Download as pdf or txt
Download as pdf or txt
You are on page 1of 59

Section 3 – Well Performance

CONTENTS
Page

3 WELL PERFORMANCE 1
3.1 SUMMARY 1
3.2 INTRODUCTION 1
3.3 DEFINITION OF WELL PERFORMANCE 1
3.3.1 Inflow Performance Relationship 3
3.3.2 Vertical Lift Performance 3
3.3.3 Overall Pressure Drops 4
3.3.4 Well Design For Life Of The Well 4
3.4 PRESSURE/VOLUME/TEMPERATURE 5
3.4.1 Direct Use Of Laboratory Data 8
3.4.2 Equation Of State 8
3.4.3 Empirical Correlations - Untuned 10
3.5 INFLOW PERFORMANCE 11
3.5.1 Radial Inflow Equation And Skin 12
3.5.2 Vogel Inflow Performance 13
3.5.3 Fetkovich 13
3.5.4 Jones 14
3.5.5 Hydraulically Fractured Wells 14
3.5.6 Horizontal Wells 15
3.5.7 Formation Damage 16
3.5.8 Perforation And Deviation Skin 16
3.5.9 Multiple Zone Completions 18
3.5.10 Gravel Pack Completions 18
3.5.10.1 Cased Hole Gravel Packs 18
3.5.10.2 Open Hole Gravel Pack 19
3.6 STIMULATION 20
3.6.1 Selection Of Stimulation Candidates/Methods 22
3.6.2 Acidising 24
3.6.2.1 Breakdown Treatments 25
3.6.2.2 Scale And Corrosion Removal 25
3.6.2.3 Matrix Acidising Of Sandstones 26
3.6.2.4 Matrix Acidising Of Carbonates 28
3.6.3 Types Of Acids And Additives 29
3.6.4 Acid Placement Methods 29
3.6.4.1 Bullheading 29
3.6.4.2 Coiled Tubing/Snubbing Workstring 30
Section 3 – Well Performance

3.7 VERTICAL LIFT PERFORMANCE 31


3.7.1 Correlation Selection 32
3.7.1.1 Fancher-Brown Correlation 33
3.7.1.2 Hagedorn And Brown 33
3.7.1.3 Duns And Ros 33
3.7.1.4 Orkiszewski 33
3.7.1.5 Beggs And Brill 34
3.7.1.6 Ansari 35
3.7.1.7 Petroleum Experts Correlations 35
3.7.1.8 Gray 35
3.7.2 Use Of Correlations For Well Dying Prediction 35
3.7.3 Tuning Pressure Drop Correlations 36
3.7.4 Natural Flow 37
3.7.5 Shut-In Conditions 37
3.7.6 Artificial Lift 39
3.7.6.1 Gas Lift 39
3.7.6.2 Electrical Submersible Pumps 43
3.7.6.3 Jet Pumps 44
3.7.7 Heat Transfer 46
3.7.8 Erosion 47
3.7.9 Tubing Selection 49
3.7.9.1 Natural Flow Production 49
3.7.9.2 Tapered Strings 51
3.7.9.3 Pumped Production 51
3.7.9.4 Gas Lift 52
3.8 WELL AND FIELD OPTIMISATION 53
3.8.1 Practical/Theoretical Optimisation 54
3.8.1.1 Practical Methods 55
3.8.1.2 Theoretical Methods 55
3.8.1.3 Combined Method 55
3.9 REFERENCES 56
Section 3 – Well Performance

3 WELL PERFORMANCE

3.1 Summary
The contents of this section give engineers an insight into the aspects of engineering which enable a
well’s performance to be assessed. It relates these to the computer programmes available and
normally used to determine well performance. The wells performance is critical for the completion
engineer to design the completion to cater for the pressures, flow rates, temperatures and producing
conditions to meet with the planned life of the completion.

3.2 Introduction
The purpose of this section of the manual is to define well performance and how it is assessed and
modelled. It has been written from a users viewpoint rather than any theoretical perspective. The
more theoretical aspects are covered in many available industry publications.
This section of the manual should be used in conjunction with the well performance software
‘Prosper’ Manual and the Near Wellbore Completion Manual. This section does not attempt to
cover any of the mechanics of using Prosper, rather what issues need to be addressed when
attempting accurate and representative well performance modelling. There are some areas of
overlap, between the Prosper manual and this section, however these have been included to
emphasise particular points.
The Prosper Software can be accessed at:
Prosper Software

The Near Wellbore Completion Manual can be accessed at:


Near Wellbore Completion Manual

3.3 Definition Of Well Performance


Following the review of the reservoir and assessing what is known about it, the planning process
then focuses on well design. A major component of this is the selection and design of the
completion to be used. There are a large number of tools and techniques available to the completion
engineer to address the specific properties of the reservoir and to maximise the productivity of the
planned wells. However, central to the entire well design process is the means of defining well
performance.
The classic approach to expressing well performance is a graph of the well bottomhole flowing
pressure versus the produced fluid rate. The following example is for an oil well:

Page 1
Section 3 – Well Performance

Figure 3.1 – Typical Inflow And Outflow Plot

Superimposed on the graph are two well properties indicated by blue and red lines.
The blue line represents what the reservoir will deliver to the well and is termed the inflow
performance relationship (IPR). The concept was introduced in the 1920s with the development of
the bottomhole pressure gauge. From the gradient of this line the productivity index, or PI, can be
calculated and is quoted as stock tank barrels per day per psi drawdown. This gives a means of
comparing productivity between wells.
The red line defines what the bottomhole flowing pressure has to be in order for the well to produce
at the rate on the bottom axis. This relationship accounts for the pressure required to get the
produced fluids up through the tubing used to construct the well and is termed the vertical lift
performance (VLP) or tubing performance. If the datum point for each system is the same, the point
at which the two lines intersect defines the rate at which the well will flow and the bottom hole
pressure at this point. This is an example of nodal analysis, so called because the whole system
performance is described by two independent subsystems. The two subsystems are upstream and
downstream of a common node. It is possible for this node to anywhere in the system, for example
when looking at facilities performance as well, it is often useful to define the node as the wellhead
or where the flowline meets the manifold. Such concepts are explored more in the well optimisation
section 3.8.
Having defined the means of describing well performance, the design of the well can be considered
in order to optimise this. In this context, optimisation does not mean the same as maximisation.
High rate wells can certainly be designed, however, this may have implications on the performance
of the reservoir throughout field life that may reduce overall hydrocarbon recovery. The key issue is
that a means of reviewing well performance is available to contribute to the overall review of the
best production strategy for the reservoir.

Page 2
Section 3 – Well Performance

As shown in the graph above, there are two parameters to consider when assessing well
performance:
• The inflow into the well (described by the IPR)
• The outflow through the well to surface (described by the VLP)

An analogous relationship exists for injection wells.

3.3.1 Inflow Performance Relationship


The reservoir will have specific properties defined by the geological processes occurring as it was
laid down. Similarly, the hydrocarbon fluid will have properties defined by its source and the
environment in which it was converted from Kerogen. The issues the completions engineer has
control over are:
• Geometry of the well
• What sections are allowed to flow
• Order in which this occurs
• Selection of the interface between the reservoir and wellbore

For example, a horizontal well may be selected with flowing sections towards the toe end and
beginning heel end. The toe may be produced first to ensure maximum clean-up from these sections
through a wire wrapped completion to prevent the production of sand from the reservoir.
As well testing became common a large effort took place within the industry to define mathematical
equations that describe the IPR based on reservoir and fluid properties. From this work, reservoir
and fluid parameters have been identified that control the inflow. In turn these can be used in order
to predict the potential IPRs of new wells as a basis of effective well design programs. This chapter
will discuss the various IPR relationships available and criteria for their applicability to a given
well. How the appropriate IPR can then be used to evaluate the potential well performance for a
specific well design will then be discussed with a proposed methodology.

3.3.2 Vertical Lift Performance


The vertical lift performance (VLP) or tubing performance is controlled by both the properties of the
produced hydrocarbon and the construction of the well. Whilst the completions engineer has no
control over the hydrocarbon characteristics, defined by the pressure - volume - temperature
relationship or PVT, they have options to use different tubing sizes and artificial lift methods in
order to optimise production. The two most common methods of artificial lift are using gas lift,
where produced gas is injected at a series of points down the well, or running a pump of some sort,
(refer to section 4). Both methods have the effect of reducing the hydrostatic column pressure in the
well. Without this to overcome, produced fluids can be far more easily flowed to surface.
The dimensions of the tubing that the produced fluids flow through to surface:
• The material used
• The degree of corrosion
• Other factors such as scale build up also effect the VLP

Clearly the completion engineer has far more control over these along with the wellhead flowing
pressure. By considering all of these parameters, the most appropriate well design can be identified.

Page 3
Section 3 – Well Performance

3.3.3 Overall Pressure Drops


It is useful to take an overall view of the pressure drops in the field/well. This enables the real
influences on the field/well performance to be identified. For example in a gas well in a tight
reservoir, the majority of the pressure drops are frictional in the reservoir. In an oil well, hydrostatic
pressure drops in the tubing are more important. This allows both the effort in prediction to be
focused on the areas where it maters, but also the areas where the completion designer can have an
impact on improving performance, (refer to Figure 3.2).

Figure 3.2 – Generalised Pressure Drops In Hydrocarbons Production

3.3.4 Well Design For Life Of The Well


A major issue to be considered during the well and completion design process is how the well
should perform for the entire period of its operational life. Factors that impact on this are:
• The performance of the reservoir throughout the field life. For example, pressure
support (or lack of it) or water/gas coning.
• Reservoir constraints. For example on the Harding field, the reservoir is so conductive
that rates in excess of 40,000bpd could be achieved. However, gas and/or water coning
would occur at these rates and, therefore, the wells do not need to be designed to handle
these rates. Likewise sometimes bottomhole pressures are deliberately maintained above
the bubble point in order to prevent relative permeability effects.
• The fluids handling and how this may change, e.g. there is no point in designing a large
well to cope with high GORs if the wells will have to be constantly choked back
because the gas cannot be managed at surface.
• Production problems. Scales, asphaltenes, wax etc. all impact the well design and
productivity. Asphaltene deposition occurs at a certain pressure and, therefore, the
bottom hole pressure is maintained above this point in order to prevent asphaltene
depositing in the reservoir or at the bottom of the well (e.g. perforations).
• The likelihood of workovers. This will allow for a tubing size change or implementation
of artificial lift.

Page 4
Section 3 – Well Performance

3.4 Pressure/Volume/Temperature
The Pressure Volume Temperature (PVT) relationship describes how a fluid behaves under
changing conditions. With an accurate PVT relationship, the density, viscosity and gas-oil ratio for
the fluid under expected pressures and temperatures can be reliably extracted. This is then used to
determine the inflow performance and more importantly the tubing performance.
As a hydrocarbon fluid is produced, the temperature and pressure changes. These changes will
initially only cause the oil or gas to change viscosity and density. At a certain point, however, the
fluid will change from a single phase to two phase. For a black-oil fluid, the gas will start to come
out of solution at the bubble point. For a condensate, condensate will start to come out of solution at
the dew point. For a given fluid composition, these points (the saturation points) define the phase
envelope. Outside the phase envelope, the fluid is single phase, inside the fluid is two phase. The
critical point is the point on the phase diagram where to the left the fluid that first comes out of
solution is a gas. To the right, the fluid that first comes out of solution is a liquid. Figure 3.3 shows
an example of a phase envelope.

Figure 3.3 – Example Phase Envelope

Page 5
Section 3 – Well Performance

The definitions of the different fluids are:


• A black-oil or volatile oil is one where with a drop in pressure, gas will come out of
solution. The difference between a black oil and a volatile oil is purely arbitrary and
relates to the higher GOR and formation volume factor (FVF) of a volatile oil compared
to a black oil. The FVF for a volatile oil will be above approximately 1.5, below this it
will be a black-oil.
• A retrograde condensate is one where with a drop in pressure, liquid will first come out
of solution. Note most retrograde condensates will exhibit the behavior where as the
pressure is reduced further, the liquid may vaporise again.
• A gas is where a drop in pressure will not result in the phase envelope being crossed.

These definitions apply to the conditions in the reservoir (isothermal). In the tubing, cooling will
occur and therefore the fluid will almost certainly cross the phase envelope at some point. Therefore
a gas will produce some liquid (condensate) before it reaches the surface.
It is possible to have all of these fluids with the same composition, it is just the initial pressure and
temperature that may change. For example in the phase envelope, (refer to Figure 3.3), a reservoir
temperature of 650°F or below would result in a black or volatile oil. A reservoir temperature of
between 650°F and 858°F would be a retrograde condensate and a reservoir temperature of above
858°F would be a dry gas. Note: In this example a black-oil fluid is likely as most reservoir
temperatures are below 650°F.
Below the saturation pressure, the proportions of the phases will change. This can be examined on
the phase envelope plot (as in Figure 3.4).

Figure 3.4 – Fluid Behaviour From Reservoir To Surface

As the fluid proportions and properties change, this will clearly affect the well performance.

Page 6
Section 3 – Well Performance

The reservoir or near wellbore performance will be effected by:


• Viscosity
• Expansion or contraction of the fluid (included in the formation volume factor)
• Any relative permeability effects due to liquid or gas break-out (note: Prosper cannot
handle gas relative permeabilities and these will have to be estimated separately)

The tubing performance on the other hand will be effected by:


• Density of the fluid(s) - accounted for by the FVF and oil/gas gravity
• Proportion of gas to liquid (the GOR or CGR)
• Viscosity (to a lesser extent)

The PVT data is therefore critical to the well performance predictions. Historically most errors in
well performance prediction have been attributed to poor or inaccurate PVT data.
Regardless of which PVT system is used to describe the fluids within Prosper, it is vital that the
original laboratory PVT data is both representative of the reservoir fluids and covers well
performance issues as well as reservoir performance issues:
a) Ensure that the well has cleaned-up adequately (as evidenced by stable flow with a
steady GOR and water cut).
b) There are two options, either, bottom hole single phase samples, or separator samples,
recombined according to the GOR. Sampling at the separator introduces more errors
than downhole sampling as the proportions of oil and gas have to be accurately
measured. Whichever method is chosen, ensure that the fluid is single phase at the
sandface, otherwise condensate or gas breakout may make the produced fluids
unrepresentative. Sampling of fluids where the reservoir pressure is close to the bubble
or dew point is always going to be error prone and this uncertainty must be
acknowledged.
c) The value of accurate samples is huge. For example, on Pompano, approximately $20
million US could have been saved if the paraffin content had been accurately known and
the expensive TFL completions avoided. The only oil sample was unfortunately ‘lost’ in
the laboratory.
d) Multiple samples should be checked against each other. The bubble point is the most
useful consistency check.
e) The PVT analysis should include a constant composition expansion experiment for
conditions between the bubble point and separator pressure. Note: The reservoir
engineer is more likely to be interested in differential expansion where at each stage of
the expansion, the produced gas is removed and, therefore, the composition changes.
The two methods will produce different GORs, and Formation Volume Factors and,
hence, different predictions about well performance. Constant composition is more valid
in the tubing as the oil and gas will be in constant contact with each other.
f) Any PVT model may be ideal for the reservoir or the facilities, but not the tubing. For
example, the correct fluid density may not be critical for either the reservoir or the
facilities model. However, for the tubing the correct fluid density over a large range of
pressures and temperatures is absolutely vital. Likewise, the reservoir fluids do not need
to include a large sensitivity to temperature and facilities correlations do not necessarily
need to cover a large range in pressure.

Page 7
Section 3 – Well Performance

3.4.1 Direct Use Of Laboratory Data


It is possible to use laboratory data, if it is available, over the expected range of pressures and
temperatures. Interpolation will be performed between pressure and temperature data points, but so
long as there are sufficient data points and that for each temperature, there is a data point at the
point bubble, the errors will not be substantial. The real problem comes from the inflexibility this
approach introduces. It does not allow the user to alter the composition to any extent, i.e. gas lift, or
changing GORs cannot be incorporated.

3.4.2 Equation Of State


An Equation of State (EoS) is a mathematical method for modelling a fluid based on the
components in the fluid. As a reservoir hydrocarbon contains hundreds of different components, the
EoS is usually limited to the major components or groups of components. The properties of these
components and these ‘pseudo components’ must be specified. The equations of state were
originally developed for pure substances but through time their use was extended to mixtures. With
mixtures (e.g. reservoir fluids) some method of introducing a measure of the polar and other
interactions between pairs of dissimilar molecules is required. The binary interaction coefficients
change the ideal Equation of State to match the reality of many mixtures.
The Equation of State used by Prosper is the Peng-Robinson (P-R). The Prosper programme can be
accessed at:
Prosper Software

with templates available at

Integrated Assest Modelling

It has been well established that the capability of two-parameter equations of state, such as P-R, in
predicting the liquid density can be improved by introducing the ‘volume shift’ parameter. The
method is particularly attractive because it does not change the predicted phase equilibrium, but
affects the phase densities by shifting the volume axis. The accurate prediction of density is vital for
tubing performance predictions. However, this technique is arbitrary and non-rigorous/unscientific.
The inclusion of the third parameter in EoS, may deteriorate the predicted density at some
conditions. This could occur for systems with high concentrations of supercritical compound(s),
particularly methane. Moreover, using a constant shift parameter for light hydrocarbons, the phase
densities cannot be predicted accurately.
With an accurate EoS, in theory, the properties of the mixture can, therefore, be calculated for any
pressure and temperature.
Some of the specific issues concerning equation of state models are:
a) They may give the impression of being highly accurate. Like any model, they are only as
good as data they are based on.
b) They are only valid for a fixed input composition. Although they are very good at
predicting the phase to phase transfers vital for separator performance, they require a
feed composition. Therefore, each EoS is only valid for a single GOR. They are,
therefore, difficult to use for gas lift completions and any model where the reservoir
GOR may change (e.g. gas cap expansion).
c) As the equations of state are often produced specifically for reservoir or facilities
engineers, they may, therefore, be invalid for tubing conditions. This is particularly the
case if they have been tuned excessively using volume shift or binary interaction
coefficients. In the example below (refer to Figure 3.5), the bubble point is ridiculously
Page 8
Section 3 – Well Performance

high at low temperatures and may indicate other potential problems at likely conditions.

Page 9
Section 3 – Well Performance

d) It is known that many EoS models poorly represent the liquid density of fluids and
errors of around 5% are commonplace. Errors in liquid densities would not significantly
effect either reservoir models or most facilities models, but will seriously effect the
tubing performance.
e) An EoS model is not particularly accurate at determining the viscosity of fluids. This
can be improved by the entering of ‘critical volumes’ for each component.

Figure 3.5 – Example Of Unreliable PVT Caused By Excessive Tuning

3.4.3 Empirical Correlations - Untuned


Various correlations are available that attempt to predict a fluids properties based on the fluids oil
and gas gravity, and the GOR. Each correlation has been designed for a particular range of fluids
and is purely empirical in nature.
The correlations available in Prosper are:
Black Oil FVF, GOR
• Glaso
• Standing
• Lasater
• Vazquez-Beggs
• Petrosky et al

Black Oil Viscosity


• Beal et al
• Beggs et al
• Petrosky et al

Page 10
Section 3 – Well Performance

Gas Viscosity
If a gas is chosen, all the condensate drop out is assumed to occur in the separator and not in the
tubing. Therefore, any condensate hold-up problems cannot be analysed.
• Lee et al
• Carr et al

Any of these correlations can be used directly without tuning so long as the GOR and densities are
known. It would be possible to examine each of these correlations in turn to see which correlation
was developed to model a similar fluid to one in question, however, this implies a complete lack of
understanding of the reservoir fluids. At the very least, the bubble point should be matched and the
correlation chosen accordingly.

Retrograde Condensate
Prosper has its own Retrograde condensate empirical model that is capable of predicting fluid
properties and condensate gas ratios (CGRs) below the dew point. Even more than a black oil
model, this model is only realistic if it is matched to real conditions.

3.5 Inflow Performance


Efficient well design requires a proper understanding of the reservoir inflow performance (IPR) and
of how it is affected by the near wellbore completion. Completion, well kill and workover
techniques that minimise damage or enhance performance should be adopted, especially where the
objective is to maximise well deliverability.
The attainment of a near optimum, or stimulated, IPR requires a proper understanding of the causes
of skin effects and the application of techniques to avoid damage and to enhance flow efficiency,
however, a trade-off must often be made between maximising deliverability and minimising
operational problems, lifting costs and capital expenditure. Production forecasts must be adjusted to
reflect resultant inflow capabilities. By quantifying the effect of the various options on near
wellbore performance, it is possible to develop economic justification for improved operational
procedures and completion methods, or for workover and stimulation operations.
Formation damage is a major cause of production deferment, therefore, by understanding the
damage processes and associated risks, steps can often be taken to avoid damage at minimal cost to
the overall development. Where well testing results indicate that damage has already occurred,
analysis of the most probable causes allows more efficient design of measures to bypass, or remove,
the damage.
Perforation is one of the most critical steps in the completion process, which can have considerable
impact on the inflow efficiency. With the proper choice of perforation interval, charges, gun system
type, perforating method and drawdown, it is possible to consistently get close to the idealised
performance of an open hole completion, especially if tubing conveyed, underbalanced perforating
techniques are used. However, to be cost effective, the designer must select a technique that is
appropriate to meet with the production objectives and operational environment. Guidelines have
been developed to aid the less experienced engineer in making these judgments.
Stimulation techniques, such as acidizing and fracturing, can be used to remove, or bypass, damage
areas and to enhance the natural IPR. However, many stimulation treatments are expensive and
involve a degree of risk, so that a basic understanding of the processes is required to select the most
appropriate treatment. Design and operational procedures are discussed and quality control
requirements stressed.

Page 11
Section 3 – Well Performance

Sand problems plague many developments in shallow, unconsolidated formations and in over-
pressured, shaley sands. The techniques to assess the risks of sand production and possible
strategies to avoid the problem are, therefore, an integral part of development planning. Where sand
control must be installed, cased hole gravel packing is usually selected. This completion technique
will always impact on the well’s inflow capacity; however, the degree of damage incurred depends
heavily on the design and installation of the completion.
Part of the completion interval may, at some time, need to be isolated to prevent excessive gas or
water production, (refer to section 14.8). Various mechanical and chemical techniques are available
for this, and selection of the most appropriate option will impact, not only the costs, but also the
probability of success. Thus, the design and modification of the near wellbore completion is a
fundamental element of production engineering that can have substantial impact on well
deliverability, completion and workover costs, and on the daily operating costs.

3.5.1 Radial Inflow Equation And Skin


The fundamental equation of liquid flow from the reservoir into the completion is the radial inflow
equation or Darcy equation:
0.00708kh (Pr − Pw )
Q = Equation 1
 r  
Bµ  log  e  + S 
  rw  

Where:
Q = flow rate (bpd)
k = permeability (md)
h = reservoir thickness (ft)
Pr = reservoir pressure (psi)
Pw = bottom hole flowing pressure (psi)
B = formation volume factor or FVF (dimensionless)
µ = viscosity (cP)
re = drainage radius
rw = wellbore open hole radius
S = skin (dimensionless). The skin is a measure of any additional pressure drops that
are present if the well
This defines the flowrate for a given pressure drop for a vertical well fully penetrating a reservoir
interval with a constant pressure outer boundary. The skin is a catch all term for any additional
pressure drop. The kh and skin can be determined from well testing data.
This equation can be simplified even further, if all the constants are grouped together:
Q = J (Pr − Pwf ) Equation 2

Where:
J = productivity index or PI.
These equations are only applicable above the bubble point and for oil wells.

Page 12
Section 3 – Well Performance

In Prosper, this equation uses the Dietz shape factor and drainage area rather than the reservoir/no
flow boundary radius. The Dietz shape factor accounts for the fact that the drainage area may not be
circular. Each shape (round, square, rectangular etc.) will have its own number - for example a well
in the centre of a circular drainage area has a Dietz shape factor of 31.6. A table of shapes and
factors is included in the Prosper help files.
The skin combines various different aspects of the well performance:
• Formation damage skin
• Perforation skin
• Partial completion skin (i.e. if not all of a reservoir interval has been completed)
• Deviation skin
• Stimulation effects
• Reservoir heterogeneity skin
• Multiphase flow effects
• Darcy effects of sand or gravel filled perforations
• Non Darcy (or turbulence) effects associated with any of the above

Note: It is not possible to treat any of these effects in isolation. For example a well
perforated in a small proportion of the reservoir will make any poor perforation
effects more pronounced.

More details on the inflow performance can be found in the BP Near Wellbore Performance Manual
at:
Near Wellbore Completion Manual

3.5.2 Vogel Inflow Performance


This empirical equation is used for producing wells below the bubble point:
  Pwf   Pwf 
2

Q = Qb + (Qmax − Qb )1 − 0.2  −    Equation 3
  Pb   Pb  

Where:
Qb and Pb are a flowrate and a corresponding bottom hole bubble point pressure.
In order to use this correlation test data must be available.

3.5.3 Fetkovich
Q = ( 2
J o Pr − Pwf
2
) Equation 4

Jo is the PI above the bubble point adjusted for any relative permeability effects. This empirical
relationship has similar application to the Vogel IPR and should be used below the bubble point.

Page 13
Section 3 – Well Performance

3.5.4 Jones
(P r − Pwf ) = aQ 2 + bQ Equation 5

This empirical equation is an expansion of the Darcy inflow equation to include rate dependent or
non Darcy effects and is, therefore, applicable to oil and gas wells. The b term is the same for the
Darcy equation. The a term accounts for the flow velocity:
2
0.0005359bo ρ
a = 2
Equation 6
k 1.201h p rw

Where:
hp = is the completed interval.

Note: The Jones equation does not account for any flow convergence around
perforations or for any case that is not a simple open hole completion.

3.5.5 Hydraulically Fractured Wells


A fractured well performance depends on getting the fluids to the fracture face and along the
fracture. The first is dependent on the reservoir permeability and the fracture length, the second is
more dependent on the fracture conductivity.
The (dimensionless) fracture conductivity (Fcd) is dependent on the fracture width and proppant
permeability:
K fWf
Fcd = Equation 7
KX f

Where Kf and K are the permeabilities of the proppant and reservoir respectively, Wf is the fracture
width and Xf is the fracture half length (distance from the well to the tip of the fracture). Fracture
widths are usually in the order of 0.15 to 0.75 inch. A fully effective fracture should have an Fcd >
10. Note: The permeability of proppant is much less under real conditions than measured in the
laboratory due to compression, gel residues, fines entrapment, long term crushing, non Darcy effects
or contamination. A useful rule of thumb is to reduce the lab derived permeability value by a factor
of 10. The fracture width will also be reduced by embedment of the proppant into the rock. In severe
cases, this can lead to a severe reduction in the effectiveness of the fracture and techniques for
extending the width of the fracture to compensate (i.e. tip-screen out techniques) should be used.
The model used within Prosper is a transient model and accounts for the early time being dominated
by fracture conductivity and the later time being dominated more by overall reservoir performance.
An example of the effect this has is shown in Figure 3.6.

Page 14
Section 3 – Well Performance

Figure 3.6 – Propped Fracture Transient Performance Example

In this example, only after twenty days has the well settled down to close to steady state
performance.

3.5.6 Horizontal Wells


Prosper uses three models for horizontal wells.
The first is the horizontal well model of Kuckuk and Goode. It assumes a horizontal well in a closed
rectangular drainage volume bounded by sealing surfaces. The pressure drop in the wellbore itself is
not accounted for. This is acceptable for cases where the drawdown is an order of magnitude greater
than this frictional pressure drop.
The second model is similar except that a constant pressure upper boundary is assumed. This would
be the case for a well with an active gas cap. Such a case will give higher productivities than an
upper no flow boundary.
The third method allows for pressure drops along the completion interval itself. A number of
horizontal well models are available including the preferred ones of Goode and Wilkinson2 and
Kuckuk and Goode. Both of these formulations are valid even when the horizontal wellbore length
approaches the length or width of the reservoir. The method of Goode and Wilkinson is less
recommended when frictional pressure drops are included, but is preferred when friction is ignored
(infinite conductivity). Note: In order to use Goode and Wilkinson with Infinite conductivity, set the
roughness to 0 (this turns off friction), rather than assuming friction for a smooth pipe. The other
horizontal well models are included for completeness and should not be used in practice.

Page 15
Section 3 – Well Performance

3.5.7 Formation Damage


Productivity damage is one of the major causes of production deferment. Damage can be defined as
‘any barrier to, or effect within the confines of the near wellbore area or wellbore completion
interval, that causes lower productivity than would be expected from an ideal reservoir with an ideal
completion’. Hence, plugged perforations, screens or gravel packs, blocked pores due to solids or
emulsions, reduced absolute or relative permeability, or increased water saturation in the near
wellbore area, all constitute productivity damage.
The most common causes of damage are from drilling, completion and workover operations carried
out under overbalanced conditions resulting in an influx of solids and fluids into the formation. The
effects of this can be calculated for both open hole and perforated wells.
A detailed discussion on the causes of formation damage can be found in the Near Wellbore Manual
found at:
Near Wellbore Completion Manual

Damaged can often be minimised or avoided by the use of solids free or non-damaging wellbore
fluids, adopting underbalanced perforating techniques or alteration of produced fluid properties to
prevent the formation of emulsions, solids precipitation, etc. and is preferred. Post-damaged
formations can often be treated to reduce the damage by conducting stimulation operations such as
bypassing (through perforating or fracturing) or acidising, scale removal, etc. (refer to section 3.6).

3.5.8 Perforation And Deviation Skin


Prosper uses three options for models to calculate the skin in a deviated, perforated well. Locke’s
method is the original model and is included in Prosper for completeness. Mcleod’s method
includes an option of over or underbalanced simplification. The crushed zone permeability is
reduced to 10% for over balanced perforating, and 40% for underbalanced perforating. This is rather
arbitrary and a more detailed discussion is found in the subsections of the BP Near Wellbore
Performance Manual including gun performance and perforating techniques which can be found at:
Near Wellbore Manual

Karakas and Tariq is a more general technique and is recommended for most applications. It can
deal with high deviations as long as radial flow is reached away from the wellbore and, therefore,
the well is not too long in comparison to the reservoir.
The input data for Karakas and Tariq is rather involved. The possible provenances of the various
data inputs are included in Table 3.1.

Page 16
Section 3 – Well Performance

Perforation Diameter This is the diameter in the rock (not the casing). It can be extracted from
the gun vendor with a correction for the rock strength. Alternatively, the
3
equation is D = EH N 80 (3.27 − 0.611n[UCS ]) , for deep penetrating
charges.
Where:
D = the perforation hole diameter
UCS = is the unconfined compressive rock strength
EHN80= is the entrance hole diameter in N80 steel (available
from perforation catalogues).
Shots per Foot For underbalanced perforating this will be the gun’s shot per foot. If
overbalance perforating with dirty fluids, the open perforation may only
be 0.1 - 0.25 the nominal shot per foot.
Perforation Length This is a property of the gun and charge, with corrections for confining
stress and rock strength. These corrections are vital and the equations will
be incorporated into Prosper but are available in THoR and from the gun
companies. Typical numbers are between 12” and 24”.
Damaged Zone Thickness This will depend on the mud and rock properties and how the wells are
drilled. Typical values are 3” to 12”. The actual value is not particularly
critical so long as the perforations extend behind this. The actual
thickness can be determined from filter cake tests on cores or estimated
from filtrate depth of invasion calculated from mud losses but this is
difficult to achieve in practice.
Crushed Zone Thickness From CAT scans and thin sections of perforated core, this is typically 0.5"
thick.
Crushed Zone Permeability Assuming an optimum perforation underbalance, this should be the same
as the formation permeability, otherwise a figure of up to half the
formation permeability can be used.
Shot Phasing This is a choice for engineers. Low values (60 - 90 degrees) will normally
give optimum perforation efficiency.
Deviation This is also a choice for engineers.
Penetration This is the fraction of the wellbore that is open to flow. Prosper assumes
that it is the top interval only that is completed, (i.e. from the top
boundary of the reservoir downwards).
Vertical Permeability There is some uncertainty around what figure should be used. In THoR
and KT Perf, there was a natural division between short and large scale
kv/kh. Short scale Kv/kh applied to the scale of the perforations i.e. on the
scale of a few inches. Large scale kv/kh applied to the effects of
anisotropy on the deviation or partial penetration skin. Prosper assumes
that the large and small scale kv/kh are the same. In a vertical fully
completed well, the small scale kv/kh should be used, but in a high angle
or partially completed well, the large scale effects may be more important
and the lower large scale kv/kh should be used. Therefore, some judgment
is required in order to pick the correct vertical permeability. This anomaly
should be corrected when the method of Wong and Clifford is included in
Prosper v.6.
Wellbore Radius This is the drilled open hole radius.
Table 3.1 – Prosper Perforating Input Data

Page 17
Section 3 – Well Performance

3.5.9 Multiple Zone Completions


Prosper has various options for looking at multizone completions. It can, either, include or exclude
the frictional pressure drop between each zone. Note: That although crossflow is included, it is only
in the wellbore and no fluid flow is allowed between different layers in the reservoir.

3.5.10 Gravel Pack Completions


Prosper can handle, both, cased and open hole gravel packs. It can also deal with open hole
completions with sand control but no gravel. The subject of gravel packing cannot be dealt with in
detail here, only the basic drivers on productivity and how to calculate it for a gravel pack
completion. The BP sand control guidelines contain much more detail at:
Sand Control Guidelines

3.5.10.1 Cased Hole Gravel Packs


A cased hole gravel pack is where screens are installed in a cased and perforated well. Gravel is
pumped to fill the volume between the casing and the screen and also, critically, the perforations.
The cased hole gravel pack pressure drops include those of a cased and perforated completion, (i.e.
phasing, perforation length and crushed zone permeability all have a contribution to the pressure
drop). However in addition, there will be a substantial pressure drop through the gravel in the
perforations and in the gravel between the casing and the screen.
The model is for cased hole gravel pack pressure drops, therefore incorporate the perforating
parameters discussed in 3.5.7. The following variables are required for the pressure drops in the
gravel:
a) Gravel pack permeability (this can be extracted from the table Table 3.2, although due
allowance must be given to downhole conditions, e.g. any debris or compaction).
b) Perforation diameter, this is critical and the reason why big hole charges (1” diameter)
are preferred.
c) Shots per foot, again this is critical 12spf is preferred (dependent on liner size).
d) Gravel pack length - in Prosper this is the distance from the sandface (not the casing) to
the screen OD. This means that the liner size is not included in any of the calculations
and, therefore, the validity of these calculations.
e) Perforation interval
f) Perforation efficiency

US Mesh Range Sieve Opening (mm) Permeability (Darcy)


6/10 2 – 3.36 2703
8/12 1.68 – 2.38 1696
10/20 0.84 – 2.00 652
12/20 0.84 – 1.68 630
16/30 0.589 – 1.19 250
20/40 0.42 – 0.84 171
40/60 0.25 – 0.42 69
50/70 0.21 – 0.297 45
Table 3.2 – Gravel Pack Permeability (for clean, uncompressed gravel)

Page 18
Section 3 – Well Performance

Note: The criticality is getting a large number of large diameter perforations filled
with high permeability gravel. In the event that the perforations are not
effectively packed, they will fill with reservoir sand. The permeability of the
loose reservoir sand is likely to be much lower than properly sorted and clean
gravel.

3.5.10.2 Open Hole Gravel Pack


An open gravel pack is where a screen is installed across open hole and the volume between the
screen and the open hole is packed with gravel. The same model can also be used to model an open
hole sand control completion where the formation collapses around the sand screens.
The productivity of an hole gravel pack will be dependent largely on the permeability of the gravel
or formation sand. If there is no gravel and the formation is heterogeneous, the collapsed sand may
contain significant fines or clay and the permeability will be reduced as well as these fines
potentially plugging the screens.

Page 19
Section 3 – Well Performance

3.6 Stimulation
Stimulation involves reducing the skin by removing, or bypassing, damage or improving the well’s
ideal IPR by creating an enlarged contact area between the wellbore and formation. As stated above,
formation damage is much easier to avoid than remove so many of the more sophisticated mud acid
stimulations are unnecessary incremental investments to compensate for poor design and execution
of the near wellbore completion, or of well kill or workover operations. Often attempts to achieve
relatively minor cost reductions in these operations (e.g., by not filtering base fluids) result in
significant damage that causes production deferment and necessitates much more costly and risky
remedial treatments.
As shown in Figure 3.7, the fact that a well has high deliverability does not necessarily imply that it
is undamaged. Conversely, if the offtake is limited by reservoir or regulatory requirements, it may
be possible to achieve the target by reducing the wellhead pressure and thus deferring stimulation
until increasing water cuts or pressure depletion mean that the target is no longer attainable by
natural flow. Therefore, wells with deliverability close to, or just below, their production targets are
often excellent candidates for testing and acidisation. Moreover, as shown in Figure 3.7 discussed in
the Near Wellbore Manual, re-perforation is often an alternative to acidisation.

IPR Curves
0 - Undamaged IPR (skin = 0)
1 - Reperf/Stimulated (skin = 2)

2 - Damaged IPR (skin = 8)


50% Perforations Plugged

3 - Damaged IPR (skin = 16)


Deep solids invasion via
Perforations and 50% Perforations
Plugged

TPR Curves
0 – 1,000 psi WHD Pressure
1 – 1,500 psi WHD Pressure

Figure 3.7 – Prosper Analysis Of The Effect Of Damage, Re-perforation, Acidisation On Well
Deliverability

Page 20
Section 3 – Well Performance

The three most common stimulation techniques are:


• Acidizing
• Propped fracturing
• Acid fracturing

Acidizing is intended to remove damage in the near wellbore area, thus potentially reducing the skin
factor to around zero (+1 to –3).
There are a number of chemical treatments which can be used singly or in combination to remove
damage. These include:
• Organic solvents to remove wax and asphaltenes.
• Hydrochloric acid (HCl) to dissolve calcium carbonate, dolomite and iron compounds or
to break gels.
• Hydrochloric acid (HCl/HF) to dissolve fines and clays in the near wellbore area (<6”).
• Self generating mud acid, flouboric acid or buffeted mud acid for deeper damage to
achieve adequate penetration.
• NaOCl (bleach) to remove bacterial slimes from water injection wells.
• Ethene diamine tetra-acetic acid (EDTA) or scale convertors followed by HCl to remove
calcium sulphate scale.
• Flouboric acid and clay stabilizer to seal mobile fines.
• Ethyelene glycol monobutyl ether (EGMBE) mutual solvent to treat wettability and
relative permeability problems.

Fracturing creates a highly permeable flow channel with a large contact area which can result in a
significant negative skin factor (-3 to -7) and is used to achieve commercial rates from low
permeable formations, (refer to Figure 3.8).
In sandstones, the fractures are propped open with high permeability coarse frac sand (or gravel). In
carbonates, high permeability channels can sometimes be etched into the fracture face to create an
even more conductive room-and-pillar arrangement after the fracture closes. However, this acid
fracturing technique is not always effective, especially in soft chalks or naturally fractured
dolomites, so propped fractures may be more attractive in some carbonates.
For more detailed descriptions and applications for these refer to the Near Wellbore Manual at:
Near Wellbore Manual

Page 21
Section 3 – Well Performance

Figure 3.8 – Prosper Analysis Of Effect Of Fracture Half Length On Well Deliverability

3.6.1 Selection Of Stimulation Candidates/Methods


The best acidisation candidates are good wells that are producing below their potential (ie damaged
wells in which the actual IPR is less than the theoretical ideal IPR and deliverability less than rate
limitations). Acidisation will not help a poor well which has a very low kh or high oil viscosity, or
one that lacks reservoir pressure. Most acid jobs are designed to remove perforation plugging or
induced formation damage, which is particularly common in naturally fractured and vugular
reservoirs; in sandstones containing large amounts (>10%) of swelling or migrational clays loosely
attached to the sand grains; or where the well fluids were of poor quality. However, some of these
shaly, or dirty, sandstones exhibit a migrating fines problem, especially during initial clean-up and
water breakthrough (this is often termed ‘natural damage’), which necessitates a regular programme
of periodic acid stimulations.
Naturally fractured and high permeability carbonates nearly always require a clean-up acid job to
bypass deep drilling and perforation damage and to fully reopen the fractures. Once this is removed,
they will demonstrate a natural negative skin (-2 to -5).
Low permeability reservoirs (kgas <2.5 md, koil <25 md) always require fracturing, although short,
wide fractures may be beneficial in moderate permeability zones (10 to 75 md), or for bypassing
deep damage. The lower the permeability, the larger the fracture required, so that massive hydraulic
fracturing (MHF) is used for tight (micro-darcy) gas sands.
The key considerations in selecting stimulation candidates are as follows (see also Table 3.3):
1. Define the maximum allowable, feasible or desirable offtake rate.
Page 22
Section 3 – Well Performance

2. Review well inflow and outflow performance (as discussed in the Near Wellbore
manual and sections 3.5 and 3.6 of this manual) to confirm that the IPR is the major
or most easily corrected constraint to well deliverability. Comparison of the
expected and actual IPRs will identify potential stimulation candidates. The effect
of negative skin should also be considered for low permeability zones (k <50 md).
Ideally, potential stimulation candidates should be identified, or confirmed, from
well test data showing low permeabilities or high skin factors.
3. Review the remaining reserves and economic attractiveness of accelerating
production to confirm expenditure is justifiable and to scope budgetary limitations.
4. Determine the most probable cause of the observed well inflow performance and
evaluate damage mechanisms.
5. Review workover options and select the most appropriate ones for economic
analysis:
a) Reperforation:
-Through-tubing.
-TCP at high shot densities.
b) Chemical dissolution:
-Clean-up or matrix acidisations.
-Combination treatments pumped sequentially.
c) Hydro-mechanical jetting with coiled tubing using:
-Selective acidizing packers.
-Abrasajets and other jetting tools.
d) Fracture stimulation:
-“Skinfrac” to bypass deep damage.
-Standard frac to enhance IPR on low permeability rock.
-MHF to enhance IPR on tight gas.
6. Estimate treatment and workover costs and develop production forecasts for two or
three options.
7. Run economics to select the most attractive treatment.
8. Test chemical compatibility and define QA/QC procedures for treating fluids.

The remainder of this chapter will focus on the conventional stimulation techniques of acidisation
and fracturing; however, further discussion of other solvent treatments can be found in section 5 on
Production Chemistry.

Page 23
Section 3 – Well Performance

Candidate Property Method


Formation
- HCl or acid frac (or propped frac if uniformly etched or
• Carbonates
very weak
• Sandstones - clean-up acid, mud acid, propped frac
Formation Permeability
• Moderately high (>75md) - acidize (or mini-frac)
• Low (<10md) - fracturing
• Very low (<1md) - MHF
Depth Of Damage
• Plugged perforations or screens - acidize or re-perforate (or jetting)
• Shallow (<6”) - conventional acidisationn or re-perforate
• Moderate (6” to 3ft) - retard or self-generating acid
• Deep (>3ft) - fracturing
Fracture Barriers/Zone Thickness
- better frac and acidisation results; stage acid jobs on
• Thin zones (<30ft)
zones >100ft
• Thick zones (>100ft) - limited entry, sequential treatment, or positive diversion
- diverter selection and staging
- bottom of frac preferentially packed
• Height control problems - MHF
- fracs in techtronically disturbed areas (e.g. near Rockies)
- high pump rates with viscous fluids
- thin shales
- silty shales
Proximity To Contacts (OWC, GOC)
• No barriers - acidize or mini-frac
• Limited barriers - limit job size
Table 3.3 – Key Components In Selecting Stimulation Method

3.6.2 Acidising
Acidisation is the main stimulation technique for high permeability rocks (>50 md); and is used to
remove damage from the near wellbore area by breaking open plugged perforations and pore
throats; by dissolving or suspending the plugging materials; or by bypassing the damage by
dissolution of the rock matrix (in carbonates). Little benefit can be gained from attempting to
acidize an undamaged formation. This can easily be demonstrated by considering the effect of
increasing rw slightly (i.e. 100% removal of the rock) in the radial flow equation 1.
However, as shown in Figure 3.7, damage can have a dramatic effect on productivity. The impact of
damage and its removal can conveniently be estimated using equation 8.
J1 7 + S '2
= Equation 8
J2 7 + S '1

Page 24
Section 3 – Well Performance

Thus, in matrix acidizing, the intent is to restore the natural permeability, or to reduce non-Darcy
flow effects, in the region very close to the wellbore. If there are any significant amounts of reactive
material, the large surface area of a porous system means that effective acid stimulation is limited at
bottom hole temperatures to 3 to 6” (75 to 150mm) in sandstone, and 12 to 24” (300 to 600mm) in
carbonates, unless special formulations are used. This means that reperforation with high powered
guns can be evaluated as an alternative method of dealing with damaged sandstones.
There are essentially five main types of acid treatments:
• Breakdown treatments
• Scale removal
• Matrix acidisation of sandstones
• Matrix acidisation of carbonates
• Acid fracturing (addressed in section 3.6.2)

3.6.2.1 Breakdown Treatments


This involves small volumes (25 to 50gals) of weak HCl. The mechanism is a pH and hydraulic
shock combined with an increase in water saturation to remove fines. Little or no dissolution may be
involved. A small percentage of surfactant is usually added to aid in wetting and transportation of
the fines. This is also a common pretreatment technique for both matrix acidisation and fracture
stimulations. This can be applied to both sandstones and carbonates.

3.6.2.2 Scale And Corrosion Removal


Scale precipitation is a common oilfield problem, particularly in waterflooding operations. Calcium
carbonate can be deposited in producers, where decreasing pressures reduce the stability of
bicarbonates; and in water injection wells due to the temperatures increasing downhole. However,
the mixing of incompatible waters is also a major cause of scaling. Calcium carbonate is the most
common type of oilfield scale and the most difficult to predict theoretically.
Calcium carbonate is generally highly soluble in acid, particularly if deposited quickly as a porous
mass. However, the thicker, harder deposits that are laid down in layers need additional contact time
and, particularly, acid agitation to aid dissolution. Therefore, stage washing (i.e. starting and
stopping the pump) or use of coiled tubing to jet the acid onto the scale is useful. A prolonged soak
(3 to 12 hours) can also be beneficial; however, the main problem is maintaining acid contact with
the scale by preventing leak-off into the formation or through wormholes in the scale.
Weak hydrochloric acid (7 1/2 to 15%) is generally used for the removal of calcium carbonate scale,
rust (Fe2O3), millscale and corrosion products (FeS). At higher temperatures (>250°F), weak
organic or phosphoric acids are often used to minimise the risk of damage to downhole equipment.
Organic acids are also used where the objective is to remove scale from a pump, or where stainless
steels are used downhole (T >150°F). Only small volumes of acid are generally required for
dissolving carbonate scale (100gal 15% HCl per ft3 of scale). Treatment of scales with a solvent, as
well as acid, may be necessary where scale and wax or asphaltenes have been deposited together.
It is important to remember that many scales are not acid soluble and, therefore, it is essential that
samples be analysed before designing a treatment.

Page 25
Section 3 – Well Performance

3.6.2.3 Matrix Acidising Of Sandstones


Most books and papers on matrix acidizing in sandstones focus on the removal of clays and fines
from the immediate wellbore areas with mud acid (HCl/HF). As discussed in 3.5.7, this type of
damage is mostly caused by:
• Solids invading the pore structure due to poor quality wellbore fluids or filtercake.
• Destabilization of the native clays by the invading fluids and filtrate.
• Transportation of migrating clay particles into the wellbore area during clean-up or
water breakthrough.

The problem is most severe, and most difficult to treat, in low permeability (<100md), high clay
(>10%) content formations, like those found in many deltaic depositional environments (e.g. Gulf of
Mexico, South East Asia, etc).
When fine materials and clays are trapped in the pore throats, they usually have to be at least
partially dissolved with hydrofluoric acid (HF) in the form of mud acid (HCl/HF) or fluoboric acid
(HBF4). Regular mud acid has a concentration 12% HCl and 3% HF and is most successful in well
consolidated sands with high quartz-to-clay ratios.
However, poor acid response was noted in many shaly, weakly consolidated formations. Therefore,
in recent years, there has been a tendency to use lower strength mud acid (6.5/1.0) or fluoboric acid
(0.1 to 0.2% HF) in formations containing large amounts of shale. Straight HCl is also widely used
for formations containing large amounts of acid soluble material and in situations where the purpose
is to dissolve acid soluble lost circulation material.
An acid selection guide for sandstone acidizing, based on the permeability and clay content of the
formation has been compliled:

HCl Solubility ≥20% 15% HCl only

Removal of Acid Soluble LCM 15% HCl only

High Permeability (100md or more)


High quartz (80%), low clay (<5%) 12% HCl and 3% HF*
High feldspar (>20%) 13.5% HCl and 1.5% HF*
High clay (>10%) 6.5% HCl and 1% HF**
High iron chlorite clay 3% HCl and 0.5% HF**
Low Permeability (10md or less)
Low clay (<5%) 6% HCl and 1.5% HF†
High chlorite 3% HCl and 0.5% HF††

* Preflush with 15% HCl


**Preflush with sequestered 5% HCl
† Preflush with 7.5% HCl or 10% acetic
†† Preflush with 5% acetic

Page 26
Section 3 – Well Performance

A hydrochloric acid preflush of approximately the same volume is used to:


a) Displace or dilute any formation water containing potassium, sodium or calcium ions
that could cause fluosilicate deposition.
b) Maintain a low pH in the near wellbore area throughout the job and clean-up to
minimise colloidal silica, ferric hydroxide, gibbsite or fluosilicate deposition due to
overspending.
c) Dissolve calcium carbonate and avoid calcium fluoride (CaF2) deposition.
d) Ensure that the high cost HF is spent on dissolving clays and not wasted reacting with
materials that can be removed by HCl.

One of the main problems is to ensure that the preflush precedes the mud acid into all perforations.
This requires special attention to diversion techniques; either doing the job in multiple stages of
preflush, pretreatment, mud acid and a positive diverter; or by using the pretreatment to open up the
entire interval and to establish a more or less uniform permeability or injectivity profile with
selective matrix diversion. Matrix diversion is also needed within the acid to ensure that the high
permeability streaks, large pores, or those first treated, do not take all the acid. Simulation results
show that this is best achieved at low injection rates (0.01 b/min/ft).
For a typical job, an organic solvent spearhead is used to remove wax and oily deposits, followed by
a 50 to 100gal/ft preflush with 5 to 15% HCl. The mud acid volume should be 50 to 150gals/ft
pumped at low rates (0.5 b/min) to get a dynamic contact time of 0.5 to 4 hours. An overflush of
50gal/ft 5% HCl, ammonium chloride or diesel (or nitrogen in a gas well) is injected to push the
spending acids deep into the reservoir in order to minimize the damage caused by the settlement of
partially dissolved solids and by-products while the acid is stationary. The acid should be produced
back as soon as possible (unless fluoboric acid is being used to stabilise shales, when a shut-in
period is needed). Matrix treatments must be injected at non-fracturing rates with the bottom hole
pressure being maintained at least 10% below the fracture propagation pressure.
Acid does not necessarily increase the permeability of a formation, indeed in some cases it can
cause more damage than it was designed to remove. HCl can reduce permeabilities in ‘dirty’
sandstones by dispersing clays and fines which may subsequently block pore throats. Iron dissolved
by the acid can reprecipitate as the pH falls, reducing permeability. Acid can also cause the
precipitation of sludges on contact with crude oil. In addition, mud acid can reprecipitate silica as it
spends; and, if the preflush is inadequate, the mixing of hydrofluoric acid with formation water, or
its contact with carbonates, can result in the precipitation of insoluble fluorides. Tests required to
ensure that these problems are avoided are discussed in detail in the Near Wellbore manual. In some
cases, acidisation may simply be an inappropriate treatment for a particular formation.
Particular care needs to be taken with formations containing large amounts of ferric ions in the form
of pyrite or chlorite clay minerals. (The acid can also pick up ferric iron from the tanks and tubulars
if these are not properly cleaned prior to the acid job.) In these situations, a sequestering or chelating
agent is required to prevent the acid from overspending (i.e. pH >2.2), or to complex the iron into
solution. Commonlyused additives include:
• Acetic acid (T <150°F)
• Citric acid (T <250°F)
• EDTA or NTA (T <350°F)
• Erythoboric acid (T <400°F)

In sour wells, there is the additional problem of sulphur deposition by the reaction of the ferrous
ions with the H2S. In slightly sour wells (<5%), it is possible to control the amount of H2S picked up
by the acid with a hydrogen sulphide control agent.

Page 27
Section 3 – Well Performance

3.6.2.4 Matrix Acidising Of Carbonates


For acidising carbonates, 100 to 250gal/ft of HCl is normally used at concentrations of between 15
and 28% and can effectively treat some 12 to 18” around the wellbore at moderate temperatures
(<250°F). Injection should be as fast as possible at non-fracturing conditions and should increase
through the job.
Two processes are often involved in matrix acidising of carbonates:
• Enlarging natural fractures.
• Creating wormholes into the formation.

Both will result in highly conductive flow channels in the wellbore region.
Wormholes are randomly occurring, long (1 to 36”), empty channels with diameters ranging from
pore size to several millimetres. They are caused by a variety of instability phenomena:
• Large pores enlarge faster and, therefore, take more acid.
• Neighboring pores coalesce.
• Viscous fingering at high viscosity ratios and capilliary fingering at low viscosities
cause the acid front to be unstable.
• The length of the acid front between spent and unspent acid.

The number and size of wormholes cannot be predicted by either theory or experiment, although
some recent work using fractal theory appears promising and may help to define an optimum
injection rate.

Figure 3.9 – Wormhole Patterns From Acidising A Non-Fractured Carbonate With HCl

Page 28
Section 3 – Well Performance

The major concern is to avoid overacidising and destabilising the rock at the perforations by
maintaining a reasonable flowrate, while limiting the total number of wormholes from any given
perforation. This is achieved by initiating dissolution at relatively slow rates (<0.02bbl/min/ft) and
then progressively increasing the pump rate to the maximum non-fracturing rate to limit the size of
the wormholes in the near wellbore area.
If the primary objective is to stimulate the natural fractures and bypass damage, the acid is often
viscosified, emulsified, foamed or treated with a fluid loss additive to aid penetration. In low
permeabilities, however, care must be taken that a reasonable injectivity rate (0.25 to 0.5b/min) can
be maintained without fracturing.
Acid concentrations and volumes should be increased with the expected depth of damage,
temperature and permeability. Hydrochloric acid is most active at around 22% concentration, so
treatments commonly use 15 to 28% HCl.
At high temperatures (<250°F), it is very difficult and expensive to inhibit high strength HCl
effectively and a less corrosive organic acid is generally used (e.g. formic acid can be effectively
inhibited at temperatures up to 400°F). Alternatively, the well tubulars can be pre-cooled by
circulation or injection of a brine preflush. At very high temperatures (>450°F), it may be advisable
to try removing the damage with a non-acidic suspending solution.

3.6.3 Types Of Acids And Additives


The variety of acids commonly used and their additives for the various treatments can be found in
the Near Wellbore Manual.

3.6.4 Acid Placement Methods


This section describes the various methods of acid placement.

3.6.4.1 Bullheading
Pumping the acid down the production tubing or drillpipe is the simplest and cheapest placement
method. It permits high injection rates and flexibility in the selection of diverting agents.

Note: However, bullheading involves risk of formation damage and inefficient acid
usage, unless the pipe is clean and scale free.

Pumping acid through tubulars is a very effective way of removing solids from the wall, many of
which will not dissolve and may be deposited in the formation (e.g. wax, pipe dope, gypsum, barite,
etc). Other pipe wall deposits will cause the acid to partially spend (calcium carbonate, millscale,
rust, corrosion products) and may induce ferric hydroxide precipitation. Moreover, unless
elastomers and nipples have been specially selected for acid exposure, there is a risk of causing
completion failures or seal leakages at wellheads, downhole safety valves, sliding sleeves and seal
assemblies.
If the tubing is to be used when acidising an oil well, it is highly desirable to preclean it with both
an organic solvent and an acid wash. This can be done by:
• Pickling the tubing prior to installation.
• Circulating through a sleeve above the packer.
• Pumping down a small wash above a set of swab cups.

Page 29
Section 3 – Well Performance

A second problem with bullheading is that any standing rat hole fluids will be injected into the
formation. These are often extremely dirty and can cause severe plugging. An additional problem
with bullheading is achieving effective and rapid clean-up of acid from the wellbore area, unless the
well is capable of flow or is completed with gaslift valves.
Nevertheless, bullheading is commonly used for treating gas wells, new completions capable of
flow and water injection wells. In these latter cases, corrosion products are usually acid soluble, so
the acid volumes are simply increased to account for some wastage; and the products are carried
well away from the wellbore by recommencing injection immediately after the acid has been
displaced.

3.6.4.2 Coiled Tubing/Snubbing Workstring


Coiled tubing is now commonly snubbed into the well under pressure to place acid without
contamination. This tubing is of small diameter and is sufficiently flexible to be coiled and pushed
into the well, refer to section 14.4.3). It can be easily cleaned prior to the job; and can be used to
displace or blow out the rat hole fluids.
Placement can be enhanced using inflatable packers or jetting tools or by moving the coiled tubing
to locate the acid where it is most required. The coiled tubing can also be used for gas lifting the
spent acid from the well and for washing critical areas, such as seal assemblies and wellheads.
The use of coiled tubing restricts the rates at which the treatment can be pumped, encouraging
operators to maintain slow injection rates. However, the limited throughbore in the check valve may
restrict the type of diversion policy that can be adopted.
Conventional tubing or a macaroni string can also be snubbed into the well for this purpose (refer to
section 14.4.4). However, the presence of couplings makes this a more lengthy and potentially a less
safe operation.

Page 30
Section 3 – Well Performance

3.7 Vertical Lift Performance


The pressure drop in any tubing or conduit for any fluid is a function of three components:
• Gravity head
• Friction loss
• Acceleration

∆Ptotal = ∆PGravity + ∆Pfriction + ∆Pacceloration


The acceleration loss term is normally small and is ignored unless there are large amounts of
expansion (i.e. a well producing to atmospheric pressure).
The friction is dependent on the fluids (i.e. the PVT), the completion (ID and roughness) and of
course the flowrate. Frictional pressure drops may be typically 10% of the overall pressure drop for
a properly sized oil well completion, although for a gas well may be significantly higher. The
roughness of the tubing will depend on the condition and metallurgy.

Tubing Material Roughness Information Source


(in)
CeramKote54 0.000027 National Engineering Laboratory Report FRI001
Tuboscope TK236 0.00008 Tuboscope Coatings Manual
New 13%Cr L80 0.0006
New Carbon Steel L80 0.0029 Measured on BP stock
Lightly Corroded Steel 0.01
Heavily Corroded Scaled Steel 0.1 Estimate based on 2.5mm pitting or scale
Table 3.4 – Roughness Of Tubing

The gravity term is the weight of the fluid. For a single phase system, it is simply the density of the
fluid. For multiphase systems, the density is a function of the relative proportions of the different
phases. If the phases are traveling at the same speed (i.e. no slip), then the proportion of each phase
will depend on the PVT properties and the pressure. In reality the gas will travel faster than the
liquid and this will increase the proportion of liquid (termed the hold-up). Correlations are used to
determine the amount of slip and, therefore, the proportion of each phase. The correlations are
dependent on accurate PVT data in order to make valid predictions of the slippage.
The relative contribution of the friction and the density will vary with production rate, (refer to
Figure 3.10), and also tubing size. Note: The average density reduces as the flowrate increases
because the slippage will reduce until the hold-up approaches the no slip hold-up. Below a certain
rate the density increases markedly and will approach the liquid density at zero flowrate because the
slippage is excessive, and in fact, the liquid will not be produced at a fast enough rate to be
produced out of the well. When this happens the liquid builds up until it is produced out in a large
slug. This is called heading and is outside of the scope of the pressure drop correlations available for
steady state flow because it is a transient or time dependent phenomena.

Page 31
Section 3 – Well Performance

Figure 3.10 – Friction And Density Contributions To Pressure Drop

This transient phenomena has four implications:


• All the Prosper pressure drop correlations will be invalid in this flow area.
• Any well tests flowing at these rates cannot be used for comparison or tuning of the
flow correlations.
• The completion and well design should avoid flowrates in this area by selection of an
appropriate choice of tubing size (see section 3.7.9).
• Shut-in wellhead pressure cannot be predicted by using the total pressure drop at a zero
rate. This will seriously under predict the shut-in pressure.

As the gravity term is typically 70-90% of the total pressure drop inside a properly designed oil
completion, getting the density of the fluids correct is vital. The PVT data must, therefore, be at
least accurate when it comes to the oil density, formation volume factor and GOR.
There are a number of techniques used to calculate the hold-up. The simplest (e.g. Gilbert’s curves)
are empirical correlations based on field data.
Of greater complexity are correlations that attempt to predict a flow regime and then use
correlations.

3.7.1 Correlation Selection


The vertical lift performance correlations available are extensive. However, each correlation has
been developed to satisfy the prediction of certain systems. No single correlation is available that is
ideal for all wells. It is, therefore, very beneficial if the correlations can be short-listed based on
their intended use (i.e. flow regime or oil density for example). From the short-list it is then possible
to select a correlation that most accurately matches real data, e.g. a well test.

Page 32
Section 3 – Well Performance

3.7.1.1 Fancher-Brown Correlation


This is a strange correlation in that its purpose was not intended to accurately predict the tubing
performance but is a non-slip (or homogeneous) flow equation. This means that it will always over-
predict the fraction of the pipe occupied by the gas. As a result it tends to under-predict the pressure
drop where slippage occurs. It will, therefore, only accurately predict the pressure drop under high
flow bubble regime conditions or single phase flow (e.g. above the bubble point).
The importance of this correlation is that it should under-predict pressure drop, therefore, when
plotted against real data it should lie to the left of the real (measured) data. If it lies to the right it
indicates there are errors, either, in: the measured data, friction, or more likely, density predictions
(i.e. PVT problems).

3.7.1.2 Hagedorn And Brown


This relatively simple flow correlation does not use a flow map. Instead it is based on data from a
1,400ft test well with air, water and dead crude over a range of viscosities from 10-110cp. It covers
bubble and slug flow and therefore should not be used for gas wells or high GOR wells. The
original Hagedorn and Brown correlation should never be used, however refinements in Prosper
have avoided some of the original problems. They also allow this correlation to be used for deviated
wells. This correlation should not be used to predict the minima of the lift curve. It tends to under-
predict the minima. There are also errors in large bore wells with a 35-70° deviation where oil/water
slippage may occur. Despite these constraints it is frequently used and accurate under the conditions
for which it was designed.

3.7.1.3 Duns And Ros


This correlation uses a flow map (i.e. determines the flow regime) to determine the hold-up and
friction. This correlation should, therefore, apply to all flow regimes. In reality, it works best in gas
and gas condensate wells, especially in mist flow and performs poorly in oil wells where slug flow
occurs or at low rates. The loading rate prediction for gassy wells (>6,000scf/stb) is generally good.

3.7.1.4 Orkiszewski
This correlation is an amalgam of various correlations including Duns and Ros. It works well over a
wide range of conditions and often appears as the closest correlation to match well test results.
Unfortunately, there is a discontinuity in the predictions at a mixture velocity of 10ft/sec. This
results in a jump in predicted pressures (as seen in Figure 3.11). This means that the correlation is of
little practical use.

Page 33
Section 3 – Well Performance

Figure 3.11 – Orkiszweski VLP Predictions

3.7.1.5 Beggs And Brill


This was the first comprehensive study of multiphase flow for inclined pipes. Small diameter pipe
(up to 1.5”) was used in experiments where the flow regime was observed. The correlation over-
predicts liquid hold-up in wells and pipelines, however, it is in widespread use as it applies to
vertical, horizontal and inclined pipes.
The main use of the Beggs and Brill correlation is for quality control. As it over-predicts the hold-
up, it over estimates pressure losses typically by between 10 and 15%. This means that actual data
should lie to left of the Beggs and Brill correlation. If results lie to the right of the correlation,
possible causes are:
• Inaccurate PVT data - particularly too low a liquid density
• Under-recorded water cuts or over estimated gas rates
• Too low a friction factor

Caution: Arbitrarily altering any parameter in order to get a match must be avoided.
There may be a temptation to increase roughness in order to obtain
consistency. However, this should only be done if there is physical, or
analytical, evidence that corrosion, scale or other deposits may be affecting
the condition of the bore of the production conduit.

Page 34
Section 3 – Well Performance

3.7.1.6 Ansari
This correlation is a ‘mechanistic’ model as it predicts pressure drops based on the mechanism, i.e.
the flow regime. This is similar to the more sophisticated traditional methods such as Duns and Ros.
However, newer mechanistic models take this even further by eliminating the more empirical steps
and considering the transitional regimes further.
The Ansari correlation should, therefore, cover all flow regimes, all sizes of tubing, all deviations
and everything from heavy oil to gas. In a comparison of this model across a number of oil and gas
fields, at a variety of conditions, this correlation was the most widely applicable although by no
means perfect, with average absolute errors of around 7%. There is still concern that this correlation
(and many others) is not fully applicable for deviated wells.
This correlation finds greatest application with low GOR, heavy or medium oil fluids such as
Harding, Foinaven and Andrew. It is acceptable for use on higher GOR fluids and gasses, but
correlations such as Gray are often better for gas.

3.7.1.7 Petroleum Experts Correlations


Petroleum Experts have developed their own hybrid correlations. According to their information,
‘Petroleum Expert’ uses the Gould et al Flow Map and for the various flow regimes the following
correlations are applied:
• Bubble flow: Wallis and Griffith
• Slug flow: Hagedorn and Brown
• Annular Mist flow: Duns and Ros
• Transition: combination of slug and mist

‘Petroleum Experts 2’ includes the features of the PE correlation plus original work on predicting
low-rate vertical lift performances and well stability.

3.7.1.8 Gray
This correlation was developed for vertical gas condensate wells and was originally used in the API-
14B subsurface safety valve sizing program. The published limits of the correlation are a flow
velocity of 50ft/sec, 3 1/2" tubing or less, and a condensate ratio of under 50bbl/mmscf. In reality
however, Gray has proven accuracy in low to moderate CGR gas wells and is useful for predicting
liquid loading. It is empirical and uses its own PVT module for condensate and liquid prediction. It
is, therefore, not compatible with compositional PVT models.

3.7.2 Use Of Correlations For Well Dying Prediction


The correlations also vary considerably in their ability to accurately predict the minimum of the lift
curve. Figure 3.12 shows a comparison of the correlations used in Prosper. Note the difference in
the shape of the curves at low flowrates. The minima predicted varies between about 2,000bpd and
8,000bpd.

Page 35
Section 3 – Well Performance

Figure 3.12 – Correlation Comparison

In general the following caveats apply:


• Do not use the correlation outside of its intended regime i.e. don’t use Duns and Ros for
oil.
• Avoid using Orkiszewski, Fancher Brown, Beggs and Brill or Hagedorn Brown for
predicting tubing lift minima.
• In the absence of a verified correlation for the field, Ansari has proven good at
predicting minima over a wide range of fluids.

3.7.3 Tuning Pressure Drop Correlations


Prosper has the capability to ‘tune’ the correlations to match real data. This achieves this by
regression on two variables:
Parameter 1 - this is a multiplier on the slippage and therefore alters the hold-up and, therefore,
the gravity term. For a discrepancy more than 5%, the density term itself is altered.
Parameter 2 - this is a multiplier for the friction term.

Page 36
Section 3 – Well Performance

Extreme caution must be exercised if tuning the VLP correlations. It is possible to get a reasonable
match with any data, simply by tuning the correlation. Tuning the VLP correlation is only valid
under the following conditions:
a) There is absolute confidence in the PVT data and that it is valid for the conditions the
well was flowing under when it was tested. For example if the PVT data is for a certain
GOR and the well test GOR has risen, the confidence in the extrapolated PVT data will
be reduced.
b) The flowrates (oil, water and gas), BHP, BHT, THT and THP are accurately known
during the tests and are steady. In order to confirm the accuracy of these measurements,
the type of meters, the flow period and whether the meters have been recently calibrated
should be critically analysed and the pressure data not confined to a spot reading. The
pressure should be watched for at least 10 minutes to ensure that it is not fluctuating and
effected by instability or heading.
c) The flow regime should not be severe slug or any other unstable state. This can be
checked by plotting the predicted BHP vs. rate. If the test point lies close to or to the left
of the minima of the curve, the correlations will not accurately match the test data and,
therefore tuning the correlations will probably render invalid any predictions for
conditions outside of the test condition.

3.7.4 Natural Flow


Prosper can easily be used to predict the natural flow performance of a well over the well’s lifetime
using the system performance calculation.
Issues to be careful of are:
• Ensure that the datum depth for the completion is the same as that used to reference the
reservoir pressure. It is accurate to use the mid depth of the completed interval.
However, unless friction in the liner is large, then it is not critical where the datum is, so
long as it is consistent.
• It is not normally necessary to include restrictions in modelling the completion. Nipples,
packers, gas lift mandrels and tubing retrievable safety valves will cause a negligible
additional pressure drop. Wireline retrievable valves should be included and any change
in tubing size.

3.7.5 Shut-In Conditions


Determining the wellhead shut-in pressure is an important part of well performance prediction but it
is not easy. The wellhead shut-in pressure is required for:
• Determining the flow control requirements at surface i.e. the wellhead and tree ratings.
• Required treating pressures, e.g. scale inhibitor squeezes, fracture treatments, etc.

A number of techniques are available:


1 Assume a gas gradient to surface and the maximum reservoir pressure. This will give
you the worst possible case under all conditions (so long as the lowest gas gradient and
the highest possible reservoir pressure are used). Note: Remember that the gas gradient
has a very strong (almost linear) relationship with pressure. The easiest method of
determining the gas gradient is to use Prosper and perform a gradient prediction with a
gas fluid (no condensate). A low rate (e.g. 0.1mmscf/d) rate can be used. The gravity
pressure loss can then be extracted from the variables available and plotted (refer to
Figure 3.13).

Page 37
Section 3 – Well Performance

Figure 3.13 – Gas Pressure Gradient Calculations Using Prosper

Clearly this will frequently tend to give an over-estimation of the wellhead pressure,
particularly if the fluids are not gassy or a lot of water is being produced. What
frequently happens is that the high gas oil ratios are associated with depletion and,
hence, high gas oil ratios (and, therefore, a simplification of gas only) do not combine
with high reservoir pressures to give high wellhead pressures. However as a first pass it
is quick, simple and conservative.
2. The next approach is a bit more subtle. If the BHP vs Rate could be extrapolated down
to a zero rate without any of the problems associated with varying hold-up at low rates,
it would be close to extracting a realistic pressure gradient. This is easily achieved using
the Fancher - Brown no-slip correlation. This correlation will give the highest
proportion of in-situ gas and, therefore, will be conservative. This gradient will be
realistic for conditions immediately after a shut-in. Unfortunately phase segregation will
occur which can result in higher pressures than immediately after. Fortunately the oil
will always drop in comparison to the gas. This means that there will be no further
phase transition from oil to gas. Therefore, as the standard volume of gas is known at
the time of shut-in, it is conservative to assume that the amount of space occupied by the
gas remains the same. This then means that the standard gas volume is (1-total hold-up)
x completion volume. The actual gas volume is this standard gas volume corrected to
the pressure in the well.
3. Use a transient model - if one can be found ?

Page 38
Section 3 – Well Performance

3.7.6 Artificial Lift


The well performance aspects of artificial lift need to be integrated with all the other aspects of
artificial lift completion design. These are covered in section 4.
Prosper can be used to assess the options for Gas Lift or Electrical Submersible Pumps or to
troubleshoot the performance of such systems when installed.

3.7.6.1 Gas Lift


Gas lift works on the principal of lowering the hydrostatic pressure by lowering the average density
of the produced fluid by introducing gas. This is easily modelled in most well performance
predictions so long as varying gas/oil ratios can be modelled accurately. This automatically excludes
the effective use of EoS models, where a different model is required for each composition.

Injection Depth
Because gas will occupy a greater volume at lower pressures, injecting a small amount of gas will
have a relatively large effect at shallow depths. Injecting gas deeper will always be beneficial.
Because the deeper gas will be at a higher pressure, more gas may, however, be required, or be
beneficial. This is demonstrated in Figure 3.14.

Figure 3.14 – Benefit Of Increasing Gas Lift Injection Point Depth

When considering what is the optimum gas lift depth, then the other completion design aspects must
be considered. These include casing design and packer setting depth which are discussed in some
detail in section 4.5.2.

Page 39
Section 3 – Well Performance

The benefit of increasing depth of injection will depend on the pressure drop correlation in use. This
is particularly the case for high angle wells. There is some concern that the current models may not
be entirely realistic when it comes to multiphase modelling at high angles. The benefit of injecting
gas very deep along highly inclined wellbores may be over-estimated by the models. This could
result in gas lift mandrels being placed where they might only be able to be serviced by tractors or
coiled tubing, without sufficient justification. Figure 3.15 shows a plot of injection depth vs. liquid
rate for a typical deviated well profile.
The big difference in the benefit of deep gas lift for correlations such as Hagedorn and Brown
compared to the Ansari correlation should be noted. In this instance it is suspected that the Ansari
correlation is closer to reality, due to its ability to handle a greater range of flow regimes.

Figure 3.15 - Gas Lift Injection Depth vs Rate For Five Untuned Correlations

Gas Injection Rate


Increasing the amount of gas injected will increase the liquid production levels up to a point. The
hydrostatic pressure will lower with more gas, but the frictional pressure will rise. There comes a
point when increasing the gas results in the frictional pressure increasing more than the hydrostatic
pressure decrease. This point is very condition and well specific. The maximum liquid production
rate is not going to be the optimum. When designing the field facilities, there will be a cost
associated with building gas compression.

Page 40
Section 3 – Well Performance

There will also be an operating cost associated with gas compression. This will include servicing
and power costs. This means that the optimum production rate will be somewhat below the
maximum. This can be determined by evaluating the cost vs. benefit of the gas and plotting this line
on the performance curve, (refer to Figure 3.16). When operating these wells, the availability of the
gas must also be considered. This is covered in section 4.5.2.

Figure 3.16 – Injection Rate vs Production Rate

Gas Lift Design


Having determined the optimum gas injection depth and rate, well performance predictions can be
used to determine the optimum gas lift design. Note: At this point it is very important to consider a
life of well approach, rather than a specific condition. Prosper can be used to determine the gas lift
design with respect to well performance. However, it is strongly recommended that the gas lift
company or gas lift expertise is included at an early stage so that the whole system (wells, facilities
and reservoir) can be integrated and best practice incorporated.
The purposes of the gas lift design are:
a) Produce a gas lift design that achieves the optimum production rate
b) Produce a design that is reliable and does not introduce heading problems, instability or
start-up problems.
c) Produce a design that is simple to troubleshoot and optimise
d) Minimise costs by avoiding unnecessary gas lift mandrels.

Prosper will assist in the first of these issues.

Page 41
Section 3 – Well Performance

The design capabilities of Prosper are quite extensive and will not be described in detail here.
However, some points that may be useful are listed below:
a) There is a choice of valve type available. All gas lift valves are sensitive to both tubing
and casing pressure, however, casing pressure operated valves are more sensitive to
casing pressure and vice versa. The selection should not be dictated by performance and
uncertainty. A casing sensitive valve requires a drop in casing pressure to close the
valve, therefore potentially more valves may be required or a higher unloading pressure.
Figure 3.17 shows a plot of a gas lift design using a casing pressure drop to close the
unloading valves. However, it is less sensitive to changes in reservoir properties as the
casing pressure is much easier to determine than the tubing pressure. Tubing pressure
operated valve systems are conversely harder to design as they are more dependent on
changes in reservoir properties.
b) The number of valves required is dependent on the separator or wellhead pressure
during the unloading and the assumption about the static gradient of the load fluid. The
gradient may be the completion fluid gradient, however this can produce a design that is
overkill for all other conditions. The possibility should be given to designing the gas lift
system so that the unloading (i.e. removing the completion fluid from the tubing) can be
performed by a nitrogen lift and routine kick-off performed as normal. This means that
the static gradient would become the worst case static gradient of the reservoir fluid.
This can be assumed to be 100% formation water, or whatever.
c) Consider the point at which gas lift is required. During the early field life, there may be
little or substantial benefit in gas lift depending on the conditions and, especially, the
GOR of the produced fluid.
d) Do not design the system just for high rates and high reservoir pressures. This will push
the use of large orifice sizes which may introduce heading problems later on in the well
life.
e) Depending on the access difficulties, it may be useful to consider changing the orifice
valve position with time. As the reservoir depletes, the use of a deeper injection position
may be possible and beneficial. This implies that, either, two orifice valves are installed
from the first day or the lower mandrel is installed with a dummy valve.
f) The pressure drops with gas between the casing and the tubing are usually not
significant. However, Prosper has the capability to include these as required.

Page 42
Section 3 – Well Performance

Figure 3.17 – Gas Lift Design Plot

3.7.6.2 Electrical Submersible Pumps


A pump can be introduced into the completion in order to increase the energy of the system. The
pumps are centrifugal multistage pumps powered by a three phase motor. The speed of the motor is
dependent on the frequency of the power supply. The well performance aspects of such pumps and
motors are addressed in this section, however it is important to integrate the facilities, completion
design and reservoir issues. These are considered further in section 4.6.3.
Prosper works in a similar manner to the hand calculations that can be performed for ESP design:
1. First the total dynamic head requirement is calculated. This is head required to lift
the fluids from the suction to the discharge of the ESP. The head is required rather
than the pressure as the pump will deliver a constant head independent of the fluid,
whereas the pressure will be dependent on the fluid being pumped.
2. A pump stage is then selected in order to fit inside the casing and deliver the
flowrate. The number of stages is then chosen in order to deliver the total dynamic
head. With pump efficiency this determines the motor power requirements.
3. A motor and cable are then chosen to deliver the pump power requirements. With
motor efficiency and cable losses, this determines the surface power requirements.
4. Sensitivities are performed to see over what range of conditions the pump and
motor can operate. If necessary a variable speed drive can be used in order to extend
the range of conditions the pump and motor can cope with.

Page 43
Section 3 – Well Performance

The following issues must be considered during this process:


a) The position of the pump must be considered and its implications - see section 4.6.3.
b) The power requirements will depend largely on the fluids being pumped and the
reservoir pressure. This must be assessed over time, so that facilities are not under
designed. In one case, the reservoir depletion was not accounted for in the facilities
design and as a result the power available was only 1/4 of what it should have been.
c) The design of the pump/motor should be based on expected changes and uncertainty in
conditions - but over a much shorter time frame than the facilities design. The average
run life of an ESP is 2-4 years. There is therefore no point in designing it for the
reservoir pressure expected in 10 years time!
d) Free gas at the pump suction may seriously limit the performance of the pump. This
will, either, limit the available drawdown or require gas handling pumps or a gas
separator. Prosper can output the free gas liquid ratio at the pump intake.
e) The fluids will have an effect on the pump efficiency. With emulsions, the pump
efficiency may only be 10%, compared to a typical figure of 40% for less viscous
fluids8. This poor efficiency is however not all bad news. The other 90% will be
transferred to heat energy in the fluid and may increase free gas above the pump, lower
viscosities and aid in oil/water separation at surface. Prosper can not directly perform
these calculations, however the energy ‘lost’ will be known and can easily be converted
into heat energy of the fluid by using the heat capacity of the fluid. This temperature
increase can then be manually entered into Prosper.
f) The range of pumps and motors in the Prosper database is not complete and may not be
up-to-date. All designs should be checked by an ESP vendor(s) independently to see if
the same recommendations are made.

3.7.6.3 Jet Pumps


Jet pumps cannot be modelled in Prosper, however, they are an effective means of artificial lift and
should not be ignored. They operate on the venturi principle (refer to Figure 3.18). A power fluid is
pressurised and exits the nozzle at high velocity. The velocity of this fluid creates a low pressure
area in the throat. This low pressure draws in the reservoir fluid and the fluid exits the throat into
the diffuser, where expansion converts velocity into pressure.

Figure 3.18 – Typical Jet Pump Performance

Page 44
Section 3 – Well Performance

The performance of the jet pump is dependent on:


• Density of the power fluid
• Nozzle area (An)
• Throat area (At)
• Power fluid rate and pressure
• Head requirements i.e. reservoir performance, tubing performance, wellhead pressure
and power fluid rate

Note: As the power fluid is always mixed with the reservoir fluids it will have to be
produced and this will effect the tubing performance, particularly if a heavy
power fluid such as water is used.

BP have access to a jet pump simulator program called ‘Jet Pump Simulator’. It is a simple model
using a few vertical lift performance options, a few untuned PVT correlations, a simple string and a
straight line PI. It is however useful for artificial lift screening and jet pump selection. Various
sensitivities can be performed using this model, and the two jet pump parameters (throat and nozzle
areas) can be selected and input as well as. The program will then calculate the operating point
(flowrate) and pressures with these details. In the example below (Table 3.5), the power fluid is
water and the sensitivities have been performed on the nozzle and throat combinations.

Nozzle Throat Formation Power Pump Suction FBHP Pump Power Efficiency
Number Identifier Rate (stb/d) Fluid Rate Pressure (psia) Fluid Inlet Fraction
(stb/d) (pisa) Pressue (pisa)
3 0 1750 3939 3922 4297 10098 0.222
3 1 2441 3989 3763 4138 10098 0.308
3 2 2926 4024 3652 4027 10098 0.375
3 3 2991 4029 3637 4012 10098 0.385
3 4 2354 3983 3783 4158 10098 0.279
3 0 2169 5126 3826 4201 10095 0.240
3 1 2925 5196 3652 4027 10095 0.331
3 2 3328 5233 3559 3934 10095 0.384
3 3 3131 5215 3605 3980 10095 0.357
3 4 1940 5104 3878 4254 10095 0.215
Table 3.5 – Example Output From Jet Pump Simulator

The throat and nozzle identifiers are the manufacturer’s classification of the throat and nozzle areas.
For example a size 3 Nozzle has an area of 0.004in2. Note: there are differences between the
manufacturers in their throat/nozzle identifiers and the corresponding sizes.
Contact the well design team in SPR for further details of how to use and access this program.

Page 45
Section 3 – Well Performance

3.7.7 Heat Transfer


The accurate modelling of heat transfer is important for a number of reasons:
• The temperature will effect the well performance as it will determine the viscosity of
fluids and the amount of gas held in solution.
• Predictions of the maximum wellhead temperature so that this can be used for subsea or
facilities input (e.g. materials, pipeline performance or production chemistry impact).
• Production chemistry impact in the completion, for example; wax, hydrate and
sometimes scale formation will be dependent on the temperature of the fluids being
produced.
• Materials selection, certain materials such as elastomers will have temperature limits. If
these materials are used away from the reservoir section it may be unduly pessimistic to
assume that the temperature is still that of the reservoir.
• For the modelling of tubing and casing stresses.

For these reasons, it is worthwhile spending some effort in determining the heat transfer. Heat
transfer will comprise:
• The transport of fluids by production or injection
• Conduction between the fluids, the tubing, the casing, cement and the rock
• Convection of fluids in the annuli
• Phase to phase transfers - e.g. oil to gas
• Expansion or contraction of fluids (especially gases)

There are three main choices for heat transfer modelling:


a) Assume a fixed temperature profile, e.g. as extracted from well test data. This will often
be sufficient for determining the pressure drops in the completion so long the flowrates
are similar to the well test data.
b) Assume a constant heat transfer coefficient. Prosper has the capability to model heat
transfer with one overall heat transfer coefficient. Typical values are 4-6 BTU/h/ft2/F.
The heat capacity of the fluids is also required. The heat capacity of the oil will vary
considerably (up to a factor of 2) between a light and a heavy oil. This is probably the
best method to use where there is some temperature data from well tests. The well test
configuration can be modelled and a trial and error approach used to match the heat
transfer coefficient with the measured temperature data. Care should be taken with
offshore wells however as an overall heat transfer coefficient will likely over simplify
the model.
c) Model the well in detail. Either Prosper or Welltemp (part of the Wellcat) package can
be used for detailed temperature modelling. Considerable detail is required including all
casing strings, annular contents, and rock types and properties. Unless all of this
information is rigorously included, the accuracy will not be any higher than option c
above. This option is the only method that will allow an estimate of the transient heat
transfer effects rather than steady state.

Page 46
Section 3 – Well Performance

3.7.8 Erosion
Corrosion and erosion are fundamentally related to the flowrate and flow type, therefore, Prosper is
the ideal tool in which to assess the chances of erosion or erosion-corrosion. There are no exact or
theoretical relationships between the flowrate and the resulting erosion. However, BP have further
developed the empirical and conservative API erosion limit with corrections for flow type,
metallurgy, corrosion and sand.
For multiphase flow, a C factor is used, where:
C
VE = Equation 8
ρm

Where:
ρm = mixture density in lbs/ft3
VE = erosional velocity limit (ft/s) and C depends on the metallurgy and flow type.
The erosional velocity (ft/s) should be compared against the mixture velocity (Vm), where:
qL + qg
Vm = = vsl + v sg Equation 9
Ap

Where:
QL and qg = volumetric flowrates of liquid and gas (in ft3/sec)
Ap = cross sectional area (ft2)
vsl and vsg = superficial liquid and gas velocities

The superficial liquid and gas velocities and the mixture density can be extracted from Prosper by
generating the full output from the gradient calculations. Using Excel, this can be converted into a C
factor vs. depth. This can then be compared against the C factors suggested in Table 3.6.

Condition Gas Liquid Multiphase


C = 250 (CS) C = 135 (CS)
V = 70m/s (if dry and
Normally sand free C = 300 (13Cr) C = 300 (13Cr)
completely free of solids)
C = 350 (DSS) C = 350 (DSS)
C = 150 (CS) C = 100
Sand Containing Contact SPR for more details C = 200 (13Cr) C = 100
C = 200 (DSS) C = 100
Table 3.6 – Erosional Velocity Limits

Reference: Erosion Guidelines Revision 3


Notes to table:
a) Sand free is defined as less than one part per thousand barrels (pptb).
b) The present velocity limits for ‘sand containing’ conditions should only be applied up to
275pptb. Above this level the appropriate specialist should be consulted.
c) C/S - carbon or low-alloy steels, 13Cr - 13%Cr steels, DSS - duplex stainless steels.

An example of extracting the C factor from Prosper is shown in Figure 3.19.

Page 47
Section 3 – Well Performance

Figure 3.19 – C Factor vs Depth Extracted From Prosper For A Variety Of Conditions

Page 48
Section 3 – Well Performance

3.7.9 Tubing Selection


3.7.9.1 Natural Flow Production
There are various rules of thumb for selecting tubing sizes. As a first pass, Table 3.7 can be used:
Tubing Size Liquid Flowrate Range Gas Flowrate Range
2 3/8” Less than 1mbp/d Less than 2mmscf/d
2 7/8” 1-3mb/d 2-8mmscf/d
3 ½” 2-6mb/d 3-14mmscf/d
4 ½” 3-10mb/d 6-30mmscf/d
5 ½” 5-17mb/d 8-50mmscf/d
7” 7-25mb/d 12-150mmscf/d
9 5/8” 15+mb/d 20+mmscf/d
Table 3.7 – Flowrate Range For Common Sized Tubulars

Note: These ranges are very generic and especially in gas wells, the choice of tubing
size is very dependent on the wellhead pressure.

A more rigorous method is to look at the minima in the liquid rate vs. pressure plot for the
minimum flowrate and the system performance plot for the maximum. As discussed in section 3.6,
this predicts a change a flow regime. As a rule of thumb, avoid conditions where the flowrate is
below this minima. However, this does not mean the well will not flow below this point. As shown
in Figure 3.21, a high PI well will likely not flow below the minima. This is because the two curves
do not intersect. A high pressure/low PI reservoir on the other hand, will be able to flow because it
only intersects the tubing curve at one point.
However the flow is likely to be slug or severe slug. A moderate PI and pressure well actually
intersects the curve at two points. The left hand point is unstable in that any perturbation will
feedback and push the flowrate, either, higher or lower. Thus starting the well off may be difficult.
Once started, flow should be possible, but again slugging is possible. As the two curves intersect at
a low angle, this slugging could cause severe instability.

Figure 3.20 – Using Tubing Performance Curve Minima

Page 49
Section 3 – Well Performance

Both the minimum and maximum rates of a range of tubing sizes can be plotted on one graph, e.g.,
Figure 3.21.

Figure 3.21 – Example Minimum And Maximum Flowrates For Various Tubing Sizes

In this example, the reservoir depletion is indicated by the changing inflow performance curves. The
minima of the tubing curves is also very clear, with 7" having a minima at 9,000bpd, whereas 3 1/2"
tubing has a minima at around 2,000bpd. The maximum rates for initial reservoir pressures are
25,000bpd for 7" and 11,000bpd for 3 1/2". However the constraints of the erosional velocity must
also be superimposed on this model.
With initial reservoir pressure, all the tubing sizes are stable, likewise after five years. However,
after between five and ten years, the wells will cease to flow on their own. In this instance there is
not much difference in when any of the tubing sizes shown become unstable. This is because the
productivities are high. With lower productivities, or for example gas break out in the reservoir, the
benefit of a smaller tubing size may be more apparent.
For gas wells, the same approach can be used, however great care must be taken with selecting the
PVT representation of the gas. The PVT section 3.4 discusses the pitfalls in the various black-oil
correlations and their assumption that the gas density can be increased to represent the condensate.
The most important point about using this method is that care must be taken with the flow
correlation. The correlations section 3.7.1 indicates which correlations can be used to accurately
predict the instability region.
The next improvement in this approach that could be made is to examine the flow behavior to the
left of the minima to see if it is operationally acceptable to facilities, discussed below.
One aspect of well performance in producers that is often overlooked is the location of the bubble
point and the impact that this has on tubing size. There will be no minima in the lift curve if the
fluid is single phase (i.e. all liquid or all gas). In these circumstances, the bigger the tubing the better
(from a well performance stand point).

Page 50
Section 3 – Well Performance

For example, in one field as the bubble point was only 90psi, the fluid was single phase all the way
to surface. The development plan proposed had a 7" liner with 4 1/2" tubing as this was appropriate
to the production rates expected. In this case, 7" tubing would have reduced pressure drops slightly,
but would have produced a monobore completion, with the downside being an increased cost of the
7" tubing and tree.
The same approach can be used for liner sizing. If the fluids will be single phase in the liner and a
long liner is to be used, then from a well performance stand point, a large liner (e.g. 7") can be
accepted. However, this is not the only consideration for liner sizing and a full discussion in
contained in section 2.

3.7.9.2 Tapered Strings


Completions with tapered strings can be very useful. As a well has a lower pressure at surface,
evolved gas increases and expands as it rises higher up the tubing, therefore frictional pressure
drops will correspondently be highest close to surface. At this point too, velocities will also be
higher. Larger tubing close to surface has the greatest effect in reducing both these pressure drops
and velocities.
The depth of a tubing size change can be determined. Figure 3.22 shows a plot of tubing
performance for a 1,000ft section of tubing at different pressures flowing with a moderate GOR.
The top curve is at 4,000psi, the bottom curve at 400psi and the other curves every 400psi. The two
curves at each pressure are for 5 1/2" and 7" tubing. Note: as the pressure increases, the minima of
the curve moves to the right. At lower pressure, the pressure drops are also highest at the higher
rate. These two phenomena indicate that there is a benefit of using the larger tubing size close to
surface. In this example, if the tubing was 7" for pressures less than 800psi and 5 1/2" for pressures
greater than this, the completion would give stable flow down to around 5,000bpd. This is the same
as if the tubing was 5 1/2" throughout. However the pressure drops have been minimised.

Figure 3.22 – Tubing Performance Curves At Different Pressures

3.7.9.3 Pumped Production


Page 51
Section 3 – Well Performance

The effect of tubing size on pumped production wells is much less than in the case of natural flow
wells or gas lifted wells. Generally, so long as there is no instability or slugging below the pump,
and the pump can deal with a large head of liquid above the pump, there should be minimal
problems in using a larger tubing size than for the case of a naturally flowing well. Texaco have
reported problems in pumps placed above long high angle wellbores where slugs of water have built
up in an undulating well profile. The danger here is that if there was free gas, a slug of gas would
cause gas lock or severe pump wear.
If severe slugging occurs above the pump, there is the potential for the wellbore to fill with
essentially pure liquid. Centrifugal pump stages are designed to be able to handle such a condition,
however this will increase the power requirement of the motor and must be addressed in the ESP or
hydraulic submersible pump design. The slugs are then a problem for the production facilities rather
than a potential cause of the well dying. If the slugging results in the loads on the pump varying,
then in a severe case, this could lead to the pump being periodically operated outside of the
upthrust/downthrust envelope and increased pump wear occurring. In the case of positive
displacement pumps, the slugging may place additional loads on the pump and rods and, similarily,
the pump unit should be designed to be able to handle a 100% liquid (water) head.
This phenomena means that much larger tubing can be effectively used. This is of significant value
in the case of electrical or hydraulic submersible pumps deployed inside the tubing by means of
coiled tubing or cable. The use of 7", 7 5/8" or even 9 5/8" tubing then becomes possible.

3.7.9.4 Gas Lift


For gas lift conditions, the choice of tubing size is again a balance:
• Too small a tubing size and adding gas may increase friction more than it decreases the
hydrostatic pressure. The result could be a well that flows better without gas lift, or
where the production rate is severely constrained.
• Too large a tubing size and the well may suffer from heading problems or instability.

The first issue is easy to quantify using Prosper. The second issue can be addressed using Dynalift.
Dynalift is a transient well performance model. When set up with details of the well, the inflow
performance, the gas lift valve etc. it can calculate likely performance in terms of slugging and
heading. Dynalift is also available within BP.

Page 52
Section 3 – Well Performance

3.8 Well And Field Optimisation


Well optimisation does not solely apply to gas lifted or pumped wells. Wherever there is a control
that will affect the well rates, or a resource (e.g. power), that control should be optimised. For
example a single naturally flowing well producing to a fixed pressure facility is controlled and
optimised by the choke, a gas lifted well is optimised by the gas lift injection rate, an ESP by the
electrical frequency and the choke and a jet pump by the power fluid injection rate. Note: in each of
these examples, the control can be changed at will from surface.
There is also another form of optimisation that can be applied by well interventions or workover.
For example, the tubing size can be changed in a tophole workover, the pump changed out or the
gas lift valve operating depth altered.
On a single well basis, optimisation is fairly straight forward. The optimum rate is that where the
production cash flow is maximised. Note that this doesn’t necessarily mean the highest production
rate as the expenditure used to get the fluids from the reservoir to the point of sale must be
accounted for. For example, the rate can be increased by fully opening the choke and lowering the
separator pressure, however this may introduce additional pumping costs or result in additional
flaring. The same applies for gas lifted or pumped production where the gas compression or power
requirements must be addressed. Figure 3.17 shows an example of optimising a single gas lifted
well.
Where more than one well is involved, the process becomes a lot more complex. This time the field
cash flow must be maximised. This must account for:
a) Value of the oil production
b) Cost of water production
c) Effect of increased production on pipeline/flowline pressure drops
d) Cost of power or gas compression
e) Facilities limits (oil, water, gas, power)
f) Downhole production constraints (asphaltene, scale, multiphase in the reservoir etc.,
coning etc.)
g) Facilities performance - e.g. what is the consequence if the separator pressures are
changed

In Figure 3.23 and Figure 3.24, an example optimisation is shown for four gas lifted wells with
varying water cuts. The first plot shows the performance curves for the four wells (note: two of the
wells cannot flow without the assistance of gas lift). In order to optimise the allocation of a limited
amount of gas to these wells, care must be taken. Figure 3.24 shows the actual optimised system but
that it is not simply a case of adding gas to the well where there is the most benefit. Well D actually
benefits the most (in terms of liquid rate) from gas lift due to its high water cut, but it is best to
leave this well shut in until 12mmscf/d of lift gas is available. At this point, some of the other wells
have to be slightly robbed of gas in order to kick off this well.

Page 53
Section 3 – Well Performance

Figure 3.23 – Example Gas Lift Performance Curves Of Four Wells

Figure 3.24 –Gas Allocation For Example Four Wells

3.8.1 Practical/Theoretical Optimisation


Page 54
Section 3 – Well Performance

There are three means of optimising a field performance:

3.8.1.1 Practical Methods


This relies on well tests in order to measure the effect of changing control values on the well or
facilities. For example, in order to assess the effect a change in separator pressure has on a well, a
multi-rate well test can be performed and a plot of wellhead pressure vs. rate produced. The
wellhead pressure can then be mapped to the separator pressure.
This approach relies on accurate well tests over a range of conditions for each well and then a
method (by hand or computer) of optimising the total field performance. The well tests would have
to be repeated on a regular basis and whether conditions changed. Such a situation is not always
possible due to the lack of a test separator or it being required for other purposes or the shear
number of wells.

3.8.1.2 Theoretical Methods


Prosper and GAP have the ability to model fairly complex fields with a variety of different wells
and to optimise the allocation of resources in the optimum way.
In order to do this accurately, the wells and facilities must be accurately modelled and the
performance predictions must be accurate. Particular care should be exercised in tuning the lift
curve correlations on individual wells. It is recommended that a generic well model is constructed to
best match all the wells and that if tuning is required it is then based on a large dataset and
individual measurement errors on each well should not grossly effect the result. Refer to section
3.7.3 for further details on tuning correlations.

3.8.1.3 Combined Method


A theoretical model such as GAP is flexible in that it can offer predictions and answers to what-if
scenarios and is easily changed. It is vital that the model is however compared with reality. It is
particularly recommended to use multi-rate well tests on gas lifted wells as frequently there is
discrepancies between theory and reality. If there are discrepancies, then these can be investigated in
terms of the model, facilities measuring errors and the likes.

Page 55
Section 3 – Well Performance

3.9 References
Prosper Software
Prosper Software

Integrated Asset Modelling

Integrated Asset Modelling

Near Wellbore Completion Manual


Near Wellbore Completion Manual

Sand Control Guidelines


Sand Control Guidelines

Erosion Guidelines

Erosion Guidelines Revision 3

Stimulations Website
Stimulations

Page 56
Section 3 – Well Performance

Page 57

You might also like