static
heads of water at the inlet and outlet of the core (the equivalent of the static pressure),
L is the length of the core. K is the constant of proportionality. It is constant for a
particular sand pack. When other workers replicated the experiment, the results were
different to those of Darcy. This was accounted for by inclusion of the viscosity of the
flowing fluid and the equation becomes:

kA(h1 - h2 )
Q=
mL

where the original terms have the same meaning and  is the viscosity of the fluid in
centipoise.

On a more theoretical basis, Poiseuille formulated the relationship between flow rate
and pressure drop for fluid flowing in a pipe. The form of the relationship is

©HERIOT-W ATT UNIVERSITY June 2014 v1


136

(3)

where Q is the flowrate, r is the radius of the tube,  is the viscosity of the fluid and L
is the length of the tube. In this case the dependence of the flowrate / pressure drop
relationship can be seen to be dependent on the radius of the tube. In a similar manner,
the radius of the pores in a rock dictate the nature of the relationship, specifically, the
radius of the pore throats is of most significance, since these are the smallest radii and
therefore affect the flowrate/ pressure drop relationship most.

The practical unit is the millidarcy (mD) which is 10-3 Darcy. Formation
permeabilities vary from a fraction to more than 10000 milli-Darcies. At the low end
of the range, clays and shales have permeabilities of 10-2 to 10-6 mD. These very low
permeabilities make them act as seals between more permeable layers.

4.2 Factors Affecting Permeability


Permeability along the flat surfaces will be higher, than the permeability in a direction
perpendicular to the flat surfaces of the grains. In a reservoir, the permeability
horizontally along the bed is usually higher than the permeability vertically across the
bed because the process of sedimentation causes the grains to be laid down with their
flattest sides in a horizontal position (minimising the area exposed to the prevailing
currents during deposition). Figure 10 illustrates the concept.

If sand grains of generally flat proportions are laid down with the flat sides non-
uniformly positioned and located in indiscriminate directions, both porosity and
permeability may be very high. To illustrate, if bricks are stacked properly, the space
between the bricks is very small; if the same bricks are simply dumped in a pile, the
space between the bricks might be quite large.

Figure 10 Directional Permeability

©HERIOT-W ATT UNIVERSITY June 2014 v1


137

The shape and size of sand grains are important features that determine the size of the
openings between the sand grains. If the grains are elongated, large and uniformly
arranged with the longest dimension horizontal, permeability to fluid flow through the
pore channels will be quite large horizontally and medium-to-large vertically. If the
grains are more uniformly rounded, permeability will be quite large in both directions
and more nearly the same. Permeability is found generally to be lower with smaller
grain size if other factors (such as surface tension effects) are not influential. This
occurs because the pore channels become smaller as the size of the grains is reduced,
and it is more difficult for fluid to flow through the smaller channels.

This directional perspective to any property is termed anisotropy. As shown above


permeability is a directional property and gives rise to different permeabilities
depending on the shape and depositional characteristics. Very dramatic anisotropy is
generated if a rock is fractured. These anisotropic perspectives are illustrated in Figure
11. Porosity is a non directional property and therefore is isoptropic.

Figure 11 Directional permeability.

4.3 Generalised Form Of Darcy’s Law


A three dimensional rock can be defined within the co-ordinate system illustrated in
Figure 12.

©HERIOT-W ATT UNIVERSITY June 2014 v1


138

Figure 12 Co-ordinate system for rock permeability

The x and y co-ordinates increase from zero to the left and out from the page; the z co-
ordinate increases downwards. The flow velocity in a particular direction can be
defined as the flowrate in that direction divided by the area open to flow. In any
direction, s, the flow velocity is termed Vs and is equated to the static pressure
gradient in that direction (i.e. the change in pressure, dP, over a small element of
length, ds in that particular direction) minus a contribution from the difference in head
(because of the difference in elevation) of the fluid across the section ds. Therefore,

(4)

and the change in elevation head is equal to the sine of the angle to the horizontal

The Darcy units are:

4.4 Dimensions Of Permeability

©HERIOT-W ATT UNIVERSITY June 2014 v1


139

From Darcy’s equation,

the dimensions of each term can be deduced in terms of length, L, mass, M and time, T

Therefore, the equation in terms of the dimensions (and keeping permeability as k) is

L kLT M ML
= ( 2 2 - 3 2)
T M LT LT

L K
=
T LT

K = L2
(5)
It can be seen that the dimensions reflect the nature of the constant of proportionality
and it should not be confused with, for example, the area open to flow, A, of the end of
a core or a sand pack. In terms of metric units, since 1 atm = 14.73 psi = 1.013 bar
and 1 cp = 10-3 Pas it follows that

1D = 9.87 x 10-13m2 ~ 1 x 10-12m2


1mD = 9.87 x 10-16m2 ~ 1 x 10-15m2
Other units of inches2 or cm2 could be used but they are all too large for porous media
and they would also require conversion to relate to permeabilities quoted in other
units. Darcys and milliDarcys are most commonly used.

4.5 Assumptions For Use Of Darcys Law


The simple Darcy Law, as used to determine permeability, only applies when the
following conditions exist:

(i) Steady state flow


(ii) Laminar flow

©HERIOT-W ATT UNIVERSITY June 2014 v1


140

(iii) One phase present at 100% pore space saturation


(iv) No reaction between fluid and rock
(v) Rock is homogenous

1. Steady state flow, i.e. no transient flow regimes. This becomes unrealistic in terms
of flow in a reservoir where the nature of the fluids and the dimensions of the reservoir
may produce transient flow conditions for months or even years. For laboratory based
tests, the cores are small enough that transient conditions usually last only a few
minutes.

2. Laminar flow, i.e. no turbulent flow. For most reservoir applications this is valid
however near to the well bore when velocities are high for example in gas production
turbulent flow occurs. Sometimes it is termed non- darcy flow. Figure 13

Figure 13 Effect of Turbulent Flow on Measured Permeability

3. Rock 100% saturated with one fluid, i.e. only one fluid flowing.

©HERIOT-W ATT UNIVERSITY June 2014 v1


141

In the laboratory this can be achieved by cleaning cores, however, there will be a
certain connate water saturation in the reservoir, and there may be gas, oil and mobile
water flowing through the same pore space. The concept of relative permeability can
be used to describe this more complex reservoir flow regime. Relative permeability is
discussed later.

4. Fluid does not react with the rock, i.e. it is inert and there is no change to the pore
structure through time.

There are cases when this may not happen, for example when a well is stimulated
during an hydraulic fracturing workover. The fluids used may react with the minerals
of the rock and reduce the permeability. In such cases, tests on the rock to determine
the compatibility of the treating fluids must be conducted before the workover.

5. Rock is homogeneous and isotropic, i.e. the pore structure and the material
properties should be the same in all directions and not vary. In reality, the layered
nature and large areal extent of a reservoir rock will produce variations in the vertical
and horizontal permeability.

4.6 Applications of Darcys Law


To examine the applicability of this simple relationship, approximations to the type of
flow encountered in a reservoir can be made: linear flow along a reservoir section and
radial flow into a wellbore. More complex geometries cannot be analysed using this
simple analytical equation and forms of approximating the geometry and flow are
required.

In the following expressions, the nomenclature is identical to that used above.

(i) Horizontal, linear, incompressible liquid system (Figure 14)

©HERIOT-W ATT UNIVERSITY June 2014 v1


142

Figure 14 Linear flow regime

From the basic Darcy equation

The flow rate and area open to flow are substituted for the flow velocity. The variables
are separated and integrated over the length (for the flow rate) and the pressures P1 to
P2 for the change in pressure. The pressure drop P2 minus P1 is negative and is
corrected by the negative sign on the left hand side of the equation.

(6)

The final form is as formulated by Darcy and the permeability will have the units of
Darcys if the other units are:

flow rate, Q - cm3s-1 pressure, P - atm


area open to flow, A - cm2 length, L - cm
viscosity,  - centipoise

(ii) Horizontal, linear, compressible ideal gas system

The flow regime is the same as for the linear liquid system and from the basic Darcy
equation:

©HERIOT-W ATT UNIVERSITY June 2014 v1


143

In this case, the laboratory measurement of the gas flow would usually be conducted
downstream from the core at almost atmospheric conditions (i.e. there would not be a
large pressure drop across the flow meter). It is assumed that the gas used is ideal,
however, there needs to be a correction to the volumetric flow rate measured to
account for the higher pressure in the core. Figure 15.

Figure 15 Configuration for gas permeability measurements.

The flowrate measured, Qb at ambient pressure, Pb is related to the flowrate, Q in the


core at the pressure in the core, P via the ideal gas law. If the assumption is made that
the temperature is constant, then

QP = Q b Pb

Q b Pb
Q =
P

and substituting into the equation, separating the variables and integrating produces

©HERIOT-W ATT UNIVERSITY June 2014 v1


144

(7)

2 mQb Pb L
k=
A( P12 - P2 2 ) (8)

Comparing the two expressions equations 6 and 7, it is seen that the gas flow rate is
proportional to the difference in the pressure squared, whereas the liquid flowrate is
proportional to the difference in the pressure. In well testing, the flow rates are
measured at the surface and for gas wells one of the diagnostic plots is the flowrate
versus difference in pressure squared plot. Neglecting the fact that the gas is real, it
gives an indication of the ability of the reservoir to produce gas.

kA( P12 - P22 ) kA( P1 - P2 )


Gas Q b = Liquid Q =
2 mLPb mL

The ideal gas permeability can be calculated from the liquid equation using mean
flowrate, Q measured at mean pressure.

(iii) Horizontal, radial, incompressible liquid system (Figure 16)

©HERIOT-W ATT UNIVERSITY June 2014 v1


145

Figure 16 Radial geometry with radial flow from the outer boundary to the wellbore

re is the outer boundary radius


rw is the inner boundary radius (well)
Pe is the pressure at the external boundary
Pw is the pressure at the inner boundary

Starting from the basic Darcy expression again,

Substituting for flow velocity,


Q
Vs = Vr =
A
In this case the direction of flow is in the opposite sense to the co-ordinate system,
therefore

ds = -dr

For radial geometry, the area, A, is now radius dependent therefore

A = 2rh

Substitution into the basic expression gives

(9)
separating the variables and integrating

©HERIOT-W ATT UNIVERSITY June 2014 v1


146

which gives the final form

(10)

(iv) Horizontal, radial, compressible real gas system

In this case the geometry is identical to that of the radial flow of incompressible fluid
with the modifications for the compressibility of a gas as per the linear gas flow
system.

If the assumption is made that the temperature is constant, then

QP = Q b Pb

Qb Pb
Q =
P

and substituting into the equation, 10

©HERIOT-W ATT UNIVERSITY June 2014 v1


147

separating the variables

and integrating produces

(11)

4.7 Field Units


Measurements made in the field are often quoted in field units and to ensure
compatibility with the Darcy equation, a conversion is required. The field units are
usually as follows:

Flow rate, Q - bbl/day or ft3/day


Permeability, k - Darcy
Thickness or height of reservoir, h - feet
Pressure, P - psia
Viscosity, m - centipoise
Radius, r - feet
Length, L - feet

In order to convert the Darcy equation for liquid flow,


KA( P1 - P2 )
Q =
mL
to oil field units, the following conversion factors are used:

©HERIOT-W ATT UNIVERSITY June 2014 v1


148

and these produce the following version of Darcy’s equation in field units:

bbl KA( P1 - P2 )
Q = 1.1271
day mL (12)

4.8 Klinkenberg Effect


Darcy’s Law would indicate that the permeability should be the same irrespective of
the fluid transmitted, since viscosity is included in the equation. Measurements made
on gas as against liquid for some conditions give higher permeabilities than the liquid.
This phenomenon is attributed to Klinkenberg, who attributed the behaviour to the
effect of the slippage of gas molecules along the solid grain surfaces. This occurs
when the diameter of the capillary opening (pore throat diameter) approaches the mean
free path of the gas (i.e. there is in effect only one gas molecule per capillary). Darcys
Law assumes laminar flow and viscous theory specifies zero velocity at the boundary
of the flow channel. This is not valid when the mean free path of the gas approaches
the diameter of the capillary and the result is that low pressure permeability
measurements are unrealistically high because there is insufficient gas molecules to
form a zero velocity boundary layer at the edges of the pores and to form a mass of
flowing gas within the pores. In this case, too many gas molecules flow through the
pores and the permeability appears to be higher than it actually is: the effect reported
by Klinkenberg. Since the mean free path is a function of the size of the molecule, the
permeability is a function of the type of gas used in the permeability measurement.
This gas permeability is corrected for the Klinkenberg effect by plotting the gas
permeability at each reciprocal mean pressure. This is illustrated for hydrogen,
nitrogen and carbon dioxide in Figure 17:

©HERIOT-W ATT UNIVERSITY June 2014 v1


149

Figure 17 Variation in gas permeability with reciprocal mean pressure.

Pm is the mean pressure of the gas (the mean of the upstream and downstream
pressures either end of the core p in Figure 15). In effect, if the gas pressure is raised
infinitely high, the gas will perform as an incompressible liquid would, therefore if
several measurements of permeability are made at different mean pressures, the
relationship between mean pressure and permeability can be extrapolated to the
equivalent pressure conditions of a liquid. In reality, extrapolation to infinity is
impossible, so the reciprocal mean pressure is used and the results are extrapolated to
zero reciprocal mean pressure (i.e. 1/infinitely high mean pressure). This point
corresponds to the liquid permeability. The different gasses have different slopes, but
they all extrapolate to the same equivalent liquid permeability.

The form of the equation developed by Klinkenberg is of the form


kG
kL =
b
l+
Pm
(13)
Where :

©HERIOT-W ATT UNIVERSITY June 2014 v1


150

kL = equivalent liquid permeability


kG = permeability to gas
Pm = mean flowing pressure
b = Klinkenberg constant for a particular gas and rock (slope of the gas permeability,
inverse mean pressure relationship).

The Klinkenberg effect is greatest for low permeability rocks and low mean pressures.

4.9 Reactive Fluids


Darcys Law assumes that the fluid does not react with the formation. Many formation
waters react with clays in the rock to produce a lower permeability to liquid than
would be obtained with gas. Therefore the permeability to water in the formation may
be much lower than would be determined to gas in the laboratory. Any water injected
into the formation may severely reduce the permeability due to clay swelling. The
change in permeability may be substantial, for example from several hundred
millidarcys to less than one millidarcy.

4.10 Average Reservoir Permeability


Permeability is not normally distributed but has an exponential distribution, therefore a
geometric mean is used to obtain an average reservoir permeability. Where definite
layers or units have been defined in the reservoir, their permeability can be dealt with
in a similar manner to resistors in series or parallel. We will deal with horizontal beds
in parallel.
The horizontal system reflects the sedimentary nature of the rock: the rock material
may have been segregated as it was deposited giving different sizes, shapes etc. to
different layers in the formation. The reason for determining an average permeability
is in rationalising the permeability measured on small samples in the laboratory with
the measurements made for example by well test analysis. Well test analysis cannot
test small sections of the reservoir (which would be uneconomical) on the same scale
as the laboratory tests. The results are therefore representative of flow through several
layers rather than only one.

©HERIOT-W ATT UNIVERSITY June 2014 v1


151

Linear Flow
Consider the simple linear beds in parallel (Figure 18.)

Figure 18 Linear Flow in Parallel

The average permeability can be developed using the Darcy flow equation:

QT = Q1 + Q2 + Q3 (14)

k1 A1 ( P1 - P2 ) k2 A2 ( P1 - P2 )
Q1 = ...etc.
mL Q 2 = mL (15)

QT = k å Ai ( P1 - P2 ) / mL (16)

k1 A1 ( P1 - P2 ) k2 A2 ( P1 - P2 ) k3 A3 ( P1 - P2 )
= + +
mL mL mL (17)

k=
åk A i i

åA i
(18)

If all the beds have the same width the A  h so k is the arithmetic average:

©HERIOT-W ATT UNIVERSITY June 2014 v1


152

kA =
åk h i i

åh i
(19)

This equation is commonly used to determine the average permeability of a reservoir


from core analysis.

Radial Circular Flow


This is the case of several superimposed layers flowing simultaneously in the well.
Each layer supplies a rate of Qi. The total rate of flow is QT =  Qi. In Figure 19:

Figure 19 Radial Flow in Parallel

(20)

(21)

k=
åh k i i

hT (22)

©HERIOT-W ATT UNIVERSITY June 2014 v1


153

This value can be compared with that obtained through well flow tests or pressure
build-up tests.

5 MEASUREMENT OF POROSITY/PERMEABILITY AND THE


STRESS EFFECTS ON CORE MEASUREMENTS

The measurement of porosity and permeability is conducted usually on cylindrical


cores. These are placed in a core holder which seals the core in place, and in which the
porosity and permeability can be measured. These coreholders can be operated at
ambient (laboratory conditions) to give estimates of the porosity and permeability at
surface conditions, (figure 19a). However, in reservoir engineering the impact of
reservoir stresses on reservoir flow and capacity parameters has been considered for a
number of years but, increasingly, the interest in stress related measurement has
grown. The effect of removing a core from the formation is to remove all the
confining forces on the sample, allowing the rock matrix to expand in all directions,
partially changing the shapes of the fluid-flow paths inside the core. As the rock
matrix is subjected to a stress, it will deform and alter the pore space volume as the
rock is compressed. For simplicity, the overburden will be considered to produce
hydrostatic stress (called the compacting stress) on the reservoir, i.e. a grain-to-grain
stress in the rock. Within the pores, fluid pressure acts on the surface of the grains and
reduces the grain-to-grain (or compacting) stress. Therefore in a real reservoir there is
a balance between the effect of the overburden stress and the pore pressure. This can
be described by the relationship

Pcompacting = Poverburden - Ppore pressure

where Pcompacting is the grain-to-grain stress, Poverburden is the stress produced by the
weight of the overburden at a particular depth and Ppore pressure is the pressure of the
fluids in the pores. The expression shows the balance between the overburden and the
pore pressure in compacting the rock matrix: if the pore pressure declines, the
compacting stress increases and the pore volume declines. This assumes that the
overburden remains constant which is logical over the time period of a producing
reservoir. The balance can be represented in the figure 19b

©HERIOT-W ATT UNIVERSITY June 2014 v1


154

Figure 19a Core holder to measure porosity, permeability (as is the case in the figure) or pore
volume reduction (by squeeze out of pore fluids)

Figure 19b The balance between overburden & rock stress and fluid pressure

Po = Pf + Pc Po = overburden pressure
Pf = fluid pressure

©HERIOT-W ATT UNIVERSITY June 2014 v1


155

Pc = compacting stress

The effect of the change in the balance between the overburden stress and the pore
pressure is to change the compacting stress. If there is an increase in pore pressure,
then the pore volume will increase, however, this is rare and in the main, pore pressure
declines during production and the pore spaces compact under the increasing
compacting stress. Two issues are significant: the initial porosity in the reservoir (i.e.
to correctly define the volume of oil in place) and the reduction in that porosity (or
pore volume) as the pressure declines (for material balance and simulation studies).
Figure 19c shows the relationship between porosity and depth (or stress). As the depth
(and stress) increases, the porosity declines. Care needs to be taken when assessing
porosity values: were they measured under overburden or at ambient conditions? The
shale sample shows a large change in porosity as the plate-like clay minerals are
compacted and fit together in a more congruent manner.

Figure 19c Alteration in porosity with depth of burial (or stress)

The rate of change of pore volume with pressure change can be represented by an
isothermal compressibility (assuming temperature is constant):

1 dv
Cf = -
v dP

where
Cf is the isothermal compressibility, v is the volume, dv is the change in volume and
dP is the change in pressure (the negative sign accounts for the co-ordinate system: as
the pressure increases, the volume decreases).

©HERIOT-W ATT UNIVERSITY June 2014 v1


156

As the effect of a stress on the rock matrix affects the pore volume, it also affects the
pore throat radii and the permeability of the rock. In general, an increase in stress
reduces the pore throat radii and the permeability declines. For most rocks subjected to
an hydrostatic stress, this is the case as the stress is equal in all directions. Figure 19d
shows typical permeability declines for increase in stress for sandstone.

Figure 19d The reduction in permeability for a range of sandstone samples (the porosity is in
the range 15% to 22%)

Unconsolidated material has larger absolute changes in permeability as the total strain
is greater.

©HERIOT-W ATT UNIVERSITY June 2014 v1


157

In terms of the measurement of pore compressibility it is usually conducted in a


coreholder which applies an equal compacting pressure around the core. An inner liner
ensures the power fluid (usually hydraulic oil) does not contaminate the pores of the
sample. The pore pressure is usually kept at ambient, i.e. the compacting pressure
mimics the net effect of the overburden and the pore pressure in the reservoir. This
makes the test simpler, however, there may be conditions where the compressibility of
the grains themselves plays a significant role in the system and the test may require to
be conducted at true overburden and pore pressure conditions. For the test at ambient
pore pressure conditions, an outlet is connected to the core holder and this is lead to a
pipette or a balance to measure the amount of pore fluid expelled. The pressure of the
hydraulic oil is increased in stages and for each stage the amount of fluid expelled is
measured after the rock has come to equilibrium. The data can then be analysed to
indicate the change in porosity or pore compressibility. Figure 19e illustrates the
concept.

Figure 19e Measurement of the reduction in pore volume as the external stress (or compacting
pressure) is increased

The results show the change in pore volume relative to the original pore volume, for a
given change in the compacting pressure (this assumes that changes in the compacting
pressure have the same effects as changes in the pore pressure) which can be
substituted in to the isothermal compressibility as

1 dv p
Cp = -
v p dPc
where:
Cp = pore volume compressibility
vp = initial pore volume
dvp = change in pore volume (amount of fluid expelled)

©HERIOT-W ATT UNIVERSITY June 2014 v1


158

dPc = change in compacting pressure

Typical values of pore compressibility are in the range 3 x10-6 psi-1 to 10 x10-6 psi-1,
however, soft sediments can have compressibilities in the range 10 x10-6 psi-1 to
20x10-6 psi-1 or 30 *10-6 psi-1.

6. POROSITY-PERMEABILITY RELATIONSHIPS

Whereas for porosity there are a number of downhole indirect measurement methods,
the same is not the case for permeability. The downhole determination of permeability
is more illusive. Down hole permeability is mainly obtained by flow and pressure
determination and requires other characteristics for example the flowing interval.
There has been a continued interest in porosity-permeability correlations, on the basis
if one has a good correlation of laboratory measured porosity and permeability then
down hole measurements of porosity could unlock permeability values for those
formations where recovered core has not been practical. Although porosity is an
absolute property and dimensionless, permeability is not and is an expression of flow
which is influenced by a range of properties of the porous media, including the shape
and dimensions of the grains and the porosity. Since porosity is an important
parameter in permeability it is not surprising for those rocks which have similar
particle characteristics that a relationship exists between porosity and permeability.
Figure 18 below gives examples of permeability correlations for different rock types.

©HERIOT-W ATT UNIVERSITY June 2014 v1


159

Figure 20 Permeability and Porosity Trends for Various Rock Types


(Core Laboratories Inc)

If core for a particular section cannot be recovered, or for example is formed as a pile
of sand on the rig floor, then correlations like these in Figure 18 are used. Porosity
measurements obtained indirectly from wireline methods can be used to obtain the
laboratory porosity vs down hole porosity cross plot. Using this laboration porosity
value the associated permeability value can be determined from an appropriate
correlation as in Figure 20.

7. SURFACE KINETICS
The simultaneous existence of two or more phases in a porous medium needs terms
such as the capillary pressure, relative permeability and wettability to be defined. With
one fluid only one set of forces needs to be considered: the attraction between the fluid
and the rock. When more than one fluid is present there are three sets of active forces
affecting capillary pressure and wettability.

©HERIOT-W ATT UNIVERSITY June 2014 v1


160

Surface free energy exists on all surfaces between states of matter and between
immiscible liquids. This energy is the result of electrical forces. These forces cause
molecular attraction between molecules of the same substance (cohesion) and between
molecules of unlike substances (adhesion).

Surface tension (or interfacial tension) results from molecular forces that cause the
surface of a liquid to assume the smallest possible size and to act like a membrane
under tension.

7.1 Capillary pressure theory


The rise or depression of fluids in fine bore tubes is a result of the surface tension and
wetting preference and is called capillarity. Capillary pressure exists whenever two
immiscible phases are present, for example, in a fine bore tube and is defined as the
pressure drop across the curved liquid interface. The equilibrium in force between the
molecules of a single phase is disrupted at an interface between two dissimilar fluids.
The difference in masses and the difference in the distances between the molecules of
the different phases produces an initially unbalanced force across the interface. Figure
21 shows the interface between oil and water molecules.

Figure 21 Representation of an oil water boundary

Interfacial tension deforms the outer surface of immiscible liquids to produce droplets.
If the two liquids are present on a surface, the interfacial tension deforms the liquids to
produce a characteristic contact angle as shown in Figure 22.

A wetting phase is one which spreads over the solid surface and preferentially wets the
solid. The contact angle approaches zero (and will always be less than 90˚).

©HERIOT-W ATT UNIVERSITY June 2014 v1


161

A non-wetting phase has little or no affinity for a solid and the contact angle will be
greater than 90˚

Figure 22 Interfacial tension between oil, water and a solid

The contact angle describes the nature of the interaction of the fluids on the surface:
for the oil-water system shown above, an angle less than 90˚ indicates that the surface
is water wet. If the angle were greater than 90˚ then the surface would be oil wet.

The composition of the surface also affects the interfacial tension. Figure 23 shows the
effect of octane and napthenic acid on a water droplet on silica and calcite surfaces.
The water is not affected by the change in surface in the water/octane system,
however, the napthenic acid causes the water to wet the silica surface, but to be non-
wetting on the calcite surface.

©HERIOT-W ATT UNIVERSITY June 2014 v1


162

Figure 23 The effect of a change in the surface on wetting properties

The Adhesion tension, At is defined as the difference between the solid water and
solid oil interfacial tension. This is equal to the interfacial tension between the water
and oil multiplied by the cosine of the contact angle,

At = sw - so = wo Cos wo


(23)
If a container of oil and water is considered as in Figure 24, the denser water lies
below the oil.

Figure 24 Capillary rise in an oil/water system

If a glass capillary tube of radius, r is inserted such that it pierces the interface between
the oil and water, the geometry of the tube and the imbalance in forces produced
between the glass, oil and water cause the interface to be pulled upwards into the tube.
If non wetting fluids were used, the interface in the tube may be pushed downwards.
Under equilibrium conditions, i.e. after the tube has pierced the original interface, the
adhesion tension around the periphery (2r) of the tube can be summed to give the
total force upwards. Since the interface is static, this force must be balanced by the
forces in the column of water drawn up the tube and the equivalent column of oil
outside the tube, i.e. at point C, the force (or pressure) must be the same in the tube as
outside, therefore the excess force produced by the column of water is balanced by the
adhesion tension.

©HERIOT-W ATT UNIVERSITY June 2014 v1


163

net force upwards = 2r woCos    

net force downwards = (wgh - ogh)r2 = gh(w - o)r2 (25)

the interface is at equilibrium, therefore

2r woCos=gh(w - o)r2 (26)

The capillary pressure is the difference in pressure across an interface, therefore in


terms of pressure (the pressure Pc, force acting on area r2)

(27)

It can be seen from the equations, capillary pressure can be defined both in terms of
curvature and in terms of interfacial tension, as expressed by the hydrostatic head.

(28)

where
Pc = capillary pressure
 = surface tension
 = contact angle
rc = radius of the tube
h = height of interface
w = the density of water
o = the density of oil.

©HERIOT-W ATT UNIVERSITY June 2014 v1


164

For a distribution of capillaries, therefore, the capillary pressure will give rise to a
distribution of ingress of wetting fluid into the capillaries. The relative position of the
capillary rise is given with respect to the free water level, FWL, i.e. the point of zero
capillary pressure. Figure 25 illustrates the effect of three different capillary radii on
the rise of water. Figure 26 shows the behaviour for a full assembly of capillaries and
alongside the associated capillary pressure curve. In this Figure it is important to note
five aspects.

• The free water level-the position of zero capillary pressure


• The oil -water contact
• The 100% water saturation at a distance above the free-water level due to the
capillary action of the largest tube.
• The irreducible level representing the limit if mobile water saturation
• The different radii segregate the capillary pressure and therefore the height to
which the water is drawn into the oil zone.

The zone of varying water saturation with height above the 100% free water oil
contact is called the transition zone.

The formation containing irreducible water will produce only hydrocarbons whereas
the transition zone of varying water saturation will produce water and hydrocarbons.

The shape of the capillary pressure curves in the transition zone will depend on the
nature of the rock.

©HERIOT-W ATT UNIVERSITY June 2014 v1


165

Figure 25 Capillary Rise in Distribution of Capillaries

Figure 26 Capillary Pressure Curve

It must be remembered that although concepts of capillary pressure were formulated in


terms of fine bore tubes, application in practice deals with a complex network of
interconnected pores in a matrix carrying surface chemical properties as illustrated in
Figure 1 of the pore cast of the pore space.

The height at which a wetting liquid will stand above a free level is directly
proportional to capillary pressure which is related to the size and size distribution of
the pores. It is also proportional to interfacial tension and the cosine of the contact
angle and inversely proportional to the tube radius and difference in fluid density. The
smaller the pores ie. the lower the permeability, the higher the capillary pressure.

7.2 Fluid distribution in reservoir rocks


Water is retained by capillary forces as hydrocarbons accumulate in productive
reservoirs. The water is referred to as connate or interstitial water and in water wet
rocks it coats the rock surfaces and occupies the smallest pores, whereas hydrocarbons
occupy the centre of the larger pores. The magnitude of the water saturation retained is
proportional to the capillary pressure which is controlled by the rock fluid system.

©HERIOT-W ATT UNIVERSITY June 2014 v1


166

Water wet, coarse grained sand and oolitic and vuggy carbonates with large pores have
low capillary pressure and low interstitial water contents. Silty, fine grained sands
have high capillary pressures and high water contents.

Reservoir saturation reduces with increased height above the hydrocarbon-water


contact. At the base of the reservoir there will usually be a zone of 100% water
saturated rock. The upper limit of this is referred to as the water table or water oil
contact (WOC). However, there is a non identifiable level, the free water level
representing the position of zero capillary pressure.

Figure 27shows the capillary pressure curve for a reservoir where the water saturation
reduces above the aquifer. The 100% water saturation continues some distance above
the free water level corresponding to the largest pores of the rock, hD. Above this level
both the oil and water are present and the reservoir water saturation decreases with
increased height above the hydrocarbon water contact, since the larger pores can no
longer support the water by capillary action and the water saturation falls. Between the
100% WOC and the irreducible saturation level is termed the transition zone.

©HERIOT-W ATT UNIVERSITY June 2014 v1


167

Figure 27 Capillary Pressure Curve for Porous media

Consider the capillary pressure curves for the two rocks in Figure 28. The first sample
(case 1) has a small range of connecting pore sizes. The second sample (case 2) has a
much larger range of connecting pore sizes, although the largest pores are of similar
size in both cases. Also, in case 2, the irreducible water saturation is reached at low
capillary pressure, but with the graded system, a much larger capillary pressure is
needed.

©HERIOT-W ATT UNIVERSITY June 2014 v1


168

Figure 28 Capillary Pressure Curves for Different Rocks

In addition to water transition zones, there can also be an oil/gas transition zone, but
this is usually less well defined.

Rock wettability influences the capillary pressure and hence the retentive properties of
the formation. Oil wet rocks have a reduced or negligible transition zone, and may
contain lower irreducible saturations. Low fluid interfacial tension reduces the
transition zone, while high interfacial tension extends it. Figure 29 illustrates this
effect.

Figure 29 Interfacial Tension Effect

©HERIOT-W ATT UNIVERSITY June 2014 v1


169

Saturation history influences the capillary pressure water saturation relationship and
therefore the size of the transition zone. Drainage saturation results from the drainage
of the wetting phase (water) from the rock as the hydrocarbons accumulate. It
represents the saturation distribution which exists before fluid production. The level
of saturation is dictated by the capillary pressure associated with the narrow pore and
is able to maintain water saturation in the large pore below. Imbibition saturation
results from the increase in the wetting phase (water) and the expulsion of the
hydrocarbons. In this case the saturation is determined by the large pore reducing the
capillary pressure effect and preventing water entering the larger pore. This is the
situation which occurs both when natural water drive imbibes into the formation
raising the water table level and in water injection processes. Clearly the two
saturation histories generate different saturation height profiles. Figure 30 shows the
drainage and imbibition effects on capillary rise.

Figure 30 Saturation History Effect

A large density difference between water and hydrocarbons (water-gas) suppresses the
transition zone. Conversely, a small density difference (water-heavy oil) increases the
transition zone. Figure 31 shows the differences in density for water/heavy oil and
water/gas on capillary rise. Transition zones between oil and gas are not significant
because of the large density difference between oil and gas.

©HERIOT-W ATT UNIVERSITY June 2014 v1


170

Figure 31 Fluid Density Effect

7.3. Impact of Layered Reservoirs


A characteristic of reservoirs is the various rock types making up the reservoir section.
Each rock type has its own capillary pressure characteristics. Wells penetrating such
formations will show a water saturation distribution reflecting the specific capillary
effects of each formation type. In some cases a 100% water saturation will be above a
lower water saturation associated with a lower elevation material with a higher
permeability, Figure 32.
For example well A would only indicate 100% water. Well B would penetrate the
transition zone of the top layer then a region of 100% water saturation. The saturation
profiles for well B and C are illustrated in Figure 30. The transition zone of the next
layer 2, followed by an interface of 100% saturation associated with layers 2, 3 and 4
then into 100% for the next two layers. Well D penetrates through the top and next
layer at the irreducible saturation level, into the transition zone for layer three, then
into irreducible saturation for the 4th layer.

©HERIOT-W ATT UNIVERSITY June 2014 v1


171

Figure 32 Capillary Effects in Stratified Formations

8 EFFECTIVE PERMEABILITY

8.1 Definition
The idea of relative permeability provides an extension to Darcy’s Law to the presence
and flow of more than a single fluid within the pore space. When two or more
immiscible fluids are present in the pore space their flows interfere. Specific or
absolute permeability refers to permeability when one fluid is present at 100%
saturation. Effective permeability reflects the ability of a porous medium to permit the
passage of a fluid under a potential gradient when two or three fluids are present in the
pore space. The effective permeability for each fluid is less than the absolute
permeability. For a given rock the effective permeability is the conductivity of each
phase at a specific saturation. As well as the individual effective permeabilities being
less than the specific permeability, their sum is also lower.

If measurements are made on two cores having different absolute permeabilities k1 and
k2, there is no direct way of comparing the effective permeability kw and ko curves
since for the two cores they start at different points k1 and k2. This difficulty is
resolved by plotting the relative permeability krw and kro where

©HERIOT-W ATT UNIVERSITY June 2014 v1


172

Relative Permeability =

Relative permeability is dimensionless and is reported as a fraction or percentage. On


relative permeability plots the curves start from unity in each case, so direct
comparisons can be made.

A typical set of effective permeability curves for an oil water system is shown in
Figure 33 and for a gas oil system in Figure 34.

Figure 33 Relative permeability curves for water-oil system

©HERIOT-W ATT UNIVERSITY June 2014 v1


173

Figure 34 Relative permeability curves for gas-oil system

The following points are to be noted:

The introduction of a second phase decreases the relative permeability of the first
phase: for example, kor drops as Sw increases from zero. Secondly, at the point where
the relative permeability of a phase becomes zero there is still a considerable
saturation of the phase remaining in the rock. The value of So at kro = 0 is called the
residual oil saturation and the value of Sw at krw = 0 is called the irreducible water
saturation.

The shapes of the relative permeability curves are also characteristic of the wetting
qualities of the two fluids (Figure35). When a water and oil are considered together,
water is almost always the wetting phase. This means that the water, or wetting phase,
would occupy the smallest pores while the non-wetting phase, or oil phase, would
occupy the largest pores. This causes the shape of the relative permeability curves for
the wetting and non-wetting phase to be different.

©HERIOT-W ATT UNIVERSITY June 2014 v1


174

Figure 35 Oil and Water Relative Permeability Curves for Water-Wet and Oil-Wet Systems
(Core Laboratories Inc)

This is illustrated by looking at the relative permeability to one phase at the irreducible
saturation of the other phase. The relative permeability to water at an irreducible oil
saturation of 10% (90% water) is about 0.6, Figure 33, whereas the relative
permeability to the non-wetting phase, oil, at the irreducible water saturation of 0.3
approaches 1.0. In this case it is 0.95. One practical effect of this observation is that it
is normally assumed that the effective permeability of the non-wetting phase in the
presence of an irreducible saturation of the wetting phase is equal to the absolute
permeability. Consequently, oil flowing in the presence of connate water or an
irreducible water saturation is assumed to have a permeability equal to the absolute
permeability. Similarly, gas flowing in a reservoir in the presence of irreducible water
saturation is assumed to have a permeability equal to the absolute permeability.

Relative permeability characteristics are important in the displacement of


hydrocarbons by water, and in the displacement of oil and water by gas. Such
displacements occur during primary and secondary recovery operations, as well as
during coring and core recovery.

©HERIOT-W ATT UNIVERSITY June 2014 v1


175

Relative permeability data when presented in graphical form are often referred to as
drainage or imbibition curves. (Figure 35)

Imbibition relative permeability is displacement where the wetting phase saturation is


increasing. For example, in a water flood of a water wet rock, or coring with a water
base mud.

Drainage relative permeability is where the non-wetting phase saturation is increasing.


For example, gas expulsion of oil during primary depletion or gas expansion of fluids
during core recovery, and the condition existing in the transition zone at discovery.

Water displacement of oil differs from gas displacement of oil since water normally
wets the rock and gas does not. The wetting difference results in different relative
permeability curves for the two displacements.

8.2 Water displacement of oil


Prior to water displacement from an oil productive sand interstitial water exists as a
thin film around each sand grain with oil filling the remaining pore space. The
presence of water as previously stated has little effect on the flow of oil, and oil
relative permeability approaches 100%. Water relative permeability is zero.

Water invasion results in water flow through both large and small pores as the water
saturation increases. Imbibition relative permeability characteristics influence the
displacement. Oil saturation decreases with a corresponding decrease in oil relative
permeability. Water relative permeability increases as water saturation increases.

Oil remaining after flood-out exists as trapped globules and is referred to as residual
oil. This residual oil is immobile and the relative permeability to oil is zero. Relative
permeability to water reaches a maximum value, but is less than the specific
permeability because the residual oil is in the centre of the pores and impedes water
flow.

8.2.1 Water-oil relative permeability


Accumulation of hydrocarbons is represented by drainage relative permeability curves
as the water saturation decreases from 100% to irreducible. Water relative
permeability reduces likewise from 100% to zero while oil relative permeability
increases.

©HERIOT-W ATT UNIVERSITY June 2014 v1


176

Subsequent introduction of water during coring or water flooding results in a different


set of relative permeability curves - these are the imbibition curves. The water curve is
essentially the same in strongly water wet rock for both drainage and imbibition. The
oil phase relative permeability is less during imbibition than during drainage.

The oil remaining immobile after a waterflood is influenced significantly by the


capillary pressure and interfacial tension effects of the system. It is of note that a high
residual oil saturation is a result of the oil ganglia being retained in the large pores as a
result of capillary forces. Figure 36 illustrates the pore doublet model illustrating how
oil can be trapped in a large pore. The forces to displace this droplet have to overcome
capillary forces and are too great to use pressure through pumping. The force required
can be reduced by reducing the interfacial tension which is the basis for many
enhanced oil recovery methods; for example, surfactant and miscible flooding.

©HERIOT-W ATT UNIVERSITY June 2014 v1


177

Figure 36 Pore Doublet Model

An important perspective in a displacement process is the concept of mobility ratio.


This relates the mobility of the displacing fluid relative to that of the displaced fluid.
It is therefore a ratio of Darcy’s Law for each respective fluid at the residual saturation
of the other fluid. In the context of water displacing oil.

(20)

where k'rw is the relative permeability at residual oil saturation


k'ro is the relative permeability at the irreducible water saturation.

These relative permeabilities are sometimes referred to as end point relative


permeabilities. When M is less than 1 this gives a stable displacement whereas when
M is greater then 1 unstable displacement occurs.
This topic is covered extensively in the chapter on immiscible displacement

8.3 Gas displacement of oil and gas-oil relative permeability


Gas is a non-wetting phase and it initially follows the path of least resistance through
the largest pores. Gas permeability is zero until a ‘critical’ or ‘equilibrium’ saturation
is reached (Figure 34).

Gas saturation less than the critical value is not mobile but it impedes the flow of oil
and reduces oil relative permeability. Successively smaller pore channels are invaded
by gas and joined to form other continuous channels. The preference of gas for larger
pores causes a more rapid decrease of oil relative permeability than when water
displaces oil from a water wet system. Figure 37 shows the alteration of relative
permeability as gas comes out of solution and flows at increasing saturation through
the oil reservoir. These gas/oil relative permeability curves are very significant in
relation to the drive mechanism of solution gas drive.

©HERIOT-W ATT UNIVERSITY June 2014 v1


178

Figure 37 Gas Oil Relative Permeabilities ( Core Lab)

©HERIOT-W ATT UNIVERSITY June 2014 v1

You might also like