You are on page 1of 96

Philip Ringrose

Mark Bentley

Reservoir
Model Design
A Practitioner’s Guide
The Rock Model
2

Abstract
This topic concerns the difference between a reservoir model and a
geological model. Model representation is the essential issue – ask your-
self whether the coloured cellular graphics we see on the screen truly
resemble the reservoir as exposed in outcrop:
WYSIWYG (computing acronym).
Our focus is on achieving a reasonable representation.
Most of the outputs from reservoir modelling are quantitative and
derive from property models, so the main purpose of a rock model is to
get the properties in the right place – to guide the spatial property
distribution in 3D.
For certain model designs, the rock model component is minimal, for
others it is essential. In all cases, the rock model should be the guiding
framework and should offer predictive capacity to a project.

P. Ringrose and M. Bentley, Reservoir Model Design, DOI 10.1007/978-94-007-5497-3_2, 13


# Springer Science+Business Media B.V. 2015
14 2 The Rock Model

Outcrop view and model representation of the Hopeman Sandstone at Clashach Quarry, Moray Firth, Scotland

2.1 Rock Modelling these in 3D. This is often seen as the most ‘geo-
logical’ part of the model build along with the
In a generic reservoir modelling workflow, the fault modelling, and it is generally assumed that a
construction of a rock or ‘facies’ model usually ‘good’ final model is one which is founded on a
precedes the property modelling. Effort is thoughtfully-constructed rock model.
focussed on capturing contrasting rock types However, although the rock model is often
identified from sedimentology and representing essential, it is rarely a model deliverable in itself,
2.1 Rock Modelling 15

Fig. 2.1 To model rocks, or not to model rocks? Upper image: porosity model built directly from logs; middle image: a
rock model capturing reservoir heterogeneity; lower image: the porosity model rebuilt, conditioned to the rock model

and many reservoirs do not require rock models. precisely enough to guide the modelling: rock
Figure 2.1 shows a porosity model which has body databases are generally insufficient and
been built with and without a rock model. If the dip-log data too sparse to rely on as a model
upper porosity model is deemed a reasonable foundation. Most critical to the design are the
representation of the field, a rock model is not issues identified below, mishandling of which is
required. If, however, the porosity distribution is a common source of a poor model build:
believed to be significantly influenced by the • Reservoir concept – is the architecture
rock contrasts shown in the middle image, then understood in a way which readily translates
the lower porosity model is the one to go for. into a reservoir model?
Rock modelling is therefore a means to an end • Model elements – from the range of observed
rather than an end in itself, an optional step structural components and sedimentological
which is useful if it helps to build an improved facies types, has the correct selection of
property model. elements been made on which to base the
The details of rock model input are software- model?
specific and are not covered here. Typically the • Model Build – is the conceptual model car-
model requires specification of variables such as ried through intuitively into the statistical
sand body sizes, facies proportions and reference component of the build?
to directional data such as dip-logs. These are • Determinism and probability – is the bal-
part of a standard model build and need consid- ance of determinism and probability in the
eration, but are not viewed here as critical to the model understood, and is the conceptual
higher level issue of model design. Moreover, model firmly carried in the deterministic
many of these variables cannot be specified model components?
16 2 The Rock Model

These four questions are used in this chapter adopt when not representative, particularly if
to structure the discussion on the rock model, modern dynamic environments are being
followed by a summary of more specific rock compared with ancient preserved systems. It is
model build choices. possible to collect a library of analogue images
yet still be unclear exactly how these relate to
the reservoir in hand, and how they link to
the available well data. By contrast, the ability
2.2 Model Concept to draw a conceptual sketch section is highly
informative and brings clarity to the mental
The best hope of building robust and sensible image of the reservoir held by the modeller. If
models is to use conceptual models to guide the this conceptual sketch is not clear, the process
model design. We favour this in place of purely of model building is unlikely to make it any
data-driven modelling because of the issue of clearer. If there is no clear up-front conceptual
under-sampling (see later). The geologist should model then the model output is effectively a
have a mental picture of the reservoir and use random draw:
modelling tools to convert this into a quantitative
geocellular representation. Using system defaults If you can sketch it, you can model it
or treating the package as a black box that some- An early question to address is: “what are the
how adds value or knowledge to the model will fundamental building blocks for the reservoir
always result in models that make little or no concept?” These are referred to here as the
geological sense, and which usually have poor ‘model elements’ and discussed further below.
predictive capacity. For the moment, the key thing to appreciate is
The form of the reservoir concept is not com- that:
plex. It may be an image from a good outcrop
model elements 6¼ facies types
analogue or, better, a conceptual sketch, such as
those shown in Fig. 2.2. Selection of model elements is discussed in
It should, however, be specific to the case Sect. 2.4.
being modelled, and this is best achieved by With the idea of a reservoir concept as an
drawing a simple section through the reservoir architectural sketch constructed from model
showing the key architectural elements – an elements established, we will look at the issues
example of which is shown in Fig. 2.3. surrounding the build of the model framework
Analogue photos or satellite images are then return to consider how to select elements to
useful and often compelling but also easy to place within that framework.

Fig. 2.2 Capturing the reservoir concept in an analogue image or a block diagram sketch
2.3 The Structural and Stratigraphic Framework 17

Fig. 2.3 Capturing the reservoir concept in a simple sketch showing shapes and stacking patterns of reservoir sand
bodies and shales (From: van de Leemput et al. 1996)

the absence of new seismic, a fault model may be


2.3 The Structural passed on between users and adopted simply to
and Stratigraphic Framework avoid the inefficiency of repeating the manual
fault-building.
The structural framework for all reservoir models Such an inherited fault framework therefore
is defined by a combination of structural inputs requires quality control (QC). The principal
(faults and surfaces from seismic to impart gross question is whether the fault model reflects the
geometry) and stratigraphic inputs (to define seismic interpretation directly, or whether it has
internal layering). been modified by a conceptual structural
The main point we wish to consider here is what interpretation.
are the structural and stratigraphic issues that a A direct expression of a seismic interpretation
modeller should be aware of when thinking through will tend to be a conservative representation of
a model design? These are discussed below. the fault architecture, because it will directly
reflect the resolution of the data. Facets of such
data are:
2.3.1 Structural Data • Fault networks tend to be incomplete, e.g. faults
may be missing in areas of poor seismic quality;
Building a fault model tends to be one of the more • Faults may not be joined (under-linked) due to
time-consuming and manual steps in a modelling seismic noise in areas of fault intersections;
workflow, and is therefore commonly done with • Horizon interpretations may stop short of faults
each new generation of seismic interpretation. In due to seismic noise around the fault zone;
18 2 The Rock Model

• Horizon interpretations may be extended 2.3.2 Stratigraphic Data


down fault planes (i.e. the fault is not
identified independently on each horizon, or There are two main considerations in the selec-
not identified at all) tion of stratigraphic inputs to the geological
• Faults may be interpreted on seismic noise framework model: correlation and hierarchy.
(artefacts).
Although models made from such ‘raw’ seis- 2.3.2.1 Correlation
mic interpretations are honest reflections of that In the subsurface, correlation usually begins with
data, the structural representations are incom- markers picked from well data – well picks.
plete and, it is argued here, a structural interpre- Important information also comes from correla-
tation should be overlain on the seismic outputs tion surfaces picked from seismic data. Numer-
as part of the model design. To achieve this, the ous correlation picks may have been defined in
workflow similar to that shown in Fig. 2.4 is the interpretation of well data and these picks
recommended. may have their origins in lithological, biostrati-
Rather than start with a gridded framework graphical or chronostratigraphical correlations –
constructed directly from seismic interpretation, all of these being elements of sequence stratigra-
the structural build should start with the raw, phy (see for example Van Wagoner et al. 1990;
depth-converted seismic picks and the fault Van Wagoner and Bertram 1995). If multiple
sticks. This is preferable to starting with horizon stratigraphic correlations are available these
grids, as these will have been gridded without may give surfaces which intersect in space.
access to the final 3D fault network. Working Moreover, not all these surfaces are needed in
with pre-gridded surfaces means the starting reservoir modelling. A selection process is there-
inputs are smoothed, not only within-surface fore required. As with the structural framework,
but, more importantly, around faults, the latter the selection of surfaces should be made with
tending to have systematically reduced fault reference to the conceptual sketch, which is in
displacements. turn driven by the model purpose.
A more rigorous structural model workflow is As a guideline, the ‘correct’ correlation lines
as follows: are generally those which most closely govern
1. Determine the structural concept – are faults the fluid-flow gradients during production. An
expected to die out laterally or to link? Are en exception would be instances where correlation
echelon faults separated by relay ramps? Are lines are used to guide the distribution of reser-
there small, possibly sub-seismic connecting voir volumes in 3D, rather than to capture correct
faults? fluid flow units.
2. Input the fault sticks and grid them as fault The choice of correlation surfaces used
planes (Fig. 2.4a) hugely influences the resulting model architec-
3. Link faults into a network consistent with the ture, as illustrated in Fig. 2.5, and in an excellent
concept (1, above, also Fig. 2.4b) field example by Ainsworth et al. (1999).
4. Import depth-converted horizon picks as
points and remove spurious points, e.g. those 2.3.2.2 Hierarchy
erroneously picked along fault planes rather Different correlation schemes have different
than stratigraphic surfaces (Fig. 2.4c) influences on the key issue of hierarchy, as the
5. Edit the fault network to ensure optimal posi- stratigraphy of most reservoir systems is
tioning relative to the raw picks; this may be an inherently hierarchical (Campbell 1967). For
iterative process with the geophysicist, particu- example, for a sequence stratigraphic correlation
larly if potentially spurious picks are identified scheme, a low-stand systems tract might have a
6. Grid surfaces against the fault network length-scale of tens of kilometres and might con-
(Fig. 2.4d). tain within it numerous stacked sand systems
2.3 The Structural and Stratigraphic Framework 19

Fig. 2.4 A structural build


based on fault sticks from
seismic (a), converted
into a linked fault system
(b), integrated with depth-
converted horizon picks
(c) to yield a conceptually
acceptable structural
framework which honours
all inputs (d). The
workflow can equally well
be followed using time
data, then converting to
depth using a 3D velocity
model. The key feature of
this workflow is the
avoidance of intermediate
surface gridding steps
which are made
independently of the final
interpreted fault network.
Example from the Douglas
Field, East Irish Sea
(Bentley and Elliott 2008)
20 2 The Rock Model

Fig. 2.5 Alternative (a) chronostratigraphic and (b) lithostratigraphic correlations of the same sand observations in
three wells; the chronostratigraphic correlation invokes an additional hierarchical level in the stratigraphy

with a length-scale of kilometres. These sands in So which is the preferred stratigraphic tool to
turn act as the bounding envelope for individual use as a framework for reservoir modelling? The
reservoir elements with dimensions of tens to quick answer is that it will be the framework
hundreds of metres. which most readily reflects the conceptual reser-
The reservoir model should aim to capture the voir model. Additional thought is merited, how-
levels in the stratigraphic hierarchy which influ- ever, particularly if the chronostratigraphic
ence the spatial distribution of significant approach is used. This method yields a frame-
heterogeneities (determining ‘significance’ will work of timelines, often based on picking the
be discussed below). Bounding surfaces within most shaly parts of non-reservoir intervals. The
the hierarchy may or may not act as flow barriers intended shale-dominated architecture may not
– so they may represent important model automatically be generated by modelling
elements in themselves (e.g. flooding surfaces) algorithms, however: a rock model for an inter-
or they may merely control the distribution of val between two flooding surfaces will contain a
model elements within that hierarchy. This shaly portion at both the top and the base of the
applies to structural model elements as well as interval. The probabilistic aspects of the
the more familiar sedimentological model subsequent modelling can easily degrade the cor-
elements, as features such as fracture density relatable nature of the flooding surfaces, inter-
can be controlled by mechanical stratigraphy – well shales becoming smeared out incorrectly
implicitly related to the stratigraphic hierarchy. throughout the zone.
2.3 The Structural and Stratigraphic Framework 21

Fig. 2.6 The addition of hierarchy by logical combina- right, yellow ¼ better quality reservoir sands), logically
tion: single-hierarchy channel model (top left, blue ¼ combined into the final rock model with lithofacies
mudstone, yellow ¼ main channel) built in parallel detail in the main channel only – an additional level of
with a probabilistic model of lithofacies types (top hierarchy

Some degree of hierarchy is implicit in any lithofacies types within a particular model element
software package. The modeller is required to (the main channels). The chosen solution was to
work out if the default hierarchy is sufficient to build the channel model using channel objects and
capture the required concept. If not, the workflow creating a separate, in this case probabilistic,
should be modified, most commonly by applying model which contained the information about the
logical operations. distribution of the two lithofacies types. The two
An example of this is illustrated in Fig. 2.6, rock models were then combined using a logical
from a reservoir model in which the first two property model operation, which imposed the tex-
hierarchical levels were captured by the default ture of the fine-scale lithofacies, but only within
software workflow: tying layering to seismic the relevant channels. Effectively this created a
horizons (first level) then infilled by sub-seismic third hierarchical level within the model.
stratigraphy (second level). An additional hierar- One way or another hierarchy can be
chical level was required because an important represented, but only rarely by using the default
permeability heterogeneity existed between model workflow.
22 2 The Rock Model

or diagenetic features, often (perhaps incorrectly)


2.4 Model Elements excluded from descriptions of ‘facies’.
Modelling elements are defined here as:
Having established a structural/stratigraphic model
three-dimensional rock bodies which are
framework, we can now return to the model con-
petrophysically and/or geometrically distinct
cept and consider how to fill the framework to from each other in the specific context of the res-
create an optimal architectural representation. ervoir fluid system.

The fluid-fill factor is important as it


highlights the fact that different levels of hetero-
2.4.1 Reservoir Models Not Geological geneity are important for different types of fluid,
Models e.g. gas reservoirs behave more homogeneously
than oil reservoirs for a given reservoir type.
The rich and detailed geological story that can The identification of ‘model elements’ has
be extracted from days or weeks of analysis of some parallels with discussions of ‘hydraulic
the rock record from the core store need not be units’ although such discussions tend to be in
incorporated directly into the reservoir model, the context of layer-based well performance.
and this is a good thing. There is a natural ten- Our focus is on the building blocks for 3D reser-
dency to ‘include all the detail’ just in case voir architecture, including parts of a field
something minor turns out to be important. remote from well and production data. It should
Models therefore have a tendency to be over- be spatially predictive.
complex from the outset, particularly for novice
modellers. The amount of detail required in the
model can, to a large extent, be anticipated.
There is also a tendency for modellers to seize 2.4.3 Model Element Types
the opportunity to build ‘real 3D geological
pictures’ of the subsurface and to therefore make Having stepped beyond a traditional use of
these as complex as the geology is believed to be. depositional facies to define rock bodies for
This is a hopeless objective as the subsurface is modelling, a broader spectrum of elements can
considerably more complex in detail than we are be considered for use, i.e. making the sketch of
capable of modelling explicitly and, thankfully, the reservoir as it is intended to be modelled. Six
much of that detail is irrelevant to economic or types of model element are considered below.
engineering decisions. We are building reservoir
models – reasonable representations of the 2.4.3.1 Lithofacies Types
detailed geology – not geological models. This is sedimentologically-driven and is the tra-
ditional way of defining the components of a rock
model. Typical lithofacies elements may be
2.4.2 Building Blocks coarse sandstones, mudstones or grainstones,
and will generally be defined from core and or
Hence the view of the components of a reservoir log data (e.g. Fig. 2.7).
model as model elements – the fundamental
building blocks of the 3D architecture. The use 2.4.3.2 Genetic Elements
of this term distinguishes model elements from In reservoir modelling, genetic elements are a
geological terms such as ‘facies’, ‘lithofacies’, component of a sedimentary sequence which
‘facies associations’ and ‘genetic units’. These are related by a depositional process. These
geological terms are required to capture the rich- include the rock bodies which typical modelling
ness of the geological story, but do not necessar- packages are most readily designed to incorpo-
ily describe the things we need to put into rate, such as channels, sheet sands or
reservoir models. Moreover, key elements of heterolithics. These usually comprise several
the reservoir model may be small-scale structural lithofacies, for example, a fluvial channel might
2.4 Model Elements 23

Fig. 2.7 Example


lithofacies elements; left:
coarse, pebbly sandstone;
right: massively-bedded
coarse-grained sandstone

0 GR 150 56 CNL -4
channel USF LSF CAL 16 FDC
0 1.65 2.65

FT

7900

CHANNEL
LSF
3 ft

USF
SHALE
7950

8000

Fig. 2.8 Genetic modelling elements; lithofacies types grouped into channel, upper shoreface and lower shoreface
genetic depositional elements (Image courtesy of Simon Smith)

include conglomeratic, cross-bedded sandstone 2.4.3.3 Stratigraphic Elements


and mudstone lithofacies. Figure 2.8 shows an For models which can be based on a sequence
example of several genetic depositional elements stratigraphic framework, the fine-scale components
interpreted from core and log observations. of the stratigraphic scheme may also be the
24 2 The Rock Model

Fig. 2.9 Sequence stratigraphic elements

predominant model elements. These may be 2.4.3.6 Exotic Elements


parasequences organised within a larger-scale The list of potential model elements is as diverse
sequence-based stratigraphic framework which as the many different types of reservoir, hence
defines the main reservoir architecture (e.g. other ‘exotic’ reservoir types must be mentioned,
Fig. 2.9). having their own model elements specific to their
geological make-up. Reservoirs in volcanic rocks
are a good example (Fig. 2.12), in which the key
2.4.3.4 Diagenetic Elements
model elements may be zones of differential
Diagenetic elements commonly overprint
cooling and hence differential fracture density.
lithofacies types, may cross major stratigraphic
***
boundaries and are often the predominant feature
The important point about using the term
of carbonate reservoir models. Typical diage-
‘model element’ is to stimulate broad thinking
netic elements could be zones of meteoric
about the model concept, a thought process which
flushing, dolomitisation or de-dolomitisation
runs across the reservoir geological sub-disciplines
(Fig. 2.10).
(stratigraphy, sedimentology, structural geology,
even volcanology). For avoidance of doubt, the
2.4.3.5 Structural Elements main difference between the model framework
Assuming a definition of model elements as and the model elements is that 2D features are
three-dimensional features, structural model used to define the model framework (faults, uncon-
elements emerge when the properties of a vol- formities, sequence boundaries, simple bounding
ume are dominated by structural rather than surfaces) whereas it is 3D model elements which
sedimentological or stratigraphic aspects. Fault fill the volumes within that framework.
damage zones are important volumetric struc- Having defined the framework and identified the
tural elements (e.g. Fig. 2.11) as are mechanical elements, the next question is how much information
layers (strata-bound fracture sets) with properties to carry explicitly into the modelling process. Every-
driven by small-scale jointing or cementation. thing that can be identified need not be modelled.
2.4 Model Elements 25

Fig. 2.10 Diagenetic elements in a carbonate build-up; where reservoir property contrasts are driven by differential
development of dolomitisation

Bounding
fault

Slip surfaces/
Isolated small deformation
faults and bands in best
deformation sands – some
bands open (blue)

A
e1
Zon
C
B e1
e1 Zon e2
A
Zon Zon

OWC
Minor faulting in
poorer sands
(open?)

Damage zone

Fig. 2.11 Structural elements: volumes dominated by minor fracturing in a fault damage zone next to a major
block-bounding fault (Bentley and Elliot 2008)

2.4.4 How Much Heterogeneity the key criteria for distinguishing which
to Include? model elements are required for the model
build are:
The ultimate answer to this fundamental ques- 1. The identification of potential model elements –
tion depends on a combined understanding of a large number may initially be selected as
geology and flow physics. To be more specific, ‘candidates’ for inclusion;
The Property Model
3

Abstract
Now let’s say you have a beautiful fit-for-purpose rock model of your
reservoir – let’s open the box and find out what’s inside? All too often the
properties used within the geo-model are woefully inadequate.
The aim of this chapter is too ensure the properties of your model are
also fit-for-purpose and not, like Pandora’s box, full of “all the evils of
mankind.”
Eros warned her not to open the box once Persephone’s beauty was inside[. . .]
but as she opened the box Psyche fell unconscious upon the ground. (From The
Golden Ass by Apuleius.)

P. Ringrose and M. Bentley, Reservoir Model Design, DOI 10.1007/978-94-007-5497-3_3, 61


# Springer Science+Business Media B.V. 2015
62 3 The Property Model

Pixelated rocks

3.1 Which Properties? A geological model of a petroleum reservoir


is the basis for most reservoir evaluation and
First let us recall the purpose of building a reser- engineering decisions. These include (roughly
voir model in the first place. We propose that in order of complexity and detail):
the overall aim in reservoir model design is: • Making estimates of fluid volumes in place,
To capture knowledge of the subsurface in a quan- • Scoping reservoir development plans,
titative form in order to evaluate and engineer the • Defining well targets,
reservoir.
• Designing detailed well plans,
This definition combines knowledge capture, • Optimising fluid recovery (usually for IOR/
the process of collecting all relevant information, EOR schemes).
with the engineering objective – the practical The type of decision involved affects the
outcome of the model (Fig. 3.1). Deciding how property modelling approach used. Simple aver-
to do this is the job of the geo-engineer – a aging or mapping of properties is more likely to
geoscientist with sufficient knowledge of the be appropriate for initial volume estimates while
Earth and the ability to quantify that knowledge advanced modelling with explicit upscaling is
in a way that is useful for the engineering deci- mostly employed when designing well plans
sion at hand. A mathematician, physicist or engi- (Fig. 3.3) or as part of improved reservoir dis-
neer with sufficient knowledge Earth science can placement plans or enhanced oil recovery (EOR)
make an equally good geo-engineer (Fig. 3.2). strategies.
Knowledge capture The model Engineering decisions

What do you know about How much oil/gas?


fluid distributions How producible?
geology & structure? How certain are you?
Is it economic?
What data have you got?
How uncertain is it?

Fig. 3.1 Knowledge capture and the engineering decision

Fig. 3.2 The geo-engineer (Photo, Statoil image archive, # Statoil ASA, reproduced with permission)

Fig. 3.3 Defining the well target


64 3 The Property Model

The next question is which petrophysical


properties do we want to model? The focus in Exercise 3.1
this chapter will be on modelling porosity (ϕ) Which methods to use?
and permeability (k) as these are the essential Think through the following decision
parameters in the flow equation (Darcy’s law). matrix for an oilfield development to
The methods discussed here for handling ϕ and k decide which approaches are appropriate
can also be applied to other properties, such as for which decisions?
formation bulk density (ρb) or sonic p-wave
Method (for a
velocity (vp), Volume fraction of shale (Vshale) given reservoir
or fracture density (Fd), to name but a few. interval) Choice Purpose
Table 3.1 lists the most commonly modelled Conceptual Initial fluids-in-
rock properties, but the choice should not be geological sketch place volume
limited to these, and indeed a key element of of proposed estimate
reservoir analogue
the design should be careful consideration of
Simple average Preliminary
which properties should or can be usefully of porosity, ϕ, reserves estimates
represented. Integration of dynamic data with permeability, k,
seismic and well data will generally require and fluid
saturation, Sw
modelling of several petrophysical properties
2D map of ϕ, k Reserve estimates
and their cross correlations. and Sw (e.g. for designing top-
Permeability is generally the most chal- interpolation or side facilities
lenging property to define because it is highly kriging between (number of wells,
variable in nature and is a tensor property depen- wells) platform type)
dent on flow boundary conditions. Permeability 3D model of ϕ, k Definition of
and Sw in the appraisal well
is also, in general, a non-additive property, reservoir unit drilling plan
that is: (from well data)
X
n 3D model of ϕ, k Definition of infill
kΔV 6¼ k∂v
i ð3:1Þ and Sw for each of or development
1 several model well drilling plan
elements (from
In contrast porosity is essentially an additive well data)
property: 3D model of ϕ, k Submitting
and Sw detailed well
X
n
ϕΔV ¼ ϕ∂v
i ð3:2Þ conditioned to design for final
seismic inversion approval
1
cube (seismic
where ΔV is a large scale volume, δv[n] is the facies)
exhaustive set of small scale volumes filling the 3D model of ϕ, k, Designing
Sw and facies improved oil
large-scale volume.
conditioned to recovery (IOR)
Put in practical terms, if you have defined dynamic data strategy and
all the cell porosity values in your reservoir (production additional well
model then the total reservoir porosity is pre- pressures and flow targets
rates)
cisely equal to the sum of the cell porosities
3D model of ϕ, k, Implementing
divided by the number cells (i.e. the average), Sw and facies enhanced oil
whereas for permeability this is not the case. integrating multi- recovery (EOR)
We will discuss appropriate use of various scale static and strategy using an
permeability averages in following section. dynamic data injection blend
(e.g. water
alternating gas)
3.1 Which Properties? 65

Table 3.1 List of properties typically included in geological reservoir models


66 3 The Property Model

The final important question to address is: volumes and flow units in Chap. 4 when we
Which reservoir or rock unit do we want to look at upscaling, but first we need to understand
average? There are many related concepts used permeability.
to define flowing rock intervals – flow units,
hydraulic units, geological units or simply “the
reservoir”. The most succinct term for defining 3.2 Understanding Permeability
the rock units in reservoir studies is the Hydraulic
Flow Unit (HFU), which is defined as represen- 3.2.1 Darcy’s Law
tative rock volume with consistent petrophysical
properties distinctly different from other rock The basic permeability equation is based on the
units. There is thus a direct relationship between observations and field experience of Henri Darcy
flow units and the ‘model elements’ introduced (1803–1858) while engineering a pressurized
in the preceding chapter. water distribution system in the town of Dijon,
France. His equation relates flow rate to the head
Exercise 3.2 of water draining through a pile of sand (Fig. 3.4):
Additive properties Q ¼ KAðΔH=LÞ ð3:3Þ
Additivity involves a mathematical
where
function in which a property can be
Q ¼ volume flux of water
expressed as a weighted sum of some inde-
K ¼ constant of hydraulic conductivity or coef-
pendent variable(s). The concept is impor-
ficient of permeability
tant to a wide range of statistical methods
A ¼ cross sectional area
used in many science disciplines. Additiv-
ΔH ¼ height of water column
ity has many deeper facets and definitions
L ¼ length of sand column
that are discussed in mathematics and sta-
tistical literature.
From this we can derive the familiar Darcy’s
It is useful to consider a wider selection
Law – a fundamental equation for flow in porous
of petrophysical properties and think
media, based on dimensional analysis and the
through whether they are essentially addi-
Navier-Stokes equations for flow in cylindrical
tive or non-additive (i.e. multiplicative)
pores:
properties.
k
What would you conclude about these u¼ ∇ðP þ ρgzÞ ð3:4Þ
terms? μ
• Net-to-gross ratio
• Fluid saturation
• Permeability
• Porosity Water
• Bulk density
• Formation resistivity
• Seismic velocity, Vp or Vs ΔH Sand
• Acoustic Impedance, AI
L

Abbaszadeh et al. (1996) define the HFU in


terms of the Kozeny-Carmen equation to extract
Flow Zone Indicators which can be used quanti-
tatively to define specific HFUs from well data. Fig. 3.4 Darcy’s Q
We will return the definition of representative experiment
3.2 Understanding Permeability 67

where concept is widely used, and abused, and requires


u ¼ intrinsic fluid velocity some care in its use and application. It is also
k ¼ intrinsic permeability rather fundamental – if it was a simple thing to
μ ¼ fluid viscosity estimate the correctly upscaled permeability for a
∇P ¼ applied pressure gradient reservoir unit, there would be little value in res-
ρgz ¼ pressure gradient due to gravity ervoir modelling (apart from simple volume
∇P (grad P) is the pressure gradient, which can estimates).
be solved in a cartesian coordinate system as: The upscaled (or block) permeability, kb,
is defined as the permeability of an homoge-
dP dP dP
∇P ¼ þ þ ð3:5Þ neous block, which under the same pressure
dx dy dz
boundary conditions will give the same average
The pressure gradient due to gravity is then flows as the heterogeneous region the block is
ρg∇z. For a homogeneous, uniform medium k representing (Fig. 3.5). The upscaled block per-
has a single value, which represents the meability could be estimated, given a fine set of
medium’s ability to permit flow (independent of values in a permeability field or model, or it
the fluid type). For the general case of a hetero- could be measured at the larger scale (e.g. in a
geneous rock medium, k is a tensor property. well test or core analysis), in which case the fine-
scale permeabilities need not be known.
The effective permeability is defined strictly in
Exercise 3.3 terms of effective medium theory and is an
Dimensions of permeability intrinsic large-scale property which is indepen-
What are the dimensions of perme- dent of the boundary conditions. The main theo-
ability? Do a dimensional analysis for retical conditions for estimation of the effective
Darcy’s Law. permeability, keff, are:
For the volumetric flux equation • That the flow is linear and steady state;
Q ¼ KAðΔH=LÞ • That the medium is statistically homogeneous
at the large scale.
The dimensions are When the upscaled domain is large enough,
 3 1   1   2  such that these conditions are nearly satisfied,
L T ¼ LT L
then kb approaches keff. The term equivalent
Therefore the SI unit for K is: permeability, is also used (Renard and de Marsily
ms 1 1997) and refers to a general large-scale
Do the same for Darcy’s Law: permeability which can be applied to a wide
range of boundary conditions, to some extent
u ¼ ðk=μÞ:∇ðP þ ρgzÞ encompassing both kb and keff. These terms are
The dimensions are often confused or misused, and in this treatment
we will refer to the permeability upscaled from
½  ¼ ð½ =½ Þ:½  a model as the block permeability, kb, and use
Therefore the SI unit for k is effective permeability as the ideal upscaled per-
____ meability we would generally wish to estimate
if we could satisfy the necessary conditions.
In reservoir modelling we are usually estimating
3.2.2 Upscaled Permeability kb in practice, because we rarely fully satisfy
the demands of effective medium theory. How-
In general terms, upscaled permeability refers to ever, keff is an important concept with many
the permeability of a larger volume given some constraints that we try to satisfy when estimating
fine scale observations or measurements. The the upscaled (block) permeability.
68 3 The Property Model

Fig. 3.5 Effective permeability and upscaled block permeability (a) Real rock medium has some (unknown)
effective permeability. (b) Modelled rock medium has an estimated block permeability with the same average flow
as the real thing

Full permeability tensor

é kxx kxy kxzù


ê ú
Permeable medium keff = êkyx kyy kyzú Coordinates
x
ê kzx kzy kzz ú y
ë û
z

Diagonal permeability tensor

ékxx 0 0ù
keff = 0 kyy 0 ú
ê
ê ú
êë 0 0 kzzúû

Fig. 3.6 The permeability tensor

Note that in petrophysical analysis, the term automatically associated. For example, in an
‘effective porosity’ refers to the porosity of move- upscaled heterogeneous volume there could be
able fluids excluding micro-porosity and chemi- effective porosity elements (e.g. vuggy pores)
cally bound water, while total porosity which do not contribute to the flow and therefore
encompasses all pore types. Although effective do not influence the effective permeability.
porosity and effective permeability both represent In general, kb is a tensor property (Fig. 3.6)
properties relevant to, and controlling, where, for example, kxy represents flow in the
macroscopic flow, they are defined on different x direction due to a pressure gradient in the
bases. Effective permeability is essentially a y direction. In practice kb is commonly assumed
larger scale property requiring statistical homoge- to be a diagonal tensor where off-diagonal terms
neity in the medium, whereas effective porosity is are neglected. A further simplification in many
essentially a pore-scale physical attribute. Of reservoir modelling studies is the assumption
course, if both properties are estimated at, or that kh ¼ kxx ¼ kyy and that kv ¼ kzz.
rescaled to, the same appropriate volume, they The calculation or estimation of kb is depen-
may correspond and are correctly used together dent on the boundary conditions (Fig. 3.7). Note
in flow modelling. They should not, however, be that the assumption of a no-flow or sealed
3.2 Understanding Permeability 69

side boundary condition, forces the result to be a In reality, the permeability in a rock medium is
diagonal tensor. This is useful, but may not of a highly variable property. In sedimentary basins
course represent reality. Renard and de Marsily as a whole we expect variations of at least 10
(1997) give an excellent review of effective per- orders of magnitude (Fig. 3.8), with a general
meability, and Pickup et al. (1994, 1995) give decrease for surface to depth due to compaction
examples of the permeability tensor estimated and diagenesis. Good sandstone units may have
for a range of realistic sedimentary media. permeabilities typically in the 10–1,000 mD
range, but the silt and clay rich units pull the
permeability down to around 103 mD or lower.
3.2.3 Permeability Variation Deeply buried mudstones forming cap-rocks and
in the Subsurface seals have permeabilities in the microdarcy to
nanodarcy range. Even within a single reservoir
There is an extensive literature on the measurement unit (not including the shales), permeability may
of permeability (e.g. Goggin et al. 1988; Hurst and range by at least 5 orders of magnitude. In the
Rosvoll 1991; Ringrose et al. 1999) and its appli- example shown in Fig. 3.9 the wide range in
cation for reservoir modelling (e.g. Begg et al. observed permeabilities is due both to lithofacies
1989; Weber and van Geuns 1990; Corbett et al. (heterolithic facies tend to be lower than the sand
1992). All too often, rather idealised permeability facies) and due to cementation (each facies is
distributions have been assumed in reservoir highly variable mainly due to the effects of vari-
models, such as a constant value or the average of able degrees of quartz cementation).
a few core plug measurements.

No-flow boundary Open boundary 3.2.4 Permeability Averages

Due to its highly variable nature, some form of


P1 P2 P1 P2 averaging of permeability is generally needed.
The question is which average? There are well-
known limits for the estimation of keff in ideal
Fig. 3.7 Simple illustration of flow boundary conditions:
systems. For flow along continuous parallel layers
P1 and P2 are fluid pressures applied at the left and right the arithmetic average gives the correct effective
hand sides and arrows illustrate flow vectors permeability, while for flow perpendicular to

Permeability (md)
10−6 10−3 1 103
Homogeneous sand

Holocene aquifers:
1 Fluvial sand

1 Silty clay

1 Clay

Fig. 3.8 Typical ranges of North Sea oil reservoirs:


permeability for near-
surface aquifers and North 2 Fluvial
Sea oil reservoirs:
1 ¼ Holocene aquifers 2 Deltaic
(From Bierkins 1996),
? Deep Basin mudstones
2 ¼ Example North Sea
datasets (anonymous)
70 3 The Property Model

0.20
Clean sands
Heterolithic sands
0.15
Sandy Heterolithics
Freq. Mixed Heterolithicss
0.10

0.05

0.00

10
1

100

1000
time-out

0.1

10000

100000
Permeability (md)

Fig. 3.9 Probe permeameter measurements from a highly variable, deeply-buried, tidal-deltaic reservoir interval
(3 m of core) from offshore Norway

Fig. 3.10 Calculation of


effective permeability
using averages for ideal
layered systems: (a) The
arithmetic average for flow
along continuous parallel
layers; (b) The harmonic
average for flow
perpendicular to
continuous parallel layers
(ki and ti are the
permeability and thickness
of layer i)

continuous parallel layers the harmonic average is More precise limits to keff have also been
the correct solution (Fig. 3.10). proposed, such as the arithmetic mean of har-
If the layers are in any way discontinuous or monic means of each row of cells parallel to
variable or the flow is not perfectly parallel or flow (lower bound) and vice versa for the upper
perpendicular to the layers then the true effective bound (Cardwell and Parsons 1945). However,
permeability will lie in between these averages. for most practical purposes the arithmetic and
This gives us the outer bounds to effective harmonic means are quite adequate limiting
permeability: values, especially given that we seldom have an
exhaustive set of values to average (the sample
kharmonic  keff  karithmetic
problem, discussed in Sect. 3.3 below).
3.2 Understanding Permeability 71

The geometric average is often proposed as a where p ¼ 1 corresponds to the harmonic


useful or more correct average to use for more mean, p ~ 0 to the geometric mean and p ¼ 1
variable rock systems. Indeed for flow in a to the arithmetic mean (p ¼ 0 is invalid and the
correlated random 2D permeability field with a geometric mean is calculated using Eq. (3.6)).
log-normal distribution and a low variance the For a specific case with some arbitrary hetero-
effective permeability is equal to the geometric geneity structure, a value for p can be found (e.g.
mean: by finding a p value which gives best fit to results
" # of numerical simulations). This can be a very
X
n
useful form of the permeability average. For
kgeometric ¼ exp lnki =n ð3:6Þ
1
example, after some detailed work on estimating
the permeability of a particular reservoir unit or
This can be adapted for 3D as long as account facies (based on a key well or near-well model)
is also taken for the variance of the distribution. one can derive plausible values for p for general
Gutjahr et al. (1978) showed that for a log- application in the full field reservoir model (e.g.
normally distributed permeability field in 3D: Ringrose et al. 2005). In general, p for kh will be
  positive and p for kv will be negative.
keff ¼ kgeometric 1 þ σ 2 =6 ð3:7Þ
Note that for the general case, when applying
where σ2 is the variance of ln(k). averages to numerical models with varying cell
Thus in 3D, the theoretical effective perme- sizes, we use volume weighted averages. Thus,
ability is slightly higher than the geometric aver- the most general form of the permeability esti-
age, or indeed significantly higher if the variance mate using averages is:
Z Z 1=p
is large.
An important condition for keff  kgeometric is kestimate ¼ kp dV dV h1 < p < 1j
that correlation length, λ, of the permeability
ð3:9Þ
variation must be significantly smaller than the
size of the averaging volume, L. That is: where p is estimated or postulated.

λx λ y λz  L x L y L z
3.2.5 Numerical Estimation of Block
This relates to the condition of statistical Permeability
homogeneity. In practice, we have found that λ
needs to be at least 5 times smaller than L for For the general case, where an average perme-
kb ! kgeometric for a log-lognormal permeability ability cannot be assumed, a priori, numerical
field. This implies that the assumption (some- methods must be used to calculate the block
times made) that kgeometric is the ‘right’ average permeability (kb). This subject has occupied
for a heterogenous reservoir interval is not gen- many minds in the fields of petroleum and
erally true. Neither does the existence of a log- groundwater engineering and there is a large
normal permeability distribution imply that the literature on this subject. The numerical methods
geometric average is the right average. This is used are based on the assumptions of conserva-
evident in the case of a perfectly layered system tion of mass and energy, and generally assume
with permeability values drawn from a log nor- steady-state conditions. The founding father of
mal distribution – in such a case keff ¼ karithmetic. the subject in the petroleum field is arguably
Averages between the outer-bound limits to keff Muskat (1937), while Matheron (1967) founded
can be generalised in terms of the power average much of the theory related to estimation of flow
(Kendall and Staurt 1977; Journel et al. 1986): properties. De Marsilly (1986) gives an excellent
hX i1=p foundation from a groundwater perspective and
kpower ¼ kip =n ð3:8Þ Renard and de Marsily (1997) give a more recent
review on the calculation of equivalent
72 3 The Property Model

Fig. 3.11 Periodic P(i,0)=P(i,nz)


pressure boundary x
conditions applied to a
periodic permeability field,
involving an inclined layer.
Example boundary cell z
pressure conditions are
shown
P(0,j)=P(nx,j)+DP P(nx+1,j)=P(1,j)-DP

P(i,nz+1)=P(i,1)

permeability. Some key papers on the calculation Usually, at the fine scale we assume the local
of permeability for heterogeneous rock media permeability is not a tensor so that only one value
include White and Horne (1987), Durlofsky of k is required per cell.
(1991) and Pickup et al. (1994). We then wish to know the upscaled block per-
To illustrate the numerical approach we take an meability for the whole system. This is a relatively
example proposed by Pickup and Sorbie (1996) simple step once all small scale Darcy equations
shown in Fig. 3.11. Assuming a fine-scale grid of are known, and involves the following steps:
permeability values, ki, we want to calculate the 1. Solve the fine-scale equations to give pressures,
upscaled block permeability tensor, kb. An Pij for each block.
assumption on the boundary conditions must be 2. Calculate inter-block flows in the x-direction,
made, and we will assume a period boundary con- using Darcy’s Law.
dition (Durlofsky 1991) – where fluids exiting one 3. Calculate total flow, Q, by summing individual
edge are assumed to enter the opposite edge – and flows between any two planes.
apply this to a periodic permeability field (where 4. Calculate kb using Darcy’s Law applied to the
the model geometry repeats in all directions). This upscaled block.
arrangement of geometry and boundary conditions 5. Repeat for the y and z directions.
gives us an exact solution. For the upscaled block this results in a set of
First a pressure gradient ΔP is applied to the terms governing flow in each direction, such that:
boundaries in the x direction. For the boundaries 0 1
parallel to the applied pressure gradient, the peri- 1@ ∂P ∂P ∂PA
ux ¼  kxx þ kxy þ kxz
odic condition means that P in cell (i, 0) is set as μ ∂x ∂y ∂z
equal to P in cell (i, nz), where n is the number of 0 1
cells. A steady-state flow simulation is carried 1@ ∂P ∂P ∂PA
out on the fine-scale grid, and as all the uy ¼  kyx þ kyy þ kyz ð3:11Þ
μ ∂x ∂y ∂z
permeabilities are known, it is possible to find 0 1
the cell pressures and flow values (using matrix 1@ ∂P ∂P ∂PA
computational methods). uz ¼  kzx þ kzy þ kzz
μ ∂x ∂y ∂z
We then solve Darcy’s Law for each fine-
scale block: For example, the term kzx is the permeability in
! the z direction corresponding to the pressure gradi-
u ¼ ð1=μÞ k :∇P ð3:10Þ
ent in the x direction. These off-diagonal terms are
where intuitive when one looks at the permeability field.
! Take the vertical (x, z) geological model section
u is the local flow vector
shown in Fig. 3.12. If the inclined orange layers
μ is the fluid viscosity
have lower permeability, then flow applied in the
k is the permeability tensor
+x direction (to the right) will tend to generate a
ΔP is the pressure gradient flux in the –z direction (i.e. upwards). This results
3.2 Understanding Permeability 73

Note to graphics – Ensure grid edges align – redraft if necessary?


a Ripple laminaset
20mD

100mD 48 - 7.1
k
- 7.1 34

b Trough crossbed set


1200mD

100mD 759 - 63
k
- 63 336

Fig. 3.12 Example tensor permeability matrices calculated for simple 2D models of common sedimentary structures

in a vertical flow and requires a kzx permeability properties, and our aim is therefore to represent
term in the Darcy equation (for a 2D tensor). the expected flow behaviour. The effects of
Example solutions of permeability tensors geological architecture on flow are frequently
for simple geological models are given by neglected – for example, it may be assumed
Pickup et al. (1994) and illustrated in that a Gaussian random field represents the
Fig. 3.12. Ripple laminasets and trough cross- inter-well porosity and permeability architecture.
beds are two widespread bedding architectures More advanced, geologically-based, flow
found in deltaic and fluvial depositional settings modelling will, however, allow us to assess the
– ripple laminasets tend to be 2–4 cm in height potential effects of geological architecture on
while trough cross-bed sets are typically flow, and attempt to capture these effects as a
10–50 cm in height. These simple models are set of upscaled block permeability values. Struc-
two dimensional and capture typical geometry tural architecture in the form of fractures or small
and permeability variation (measured on out- faults may also generate pervasive tendencies for
crop samples) in a section parallel to the depo- strongly tensorial permeability within a rock
sitional current. In both cases, the tensor unit. By aligning grid cells to geological features
permeability matrices have relatively large (faults, dominant fracture orientations, or major
off-diagonal terms, 15 and 8 % of the kxx bed-set boundaries) the cross-flow terms can be
value, respectively. The negative off-diagonal kept to a minimum. However, typically one
terms reflect the chosen coordinate system aligns the grid to the largest-scale geological
with respect to flow direction (flow left to right architecture (e.g. major fault blocks) and so
with z increasing downwards). Vertical per- other smaller-scale features inevitably generate
meability is also significantly lower than the some cross-flow.
horizontal permeability due to the effects of
the continuous low permeability bottomset.
Geological elements like these will tend to fill 3.2.6 Permeability in Fractures
the volume within a particular reservoir unit,
imparting their flow anisotropy and cross-flow Understanding permeability in fractured
tendencies on the overall reservoir unit. Of reservoirs requires some different flow physics
course, real rock systems will have natural – Darcy’s law does not apply. Flow within a
variability in both architecture and petrophysical fracture (Fig. 3.13) is described by Poiseuille’s
74 3 The Property Model

For the explicit case, there are then several


options for how this may be done:
1. Discrete Fracture Network (DFN) models,
b where individual fractures with explicit geom-
etry are modelled in a complex network.
2. Dual permeability models, where the fracture
q and matrix permeability are explicitly
represented (but fracture geometry is implic-
itly represented by a shape factor).
w 3. Dual porosity models, where the fracture
L and matrix porosity are explicitly represented,
but the permeability is assumed to occur
Fig. 3.13 Flow in a fracture only in the fractures (and the fracture
geometry is implicitly represented by a
law, which for a parallel-plate geometry gives shape factor).
(Mourzenko et al. 1995): Fractured reservoir modelling is discussed in
detail by Nelson (2001) and covered in most
w b3 ΔP
q¼ ð3:12Þ reservoir engineering textbooks, and in Chap. 6
12μ L we describe approaches for handling fractured
reservoir models. The important thing to keep
where in mind in the context of understanding perme-
q is the volumetric flow rate, ability, is that fractures behave quite differently
w is the fracture width, (and follow different laws) from the general
b is the fracture aperture, Darcy-flow concept for flow in permeable (gran-
μ is the fluid viscosity, ular) rock media.
ΔP/L is the pressure gradient.
Note that the flow rate is proportional to b3,
and thus highly dependent on fracture aperture.
In practice, the flow strongly depends on the 3.3 Handling Statistical Data
stress state and the fracture roughness
(Witherspoon et al. 1980), but the underlying 3.3.1 Introduction
concept still holds. To put some values into this
simple equation – a 1 mm wide fracture in an Many misunderstandings about upscaled perme-
impermeable rock matrix would have an effec- ability, or any other reservoir property, are caused
tive permeability of around 100 Darcys. by incorrect understanding or use of probability
Unfortunately, fracture aperture is not easily distributions. The treatment of probability
measured, and generally has to be inferred from distributions is an extensive subject covered in a
pressure data. This makes fracture systems much number of textbooks. Any of the following are
harder to model than conventional non-fractured suitable for geoscientists and engineers wanting to
reservoirs. gain deeper appreciation of statistics and the Earth
In practice, there are two general approaches sciences: Size 1987; Isaaks and Srivastava 1989;
for modelling fracture permeability: Olea 1991; Jensen et al. 2000, and Davis 2003. Here
• Implicitly, where we model the overall rock we will identify some of the most important issues
permeability (matrix and fractures) and related to property modelling, namely:
assume we have captured the “effect of • Understanding sample versus population
fractures” as an effective permeability. statistics;
• Explicitly, where we represent the fractures in • Using log-normal and other transforms;
a model. • Use and implications of applying cut-off values.
3.3 Handling Statistical Data 75

Model
Data

Truth
Model 2

Data Reservoir
Model
(ground truth)

Model 3

Fig. 3.14 Illustration of the axiom: Data 6¼ Model 6¼ Truth (Redrawn from Corbett and Jensen 1992, #EAGE
reproduced with kind permission of EAGE Publications B.V., The Netherlands)

Our overall aim in reservoir modelling is to • ‘The previous models were all wrong, but this
estimate and compare distributions for: one must be right because it matches all the
1. The well data (observations); data we have.’
2. The reservoir model (a hypothesis or postulate); Now statement A sounds good but begs the
3. The population (the unknown “true” reservoir questions what cut-off was applied and is the
properties). geometric average indeed the appropriate aver-
We must always remember not to confuse age to use? Statement B is clearly arrogant but in
observations (data) with the model (a hypothesis) fact captures the psychology of every reservoir
and both of these with the “ground truth” (an model builder – we try to do our best with the
unknown). This leads us to one of the most available data but are reluctant to admit to the
important axioms of reservoir modelling: errors that must be present. Versions of these
statements that would be more consistent with
Data 6¼ Model 6¼ Truth
the inequality above might be:
Of course, we want our models to be consistent • ‘We were able to match the well test perme-
with the available data (from wells, seismic, and ability (kh) to within 10 % of the log-derived
dynamic data) and we hope they will give us a permeability data by applying the agreed cut-
good approximation of the truth, but too often the off and using a geometric average, and a
reservoir design engineer tries to force an artifi- power average with p ¼ 0.3 gave us an even
cial match which leads inevitably to great disap- better match to within 1 %.’
pointment. A common mistake is to try to • ‘The previous models had several serious
manipulate the data statistics to obtain an appar- errors and weaknesses, but this latest set of
ent match between the model and data. You may three models incorporates the latest data and
have heard versions of the following statements: captures the likely range of subsurface
• ‘We matched the well test permeability (kh) to behaviour.’
the log-derived permeability data by applying Figure 3.14 illustrates what the statistical
a cut-off and using a geometric average.’ objective of modelling should be. The available
76 3 The Property Model

Table 3.2 Statistics for a simple example of estimation in using the available data. To put it simply,
of the sand volume fraction, Vs, in a reservoir unit at variance refers to the spread of the data you
different stages of well data support
have (in front of you), while uncertainty refers
With 2 wells With 5 wells With 30 wells to some unknown variability beyond the infor-
Mean 38.5 36.2 37.4
mation at hand. From probability theory we can
σ 4.9 6.6 7.7
establish that ‘most’ values lie close to the mean.
SE 3.5 3.0 1.4
What we want to know is ‘how close’ – or how
Cv – 0.18 0.21
sure we are about the mean value. The funda-
N0 – 3 4
N 2 5 30
mental difficulty here is that the true (population)
mean is unknown and we have to employ the
theory of confidence intervals to give us an esti-
data is some limited subset of the true subsurface,
mate. Confidence limit theory is treated well in
and the model should extend from the data in
most books on statistics; Size (1987) has a good
order to make estimates of the true subsurface. In
introduction.
terms of set theory:
Chebyshev’s inequality gives us the theoretical
Data ∈ Model ∈ Truth basis (and mathematical proof) for quantifying
how many values lie within certain limits. For
Our models should be consistent with that data
example, for a Gaussian distribution 75 % of the
(in that they encompass it) but should aim to cap-
values are within the range of two standard
ture a wider range, approaching reality, using both
deviations from the mean. Stated simply
geological concepts and statistical methods. In
Chebyshev’s theory gives:
fact, as we shall see later (in this section and in
Sect. 3.4) bias in the data sample and upscaling 1
Pðjx  μj  κσ Þ  ð3:13Þ
transforms further complicate this picture whereby κ2
the data itself can be misleading. where κ is the number of standard deviations.
Table 3.2 illustrates this principle using the The standard error provides a simple measure
simple case of estimating the sand volume frac- of uncertainty. If we have a sample from a popu-
tion, Vs (or N/Gsand), at different stages in a field lation (assuming a normal distribution and statis-
development. We might quickly infer that the 30 tically independent values), then the standard
well case gives us the most correct estimate and error of the mean value, x, is the standard devia-
that the earlier 2 and 5 well cases are in error due to tion of the sample divided by the square root of
limited sample size. In fact, by applying the N-zero the sample size:
statistic (explained below) we can conclude that σs
the 5-well estimate is accurate to within 20 % of SEx ¼ pffiffiffi ð3:14Þ
n
the true mean, and that by the 30-well stage we still
lie within the range estimated at the 5-well stage. where σs is the standard deviation of the sample
In other words, it is better to proceed with a realis- and n is the sample size.
tic estimate of the range in Vs from the available The standard error can also be used to calculate
data than to assume that the data you have gives the confidence intervals. For example, the 95 % con-
“correct” value. In this case, Vs ¼ 36 %  7 % fidence interval is given by (x  SEx 1:96).
constitutes a good model at the 5-well stage in this The Coefficient of Variation, Cv, is a
field development. normalized measure of the dispersion of a proba-
bility distribution, or put simply a normalised
standard deviation:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3.3.2 Variance and Uncertainty Var ðpÞ
CV ¼ ð3:15Þ
Eð p Þ
There are a number of useful measures that can
guide the reservoir model practitioner in gaining where Var(p) and E(p) are the variance and
a realistic impression of the uncertainty involved expectation of the variable, p.
3.6 Vertical Permeability and Barriers 101

and kh throughout the procedure, and also gives a and care should be taken to minimise and record
sound basis for estimation of upscaled kh and kv. these errors. The TPM approach is generally
The upscaled kh (170 mD for the whole interval) more demanding but aims to minimize the
is significantly lower than the arithmetic average (inherent) upscaling errors by making estimates
kh because of the effects of sandstone connectiv- of the effective flow properties of the rock units
ity and the presence of shales and mudstone concerned. N/G ratios can be calculated at any
layers. A degree of validation that we have stage in the TPM modelling workflow.
derived a “reasonable estimate” for kh is found
in the observation that kh lies in the range
kgeometric to karithmetic (Fig. 3.36b). 3.6 Vertical Permeability
The main challenges of the Total Property and Barriers
Modelling approach are:
1. The approach requires some form of explicit 3.6.1 Introduction to kv/kh
upscaling, and upscaling always has some
associated errors; The ratio of vertical to horizontal permeability,
2. Where only log data are available (i.e. in the kv/kh, is an important, but often neglected, reser-
absence of fine-scale core data) some form voir modelling property. Too often, especially
of indirect estimate of the fine-scale sand/ when using the net-sand modelling method, a
mud ratios and rock properties is needed, value for the kv/kh ratio is assumed at the last
and this inevitably introduces additional ran- minute with little basis in reality. Figure 3.37
dom error in the estimation of N/Gres. captures a typical “history” for this parameter;
However, for challenging, heterogeneous or neglected or assumed ¼ 1 in the early stages
low-permeability reservoirs, these (generally then rapidly drops after unexpected barriers are
minor) errors are preferable to the errors encountered and finally rises again to a more
associated with the inappropriate simplifications plausible value late in the field life.
of the N/G approach. The problem of vertical permeability is also
In summary then, the widely used N/G further confounded because it is very difficult to
approach is simpler to apply and can be justified measure. Routine core plug analysis usually
for relatively good-quality reservoirs or gives some estimate of core-plug scale kv/kh but
situations where quick estimates are warranted. these data can be misleading due to severe under
The method tends to embed errors in the process sampling or biased sampling (discussed by
of re-scaling from well data to reservoir model, Corbett and Jensen 1992).

1
0.1
kv/kh ratio

0.01 ?
0.001
0.0001
0.00001
0.000001

Fig. 3.37 Typical


“history” of the kv/kh ratio
102 3 The Property Model

0.1

0.01
kv/kh
0.001

0.0001

0.00001
Hummocks Bar / Shoal Channels Heterolith
Median 0.231 0.239 0.153 0.132
Lower quartile 0.096 0.042 0.064 0.024
Min 0.0152 0.0009 7.68E -05 0.0001

Fig. 3.38 Statistics of measured kv/kh ratios from core plug pairs from an example reservoir interval

Figure 3.38 illustrates some typical core-plug situation, the geo-modeller needs to stochasti-
anisotropy data. For this example we know from cally simulate barriers and ensure they are
production data that the mean value is far too applied in the simulation model. Pervasive dis-
high (due to under sampling) and in fact the continuous thin shales and cements may also be
minimum observed plug kv/kh ratio gives a modelled as cell-value reduction factors (an
more realistic indication of the true values at effective kv/kh multiplier).
the reservoir scale. Figure 3.39 shows and example of barrier
A frequent problem with modelling or modelling for calcite cements in an example
estimating permeability anisotropy is confusion reservoir. The fine-scale barriers are first
between (or mixing the effects of) thin barriers modelled as geological objects and then assigned
and rock fabric anisotropy. The following two as vertical transmissibility values using single-
sections consider these two aspects separately. phase upscaling.
Before plunging into stochastic barrier
modelling, it is important to consider using well
3.6.2 Modelling Thin Barriers established empirical relationships that may save
a lot of time. Several previous studies have con-
Large extensive barriers are best handled explic- sidered the effects of random shales on a sand-
itly in the geological reservoir model: stone reservoir. Begg et al. (1989) proposed a
• Fault transmissibilities can be mapped onto to general estimator for the effective vertical perme-
cell boundaries; ability, kve, for a sandstone medium containing
• Extensive shales and cemented layers can be thin, discontinuous, impermeable mudstones,
modelled as objects and then transformed to based on effective medium theory and geometry
transmissibility multipliers onto to cell of ideal streamlines. They proposed:
boundaries.
k x ð1  V m Þ
Some packages allow simulation of sub- kVE ¼ ð3:30Þ
seismic faults as effective permeability reduction ðaz þ fd Þ2
factors within grid cells (see for example where
Manzocchi et al. 2002, or Lescoffit and Vm is the volume fraction of mudstone
Townsend 2005). Some modelling packages az is given by (ksv/ksh)1/2
offer the option to assign a sealing barrier ksh and ksv are the horizontal and vertical
between specified layers. For the more general permeability of the sandstone
3.6 Vertical Permeability and Barriers 103

Fig. 3.39 Example modelling of randomly distributed (b) Upscaled kv as vertical transmissibility multipliers
calcite cement barriers in an example reservoir (Reservoir (Modified from Ringrose et al. 2005, Petrol Geoscience,
is c. 80 m thick) (a) Fine-scale model of calcite barriers. Volume 11, # Geological Society of London [2005])

f is the barrier frequency 3.6.3 Modelling of Permeability


d is a mudstone dimension Anisotropy
(d ¼ Lm/2 for a 2D system with mean mudstone
length, Lm). Using advances in small-scale geological
This method is valid for low mudstone vol- modelling, it is now possible to accurately esti-
ume fractions and assumes thin, uncorrelated, mate kv/kh ratios for sandstone units. Ringrose
impermeable, discontinuous mudstone layers. et al (2003, 2005) and Nordahl et al (2005) have
Desbarats (1987) estimated effective perme- developed this approach for some common bed-
ability for a complete range of mudstone volume ding types found in tidal deltaic sandstone
fractions in 2D and 3D, using statistical models reservoirs (i.e. flaser, wavy and lenticular bed-
with spatial covariance and a range of ding). Their method gives a basis for general
anisotropies. For strongly stratified media, the estimation for facies-specific kv/kh ratios. Exam-
effective horizontal permeability, khe, was found ple results are shown in Figs. 3.40 and 3.41.
to approach the arithmetic mean, while kve was The method takes the following steps:
found to be closer to the geometric mean. Deutsch 1. Perform a large number of bedding
(1989) proposed using both power-average and simulations to understand the relationship
percolation models to approximate khe and kve between ksand, kmud and Vmud (simulations
for a binary permeability sandstone–mudstone are unconditioned to well data and can be
model on a regular 3D grid, and showed how done rapidly).
both the averaging power and the percolation 2. Input values for the small-scale models are the
exponents vary with the anisotropy ratio. typical values derived from measured core
Whatever the chosen method, it is important permeabilities.
to separate out the effects of thin barriers (or 3. A curve is fitted to the simulations to estimate
faults) from the more general rock permeability the kv or kv/kh ratio as a function of other
anisotropy (discussed below). modelled parameters: e.g. kh, Vmud, or ø.
104 3 The Property Model

Fig. 3.40 Example model of heterolithic flaser bedding (dark tones). Higher permeabilities indicated by hot
(left) with corresponding permeability model (right). colours (Modified from Ringrose et al. 2005 Petrol
Note the bi-directional sand lamina sets (green and Geoscience, Volume 11, # Geological Society of
yellow laminae) and the partially preserved mud drapes London [2005])

Fig. 3.41 Results of sub- Lenticular \ Wavy \ Flaser


metre-scale simulation of x
heterolithic bedding in tidal
z
deltaic reservoirs. Effective
permeability simulation
100
results are for the constant
Kx Kz Ky
petrophysical properties
case (i.e. sandstone and 80
Permeability (md)

A B C
mudstone have constant
permeability). Observed 60
effective permeability is
compared to bedding styles 40
and the critical points A, B
and C. A is the percolation
20
threshold for kx, and C is
the percolation threshold
for kz, while B is the 0
100
theoretical percolation Kx Kz Ky
threshold for a simple 3D
Permeability (md)

system. Thin lines are the 10


arithmetic and harmonic
averages
1

0.1

0.01
0 0.2 0.4 0.6 0.8 1
Vsand
3.7 Saturation Modelling 105

The following function was found to capture


the characteristic vertical permeability of this 3.7 Saturation Modelling
system (Ringrose et al. 2003):
3.7.1 Capillary Pressure
 Vm
kmud Vmc
kv ¼ ksand ð3:31Þ An important interface between the static and
ksand
dynamic models is the definition of initial water
where Vmc is the critical mudstone volume frac- saturation. There are numerous approaches to this
tion (or percolation threshold). problem, and in many challenging situations anal-
This formula is essentially a re-scaled geo- ysis and modelling of fluid saturations requires
metric average constrained by the percolation specialist knowledge in the petrophysics and res-
threshold. This is consistent with the previous ervoir engineering disciplines. Here we introduce
findings by Desberats (1987) and Deutsch the important underlying concepts that will enable
(1989) who observed that the geometric average the initial saturation model to be linked to the
was close to simulated kv for random shale geological model and its uncertainties.
systems, and also noted percolation behaviour The initial saturation model is usually based
in such systems. This equation captures the per- on the assumption of capillary equilibrium with
colation behaviour (the percolation threshold is saturations defined by the capillary pressure
estimated for the geometry of a specific deposi- curve. We recall the basic definition for capillary
tional system or facies), while still employing a pressure:
general average function that can be easily
Pc ¼ Pnon‐wetting phase
applied in reservoir simulation.
 Pwetting‐phase ½Pc ¼ f ðSÞ ð3:32Þ
The method has been applied to a full-field
study by Elfenbein et al. (2005) and compared to The most basic form for this equation is given
well-test estimates of anisotropy (Table 3.5). The by: pffiffiffiffiffiffiffiffi
comparison showed a very good match in the Pc ¼ ASwn b ϕ=k ð3:33Þ
Garn 4 Unit but a poorer match in the Garn 1–3
Units. This can be explained by the fact that the That is, capillary pressure is a function of the
lower Garn 1–3 Units have abundant calcite wetting phase saturation and the rock properties,
cements (which were modelled in the larger- summarized by ϕ and k. The exponent b is
scale full-field geomodel), illustrating the impor- related to the pore size distribution of the rock.
tance of understanding both the thin large-scale Note the use of the normalised water saturation:
barriers and the inherent sandstone anisotropy
Swn ¼ ðSw  Swi Þ=ðSwor  Swi Þ ð3:34Þ
(related to the facies and bedding architecture).

Table 3.5 Comparison of simulated kv/kh ratios with well test estimates from the case study by Elfenbein et al. (2005)
Modelled kv/kh: Modelled kv/kh:
Geometric average Geometric average Well test kv/kh:
Reservoir unit of simulation model of well test volume Analytical estimate Comments
Tyrihans South, well test in well 6407/1-2
Garn 4 0.031 0.043 <0.05 Test of Garn 4 interval
Garn 3 0.11 Producing interval uncertain
Garn 2 0.22 Complex two-phase flow
Garn 1 0.11
Tyrihans North, well test in well 6407/1-3
Garn 4 0.025
Garn 3 0.123 0.19 0.055 Test of Garn 1 to 3 interval
Garn 2 0.24 Analytical gas cap
Garn 1 0.12 Partial penetration model
106 3 The Property Model

Here Pc is defined by the fluid buoyancy term,


Silty Δ(ρ)gh, where h is the height above the free
sandstone water level. This equation gives a useful basis
1 for forward modelling water saturation, given
some known rock and fluid properties.
Clean uniform
J(Sw) sandstone For practical purposes we often want to esti-
mate the Sw function from well log data. There
are again several approaches to this
(Worthington 2001 gives a review), but the sim-
plest is the power law function which has the
same form as the J-function:
0
0 Sw 1 Sw ¼ C:hd ð3:38Þ
Fig. 3.42 Example capillary pressure J-functions A significant issue in reservoir modelling is
how the apparent (and true) saturation height
function is affected by averaging of well data
We can expand the Pc equation to include the and/or upscaling of the fine-scale geological
fluid properties: model data.
pffiffiffiffiffiffiffiffi To illustrate these effects in the reservoir
Pc ðSw Þ ¼ σ cos θ J ðSw Þ ϕ=k ð3:35Þ
model, we take a simple case. We must first
where define the free water level (FWL) – the fluid
σ ¼ interfacial tension water interface in the absence of rock pores, i.e.
θ ¼ interfacial contact angle resulting only from fluid forces (buoyancy and
J(Sw) ¼ Leverett J-junction. hydrodynamic pressure gradients). The effect of
Rearranging this we obtain the J-function: rock pores is to introduce another factor (capil-

Pc ðSw Þ k 1=2 lary forces) on the oil-water distribution, so that
J ð Sw Þ ¼ ð3:36Þ
σ cos θ ϕ the oil-water contact is different from the free
Figure 3.42 shows two example J-functions water level.
for contrasting rock types. A simple model for this behaviour is given by
To put this more simply, we could measure and the following saturation-height function:
model any number of capillary pressure curves, h pffiffiffiffiffiffiffiffii2=3
Pc ¼ f(S). However, the J-function method allows Sw ¼ Swi þ ð1  Swi Þ 0:1h k=ϕ ð3:39Þ
a number of similar functions to be normalized
Figure 3.43 shows example curves, based on
with respect to the rock and fluid properties and
this function, and illustrates how at least 10 m
plotted with a single common curve.
variation in oil-water contact can occur due to
changes in pore throat size. In general, for a high
porosity/permeability rock OWC  FWL. How-
3.7.2 Saturation Height Functions ever, for low permeability or heterogeneous
reservoirs the fluid contact will vary considerably
There are a number of ways of plotting the Pc ¼ as a function of rock properties, and
f(S) function to indicate how saturation varies OWC 6¼ FWL.
with height in the reservoir. The following equa- Further difficulties in interpretation of these
tion is a general form of the Pc equation, includ- functions come with upscaling or averaging
ing all the key rock and fluid terms. saturations from heterogeneous systems. For
 example, suppose you had a thinly-bedded reser-
ðρ  ρo Þgh k 1=2
Swn b ¼ w ð3:37Þ voir comprising alternating rock types 2 and 4
σ cos θ ϕ
(Fig. 3.43), then the average saturation-height
3.7 Saturation Modelling 107

Fig. 3.43 Example saturation-height functions for the listed input parameters, illustrating how an apparent change in
oil-water contact may be caused by rock property variations between wells

Table 3.6 Selected examples of tilted oil-water contacts


Field, Location Tilt of OWC (m/km) References
South Glenrock, Wy. USA <95 Dahlberg (1995)
Norman Wells, NWT, Canada 75 Dahlberg (1995)
Tin-Fouye, Algeria 10 Dahlberg (1995)
Weyburn, Sask., Canada 10 Dahlberg (1995)
Kraka, North Sea (Denmark) 10 Thomasen and Jacobsen (1994)
Billings Nose, N.Da., USA 5 Berg et al. (1994)
Knutson, N.Da., USA 3 Berg et al. (1994)

function (detected by a logging tool) would be Africa and North America, there are numerous
close to curve 3. That is, the average Sw well-documented examples. These represent
corresponds to the average k/ϕ. However, if the continental basins with appreciable levels of
thin beds were composed of an unknown random topographically driven groundwater flow, or
mix of rocks types 1, 2, 3, 4 and 5 then it would hydrodynamic gradients. Dahlberg (1995)
clearly be very difficult to infer the correct rela- provides a fairly comprehensive study of the
tionship between Sw, k and ϕ. evidence for, and interpretation of, tilted oil
water contacts. Berg et al. (1994) give good
documentation of some examples from Dakota,
3.7.3 Tilted Oil-Water Contacts USA. In the offshore continental shelf petroleum
provinces, such as offshore NW Europe the cases
Depending on which part of the world the petro- are fewer, but still evident. Table 3.6 lists a range
leum geologist is working, tilted oil-water of examples.
contacts are either part of accepted or disputed Here, we are concerned with the implications
folk law. In parts of the Middle East, North that tilted hydrocarbon-water contacts might
108 3 The Property Model

Δx
ΔHw

Hydrocarbon Δz
Aquifer
flow

Fig. 3.44 Terms defining a tilted oil-water contact (Redrawn from Dahlberg 1995 (Fig. 12.5), Springer-Verlag, New
York, with kind permission from Springer Science and Business Media B.V.)

have for a dynamic modelling of petroleum (2012), who include the terms for the effective
accumulations. For simplicity we mainly con- permeability in the aquifer, kaq, and reservoir,
sider oil-water contacts, but the theory applies kres, to derive a relationship between the
to any hydrocarbon: gas, condensate or oil. The hydrocarbon-water interface and the hydrody-
main principle governing this phenomenon is namic pressure gradient in terms of steady-state
potentiometric head. If an aquifer contains flow:
flowing water driven by some pressure gradient  
ðΔz=Δx Þ ¼ kres =kaq Δρg :ðΔHw =ΔxÞ ð3:41Þ
(Fig. 3.44), then this pressure gradient causes a
slope in the petroleum-water interface of any As can be seen from Table 3.6, the actual
accumulation within that aquifer, defined by value of the tilted oil-water contact can be quite
Hubbert (1953) as: small (most documented examples are around
10 m/km), so that uncertainties in detection
Δz=Δx  ðρw =ρw  ρo Þ:ðΔHw =ΔxÞ ð3:40Þ
become important. There are many situations
where which can give an apparent tilt in oil-water con-
ρw  ρo ¼ density of water and petroleum tact, including:
Δz/Δx ¼ slope of the hydrocarbon-water • Undetected faults (usually the first explana-
interface tion to be proposed) or stratigraphic
ΔHw/Δx ¼ potentiometric surface in aquifer boundaries;
• Variations in reservoir properties – systematic
The greater the difference in fluid density (i.e. changes in pore throat size across a field can
the lighter the petroleum), the smaller the tilt of lead to a variation in the oil-water contact of
the fluid contact. It is important to differentiate 5 m or more (Fig. 3.43);
the free-water level (FWL) from the oil-water • Misinterpretation of paleo-oil-water contacts
contact (OWC). Where the capillary pressures (marked by residual oil stains or tar mats) as
are significant (due to small pores), the differ- present-day contacts;
ence between FWL and OWC can be significant • Errors in deviation data for well trajectories.
(Fig. 3.43). In Eq. (3.40), the Δz/Δx term relates Thus, proof of the presence of a tilted oil-
to the FWL (and only approximately to the water contact requires either multiple well data
OWC). A more comprehensive treatment of this explained by a common inclined surface
topic is given by Muggeridge and Mahmode (Fig. 3.45) or multiple data types explained
3.7 Saturation Modelling 109

coherently in terms of a common hydrodynamic 3.7.3.1 Kraka Field Example


model (as in the Kraka field example discussed This small chalk reservoir in the Danish sector of
below). Possible hydrodynamic aquifer influence the North Sea provides an interesting account of
on a static (i.e. passive) petroleum accumulation the phenomenon of tilted oil-water contacts and
must also be considered alongside the concepts their interpretation. The subtle nature of the tilt
of a dynamic petroleum accumulation (e.g. on- and the use of multiple data sources to confirm an
going migration or leakage) or pressure initially doubtful interpretation are very informa-
transients in the aquifer. tive. A study of the field by Jørgensen and
Andersen (1991) included some initial
observations on a tilted oil-water contact, and a
tentative argument that it was due to tectonic
tilting during the Tertiary. A subsequent study
by Thomasen and Jacobsen (1994), give a
detailed description and a more thorough basis
for interpretation a 0.6 dip in both free water
level and oil-water contact (Fig. 3.46).
Their main observations were:
• Repeat Formation Tester (RFT) data from
three wells indicated a free-water level
(interpreted from the change in slope of
water and oil zones) falling by about 70 m
over a 2 km distance (Fig. 3.47).
• Due to the heterogeneous and fractured nature
of the chalk reservoir zone, logs from seven
Fig. 3.45 Map of the Cairo Pool oilfield, Arkansas wells show highly variable saturations
showing a hydrodynamic offset of an oil accumulation
(After Dahlberg 1995). Contours are 20 foot intervals;
(Fig. 3.48). These were interpreted by best-
black dots ¼ wells with oil in the reservoir interval, fit capillary pressure saturation functions.
open circles ¼ wells with water in the reservoir interval Difficulties in fitting a function assuming a
(Redrawn from Dahlberg 1995 (Fig. 12.5), Springer- horizontal free-water level were resolved by
Verlag, New York, with kind permission from Springer
Science and Business Media B.V.)
fitting functions to individual wells and then
identifying the implied tilt in free water level.

5600
A-7C well trajectory
Depth (feet)

5800
GOC

6000

6200

1 km

Fig. 3.46 Cross-section through the Kraka field (From Society of Petroleum Engineers Inc., reproduced with
Thomasen and Jacobsen 1994) showing interpreted fluid permission of SPE. Further reproduction prohibited with-
contacts and horizontal well to exploit down-dip reserves out permission)
(Redrawn from Thomasen and Jacobsen 1994, #1994,
110 3 The Property Model

Top Chalk (Danian D1)

Gas
5850 GOC

Oil

Danian D2)
5900

Water

Depth (ft TVD SS)


5950

6000 Top Maastrichtian

6050 OWC

Volume Fraction
Fig. 3.47 RFT data for three wells for the Kraka field
(From Thomasen and Jacobsen 1994) (Redrawn from Fig. 3.48 Type log for the Kraka field illustrating the
Thomasen and Jacobsen 1994, #1994, Society of Petro- variable oil saturations and the thick transition zone
leum Engineers Inc., reproduced with permission of SPE. (Redrawn from Thomasen and Jacobsen 1994, #1994,
Further reproduction prohibited without permission) Society of Petroleum Engineers Inc., reproduced with
permission of SPE. Further reproduction prohibited with-
out permission)

The slope of the free-water level inferred from


this was found to be close to the RFT pressure 3.8 Summary
data model.
• An intra-reservoir seismic reflection We have covered a range of issues related to
interpreted as the matrix oil-water contact petrophysical property modelling of oil and gas
was mapped around the field and extrapolated reservoirs. The theoretical principles that under-
to its intersection with top reservoir, and was lie the modelling “buttons” and workflows in
again found to be in good agreement with the geological reservoir modelling packages have
saturation model. been discussed along with many of the practical
• The orientation of the inferred hydrodynamic issues that govern the choice of parameters.
gradient (towards the SE) was found to be in To summarise this chapter, we offer a check
agreement with regional gradients from inter- list of key questions to ask before proceeding
field pressure variations. with your property modelling task:
This integrated interpretation had a significant 1. Have you agreed with your colleagues across
economic benefit in terms of the appropriate disciplines (geoscience, petrophysics and res-
placement of horizontal wells in the thicker part ervoir engineering):
of the accumulation, and in the estimation of • The key geological issues you need to
inter-well permeability in this fairly marginal address – rock heterogeneity, sedimentary
field development. barriers, faults, etc.
References 111

5. Have you run sensitivities to check important


assumptions?
6. Have you considered the effects of possible
un-detected flow barriers in the system?
A final word about the future of property
modelling – if we are looking for fit-for-purpose
models for interpreting petrophysical well data,
then we are probably talking about high-
resolution near-wellbore models (Fig. 3.49).
These models could be very detailed or could
be just a simple equation. Either way they need
to be focussed on the scale of rock property
variation – the subject of the next chapter.

References
Abbaszadeh M, Fujii H, Fujimoto F (1996) Permeability
prediction by hydraulic flow units – theory and
applications. SPE Form Eval 11(4):263–271
Begg SH, Carter RR, Dranfield P (1989) Assigning effec-
tive values to simulator gridblock parameters for het-
erogeneous reservoirs. SPE Reserv Eng 1989:455–463
Berg RR, DeMis WD, Mitsdarffer AR (1994) Hydrody-
namic effects on Mission Canyon (Mississippian) oil
accumulations, Billings Nose area, North Dakota.
AAPG Bull 78(4):501–518
Bierkins MFP (1996) Modeling hydraulic conductivity of
a complex confining layer at various spatial scales.
Water Resour Res 32(8):2369–2382
Bourbie T, Zinszner B (1985) Hydraulic and acoustic
properties as a function of porosity in Fontainebleau
sandstone. J Geophys Res 90(B13):11524–11532
Box GEP, Cox DR (1964) An analysis of transformations.
J R Stat Soc Series B: 211–243, discussion 244–252
Brandsæter I, McIlroy D, Lia O, Ringrose PS (2005)
Reservoir modelling of the Lajas outcrop (Argentina)
Fig. 3.49 Near-wellbore porosity model (1 m2 10 m) to constrain tidal reservoirs of the Haltenbanken
(Norway). Petrol Geosci 11:37–46
Bryant S, Blunt MJ (1992) Prediction of relative permeabil-
• A consistent method for handling net-to- ity in simple porous media. Phys Rev A 46:2004–2011
Buland A, Kolbjornsen O, Omre H (2003) Rapid spatially
gross (N/G) and cut-off values? coupled AVO inversion in the fourier domain. Geo-
2. Is your petrophysical data representative of physics 68(1):824–836
the rock unit (sampling problems, tails of Cardwell WT, Parsons RL (1945) Average permeabilities
distributions), and if not how will you address of heterogeneous oil sands. Trans Am Inst Mining Met
Pet Eng 160:34–42
that uncertainty? Corbett PWM, Jensen JL (1992) Estimating the mean
3. Have you used appropriate averaging and/or permeability: how many measurements do you need?
upscaling methods? First Break 10:89–94
4. Is the model output consistent with data Corbett PWM, Ringrose PS, Jensen JL, Sorbie KS (1992)
Laminated clastic reservoirs: the interplay of capillary
input? Compare the statistics of input and pressure and sedimentary architecture. SPE paper
output distributions. The variance may be as 24699, presented at the SPE annual technical confer-
important as the mean. ence, Washington, DC
112 3 The Property Model

Cosentino L (2001) Integrated reservoir studies. Editions Jørgensen LN, Andersen PM (1991) Integrated study
Technip, Paris, 310 pp of the Kraka Field. SPE paper 23082, presented at
Dahlberg EC (1995) Applied hydrodynamics in petro- the offshore Europe conference, Aberdeen, 3–6 Sept
leum exploration, 2nd edn. Springer, New York 1991
Davis JC (2003) Statistics and data analysis in geology, Journel AG, Alabert FG (1990) New method for reservoir
3rd edn. Wiley, New York, 638 pp mapping. J Petrol Technol 42.02:212–218
de Marsilly G (1986) Quantitative hydrogeology. Aca- Journel AG, Deutsch CV (1997) Rank order geostatistics: a
demic, San Diego proposal for a unique coding and common processing
Delfiner P (2007) Three statistical pitfalls of phi-k of diverse data. Geostat Wollongong 96:174–187
transforms. SPE Reserv Eval Eng 10:609–617 Journel AG, Deutsch CV, Desbarats AJ (1986) Power
Desberats AJ (1987) Numerical estimation of effective averaging for block effective permeability. SPE
permeability in sand-shale formations. Water Resour paper 15128, presented at SPE California regional
Res 23(2):273–286 meeting, Oakland, California, 2–4 April
Deutsch C (1989) Calculating effective absolute perme- Kendall M, Staurt A (1977) The advanced theory of
ability in sandstone/shale sequences. SPE Form Eval statistics, vol. 1: distribution theory, 4th edn.
4:343–348 Macmillan, New York
Deutsch CV (2002) Geostatistical reservoir modeling. Lescoffit G, Townsend C (2005) Quantifying the impact
Oxford University Press, Oxford, 376 pp of fault modeling parameters on production
Deutsch CV, Journel AG (1992) Geostatistical software forecasting for clastic reservoirs. In: Evaluating fault
library and user’s guide, vol 1996. Oxford University and cap rock seals, AAPG special volume Hedberg
Press, New York series, no. 2. American Association of Petroleum
Doyen PM (2007) Seismic reservoir characterisation. Geologists, Tulsa, pp 137–149
EAGE Publications, Houten Leuangthong O, Khan KD, Deutsch CV (2011) Solved
Durlofsky LJ (1991) Numerical calculations of equivalent problems in geostatistics. Wiley, New York
grid block permeability tensors for heterogeneous Manzocchi T, Heath AE, Walsh JJ, Childs C (2002) The
porous media. Water Resour Res 27(5):699–708 representation of two-phase fault-rock properties in
Elfenbein C, Husby Ø, Ringrose PS (2005) Geologically- flow simulation models. Petrol Geosci 8:119–132
based estimation of kv/kh ratios: an example from the Matheron G (1967) Éléments pour une théorie des
Garn Formation, Tyrihans Field, Mid-Norway. In: milieux poreux. Masson and Cie, Paris
Dore AG, Vining B (eds) Petroleum geology: North- McIlroy D, Flint S, Howell JA, Timms N (2005) Sedimen-
West Europe and global perspectives. Proceedings of tology of the tide-dominated Jurassic Lajas Formation,
the 6th petroleum geology conference. The Geological Neuquen Basin, Argentina. Geol Soc Lond Spec Publ
Society, London 252:83–107
Goggin DJ, Chandler MA, Kocurek G, Lake LW (1988) Mourzenko VV, Thovert JF, Adler PM (1995) Permeabil-
Patterns of permeability variation in eolian deposits: ity of a single fracture; validity of the Reynolds equa-
page sandstone (Jurassic), N.E. Arizona. SPE Form tion. J de Phys II 5(3):465–482
Eval 3(2):297–306 Muggeridge A, Mahmode H (2012) Hydrodynamic aqui-
Gutjahr AL, Gelhar LW, Bakr AA, MacMillan JR (1978) fer or reservoir compartmentalization? AAPG Bull 96
Stochastic analysis of spatial variability in subsurface (2):315–336
flows 2. Evaluation and application. Water Resour Res Muskat M (1937) The flow of homogeneous fluids
14(5):953–959 through porous media. McGraw-Hill, New York
Hohn ME (1999) Geostatistics and petroleum geology, (Reprinted by the SPE and Springer 1982)
2nd edn. Kluwer, Dordrecht Nair KN, Kolbjørnsen O, Skorstad A (2012) Seismic inver-
Howson C, Urbach P (1991) Bayesian reasoning in sci- sion and its applications in reservoir characterization.
ence. Nature 350:371–374 First Break 30:83–86
Hubbert MK (1953) Entrapment of petroleum under hydro- Nelson RA (2001) Geologic analysis of naturally fractured
dynamic conditions. AAPG Bull 37(8):1954–2026 reservoirs, 2nd edn. Butterworth-Heinemann, Boston
Hurst A, Rosvoll KJ (1991) Permeability variations in Nordahl K, Ringrose PS, Wen R (2005) Petrophysical
sandstones and their relationship to sedimentary characterisation of a heterolithic tidal reservoir interval
structures. In: Lake LW, Carroll HB Jr, Wesson TC using a process-based modelling tool. Petrol Geosci
(eds) Reservoir characterisation II. Academic, San 11:17–28
Diego, pp 166–196 Olea RA (ed) (1991) Geostatistical glossary and multilin-
Isaaks EH, Srivastava RM (1989) Introduction to applied gual dictionary, IAMG studies in mathematical geology
geostatistics. Oxford University Press, Oxford no. 3. Oxford University Press, Oxford
Jensen JL, Corbett PWM, Pickup GE, Ringrose PS (1995) Pickup GE, Sorbie KS (1996) The scaleup of two-phase
Permeability semivariograms, geological structure flow in porous media using phase permeability
and flow performance. Math Geol 28(4):419–435 tensors. SPE J 1:369–381
Jensen JL, Lake LW, Corbett PWM, Goggin DJ (2000) Pickup GE, Ringrose PS, Jensen JL, Sorbie KS (1994)
Statistics for petroleum engineers and geoscientists, Permeability tensors for sedimentary structures. Math
2nd edn. Elsevier, Amsterdam Geol 26:227–250
References 113

Pickup GE, Ringrose PS, Corbett PWM, Jensen JL, Size WB (ed) (1987) Use and abuse of statistical methods
Sorbie KS (1995) Geology, geometry and effective in the earth sciences, IAMG studies in mathematical
flow. Petrol Geosci 1:37–42 geology, no. 1. Oxford University Press, Oxford
Renard P, de Marsily G (1997) Calculating equivalent Soares A (2001) Direct sequential simulation and
permeability: a review. Adv Water Resour cosimulation. Math Geol 33(8):911–926
20:253–278 Thomasen JB, Jacobsen NL (1994) Dipping fluid contacts in
Ringrose PS (2008) Total-property modeling: dispelling the Kraka Field, Danish North Sea. SPE paper 28435,
the net-to-gross myth. SPE Reserv Eval Eng presented at the 69th SPE annual technical conference
11:866–873 and exhibition, New Orleans, LA, USA, 25–28 Sept 1994
Ringrose PS, Pickup GE, Jensen JL, Forrester M (1999) Weber KJ, van Geuns LC (1990) Framework for
The Ardross reservoir gridblock analogue: sedimen- constructing clastic reservoir simulation models. J
tology, statistical representivity and flow upscaling. Petrol Tech 42:1248–1297
In: Schatzinger R, Jordan J (eds) Reservoir characteri- White CD, Horne RN (1987) Computing absolute trans-
zation – recent advances, AAPG memoir no. 71., pp missibility in the presence of fine-scale heterogeneity.
265–276 SPE paper 16011, presented at the 9th SPE symposium
Ringrose PS, Skjetne E, Elfeinbein C (2003) Permeability of reservoir simulation, San Antonio, TX, 1–4 Feb 1987
estimation functions based on forward modeling of Witherspoon PA, Wang JSY, Iwai K, Gale JE (1980)
sedimentary heterogeneity. SPE 84275, presented at Validity of cubic law for fluid flow in a deformable
the SPE annual conference, Denver, CO, USA, 5–8 rock fracture. Water Resour Res 16(6):1016–1024
Oct 2003 Worthington PF (2001) Scale effects on the application of
Ringrose PS, Nordahl K, Wen R (2005) Vertical perme- saturation-height functions to reservoir petrofacies
ability estimation in heterolithic tidal deltaic units. SPE Reserv Eval Eng 4(5):430–436
sandstones. Petrol Geosci 11:29–36 Worthington PF, Cosentino L (2005) The role of cut-offs
Shuey RT (1985) A simplification of the Zoeppritz in integrated reservoir studies (SPE paper 84387). SPE
equations. Geophysics 50(4):609–614 Reserv Eval Eng 8(4):276–290
Upscaling Flow Properties
4

Abstract
To upscale flow properties means to estimate large-scale flow behaviour
from smaller-scale measurements. Typically, we start with a few
measurements of rock samples (lengthscale ~3 cm) and some records of
flow rates and pressures in test wells (~100 m). Our challenge is to
estimate how the whole reservoir will flow (~1 km).
Flow properties of rocks vary enormously over a wide range of length-
scales, and estimating upscaled flow properties can be quite a challenge.
Unfortunately, many reservoir modellers choose to overlook this problem
and blindly hope that a few measurements will correctly represent the
whole reservoir. The aim of this chapter is to help make intelligent
estimates of large-scale flow properties. In the words of Albert Einstein:
Two things are infinite: the universe and human stupidity; and I’m not sure
about the universe.

P. Ringrose and M. Bentley, Reservoir Model Design, DOI 10.1007/978-94-007-5497-3_4, 115


# Springer Science+Business Media B.V. 2015
116 4 Upscaling Flow Properties

Upscaling – from pore to field, and beyond . . .

4.1 Multi-scale Flow Modelling In multi-scale geological modelling, the


essence is that geological concepts are used to
This chapter concerns the implementation of make the transition from smaller-scale
multi-scale flow modelling for oil and gas reser- measurements to larger-scale estimates (models)
voir studies. Multi-scale flow modelling is of reservoir properties or behaviour (Fig. 4.1).
defined here as any method which attempts to Geological modelling in itself is an art form
explicitly represent the flow properties at more- requiring some intimate knowledge of the geo-
than-one scale within a reservoir. We may, for logical system – typically involving Picasso-type
example, have (a) an estimate of flow properties geologists (Fig. 4.2) with an interest in detail. For
around a single well in a specific flow unit (or upscaling we require representative geological
reservoir interval) and (b) a rationale for using models in which the geological elements (e.g.
this estimate to calculate the flow properties in layers of sandstone, siltstone, mudstone and
the whole reservoir. This rationale could simply limestone) are represented as properties relevant
be some multiplication factors transforming the for fluid modelling – porosity, permeability,
single-well flow property to the reservoir scale, capillary pressure functions, etc.
or might involve a 3-dimensional array (or grid) This process inevitably involves some simpli-
of values drawn from statistical population fication of the intricate variability of rock archi-
(which includes the single-well flow property). tecture, as we aim to group the rock elements into
4.1 Multi-scale Flow Modelling 117

Conventional reservoir model


with c.100x100x10m cells

dm-sized
borehole
sample

Geological concepts and models

Scale transition

Fig. 4.1 Scale transition in reservoir modelling and the role of geological concepts

Fig. 4.2 The art of geological modelling: “Picasso-type Picasso, Violins and Grapes, oil on canvas (1912) and
geologists” aim to represent fine detail in their art work Mark Rothko, No. 10, 1950. Oil on canvas,
while “Rothko-type geologists” aim to capture only the 229.2  146.4 cm, reproduced with permission DIGITAL
representative flow units as essential colours (Pablo IMAGE # The Museum of Modern Art/Scala, Florence)

flow units with similar properties. In the art anal- The process of transferring information
ogy this process is more like the work of Mark between scales is referred to as upscaling or,
Rothko, where broad bands of colour capture the more generally, re-scaling. Upscaling involves
essence of the object or concept being described some form of numerical or analytical method
(Fig. 4.2). for estimating effective or equivalent flow
118 4 Upscaling Flow Properties

properties at a larger scale given some set of finer The factors involved in these scale transitions
scale rock properties. Upscaling methods for sin- are enormous; certainly around 109 as we go
gle and multiphase flow are reviewed in detail by from the rock pore to the full-field reservoir
Renard and de Marsily (1997), Barker and model (Table 4.1), and important scale markers
Thibeau (1997), Ekran and Aasen (2000) and involved in reservoir modelling are best
Pickup et al. (2005). We will review the methods illustrated on a logarithmic scale (Fig. 4.4).
involved and establish the principles which guide Despite these large scale transitions, most flow
the flow upscaling process. The term downscal- processes average out the local variations – so
ing has also been used (Doyen 2007) to mean the that what we are looking for is the correct aver-
process by which smaller-scale properties are age flow behaviour at the larger scales. How we
estimated from a larger–scale property. This is do this is the rationale for this chapter.
most commonly done in the context of seismic Flow simulation of detailed reservoir models
data where, for example, a porosity value is a fairly demanding exercise, involving many
estimated from seismic impedance is used to mathematical tools for creating and handling
constrain the porosity values of thin layers flow grids and calculating the flows and
below the resolution of the seismic wavelet. In pressures between the grid cells. The mathemat-
more general terms, if we know all the fine-scale ics of flow simulation is beyond the scope of this
properties then the upscaled property can be book, and will be treated only in an introductory
estimated uniquely. Conversely, if we know sense. Mallet (2008) gives a recent review of
only the large-scale property value then there the processes involved in the creation of numer-
are many alternative fine-scale property models ical rock models and their use in flow simula-
that could be consistent with the upscaled tion. King and Mansfield (1999) also give a
property. fairly comprehensive discussion of flow simula-
We will develop the argument that upscaling tion of geological reservoir models, in terms of
is essential in reservoir modelling – whether managing and handling the grid and associated
implicit or explicit. There is no such thing as flow terms (transmissibility factors). In this
the correct value for the permeability of a given chapter, we will take as our starting point the
hydraulic flow unit. The relevant permeability existence of a numerical rock model, created by
value depends on length-scale, the boundary some set of recipes in a geological modelling
conditions and flow process. Efforts to define toolkit, and will focus on the methods involved
the diagnostic characteristics for hydraulic flow for performing multi-scale upscaling. Before we
units (HFU) (e.g. Abbaszadeh et al. 1996) pro- do that we need to introduce, or recapitulate,
vide valuable approaches to petrophysical data some of the basic theory for multiphase fluid
analysis, but HFUs should always be related to a flow.
representative elementary volume (REV). As we
will show it is not always simple to define the
REV, and when flow process are brought into 4.2 Multi-phase Flow
play different REVs may apply to different flow
processes. Hydraulic flow units are themselves 4.2.1 Two-Phase Flow Equations
multi-scale.
The framework we will use for upscaling In Chap. 3 we introduced the concept of perme-
involves a series of steps where smaller-scale ability and the theoretical basis for estimating
models are nested within larger scale models. effective permeability using averages and
These steps essentially involve models or numerical recipes. This introduced us to
concepts at the pore-scale, geological concepts upscaling for single-phase flow properties. Here
and models at the field-scale and reservoir we extend this by looking at two-phase flow and
simulations (Fig. 4.3). the upscaling of multi-phase flow properties.
4.2 Multi-phase Flow 119

Fig. 4.3 Reservoir models


at different scales (Statoil
image archives, # Statoil
ASA, reproduced with
permission)

Table 4.1 Typical dimensions for important volumes used in multi-scale reservoir modelling
Fraction of reservoir
X (m) Y(m) Z(m) Volume (m3) Cubic root (m) volume
Pore-scale model 5  105 50  105 50  105 1.25  1013 0.00005 0.00000005
Core plug sample 0.025 0.025 0.025 0.000031 0.031 0.00003
Well test volume 400 300 10 1,200,000 106 0.1
Reservoir model 8,000 4,000 40 1,280,000,000 1,086 1

For a fuller treatment of multi-phase flow The first essential concept in multiphase flow
theory applied to oil and gas reservoir systems is the principal of mass balance. Any fluid which
refer to reservoir engineering textbooks (e.g. flows into a grid cell (mass accumulation) over a
Chierici 1994; Dake 2001; Towler 2002). particular interval of time must be equal to the
A more geologically-based introduction to mass of fluids which have flowed out. This prin-
multi-phase flow in structured sedimentary ciple may be rather trivial for single phase flow,
media is given by Ringrose et al. (1993) and but becomes more critical for multiphase flow,
Ringrose and Corbett (1994). where different fluids may have different
120 4 Upscaling Flow Properties

Gamma-log resolution

Fluvial channel width


Image-log resolution

Reservoir thickness
Medium sand grain

Seismic resolution
Lamina thickness
Pore-scale model

Bedset thickness

Well test interval

Reservoir width
Core plug
10-5 10-4 10-3 0.01 0.1 1 10 100 1000
Metres

Fig. 4.4 Important length scales involved in reservoir modelling

densities, viscosities, and permeabilities. What where:


goes in must be balanced by what comes out, o and w refer to the oil and water phases,
and for a complex set of flow equations the krw and kro are the relative permeabilities of each
zero-sum constraint for each grid cell is essential. phase,
Fluid flow in porous media is represented by μ and ρ are fluid viscosity and density,
Darcy’s Law (Sect. 4.3.2) which relates the fluid Pc is the capillary pressure,
velocity, u, to the pressure gradient and two ∇Po is the gradient of pressure for the oil phase
terms representing the rock and the fluid: This set of equations is non-linear as the krw,
kro and Pc terms are all functions of phase satura-
u ¼ k=μ:∇ðP þ ρgzÞ ð4:1Þ tion, Sw, which is itself controlled by the flow
rates. Thus, in order to solve these equations for a
The pressure term comprises an imposed pres- given set of initial and boundary conditions,
sure gradient, ∇(P), and a pressure gradient due numerical codes (reservoir simulators) are used,
to gravity, ∇(ρgz). In Cartesian coordinates the in which saturation-dependent functions for krw,
gradient of pressure, ∇P, is resolved as: kro and Pc are given as input, and an iterative
numerical recipe is used to estimate saturation
dP dP dP and pressure. Figure 4.5 shows a typical set of
∇P ¼ þ þ ð4:2Þ
dx dy dz oil-water relative permeability curves with the
endpoint terminology.
The rock (or the permeable medium) is Note that the total fluid mobility is <1 (mobil-
represented by the permeability tensor, k, and ity is the permeability/viscosity ratio for the
fluid by the viscosity, μ. flowing phase). That is, the permeability of a
When two or more fluid phases are flowing, it rock containing more than one phase is signifi-
becomes necessary to introduce terms for the cantly lower than a rock with only one phase.
density, viscosity and permeability of each Clearly the fluid viscosity is a key factor but the
phase and for the interfacial forces (both fluid- fluid-fluid interactions also play a role. The
fluid and fluid-solid). For two-phase immiscible functions are drawn between ‘endpoints,’ which
flow (oil and water), the two-phase Darcy equa- are a mathematical convenience, but are also
tion and the capillary pressure equation are based on physical phenomena – the point at
used: which the flow rate of one phase becomes insig-
nificant. However, the endpoint values them-
uo ¼ kkro =μo :∇ðPo þ ρo gzÞ ð4:3Þ selves are not physically fixed. For example,
there exists a measurable irreducible water satu-
uw ¼ kkrw =μw :∇ðPw þ ρw gzÞ ð4:4Þ ration, but its precise value depends on many
things (e.g. oil phase pressure or temperature).
Pc ¼ Po  Pw ð4:5Þ Many of the problems and errors in upscaling
4.2 Multi-phase Flow 121

Endpoint kro

Relative Permeability
Total fluid
mobility Endpoint krw

kro
krw

0 1
Water Saturation
Also called Swi Swc Swor Sor = 1 - Swor

Fig. 4.5 Example oil-water relative permeability functions

Fig. 4.6 Example gas-oil 1


relative permeability
functions
Straight-line
Relative Permeability

approximation

krog krg

0 Gas Saturation 1

Sgc Sgmax Sgmax = 1 - Sor - Swi

arise from poor treatment or understanding of Typical values for a water-wet light oil might be:
these endpoints.
The most common functions used for relative kro ¼ 0:85ð1  Swn Þ3 and krw ¼ 0:3ðSwn Þ3
permeability are the Corey exponent functions:
ð4:9Þ
kro ¼ Að1  Swn Þx ð4:6Þ A similar set of functions can be used to
describe a gas-oil system (Fig. 4.6), where the
krw ¼ BðSwn Þy ð4:7Þ functions are bounded by the critical gas satura-
tion, Sgc, and the maximum gas saturation, Sgmax.
where Swn is the normalized saturation: However, gas-oil relative permeability curves
tend to have less curvature (lower Corey
Swn ¼ ðSw  Swc Þ=ðSwor  Swc Þ ð4:8Þ exponents) and sometimes straight-line functions
122 4 Upscaling Flow Properties

are assumed, implying perfect mixing or a fully- distributions have a fairly flat function (as for
miscible gas-oil system. the 1,000 mD curve in Fig. 4.7), while highly
These functions describe the flows and variable pore size distributions have a gradually
pressures for multi-phase flow. The third equa- rising function (as with the 50 mD curve in
tion required to completely define a two-phase Fig. 4.7). The capillary entry pressure is a func-
flow system is the capillary pressure equation. tion of the largest accessible pore. Different Pc
For the general case (any fluid pair): curves are followed for drainage (oil invasion)
and imbibition (waterflood) processes.
Pc ¼ Pnonwetting phase We summarise our introduction by noting that
 Pwettingphase ½Pc ¼ f ðSÞ ð4:10Þ the complexities of multi-phase flow boil down
to a set of rules governing how two or more
Capillary pressure, Pc, is a function of phase phases interact in the porous medium. Figure 4.8
saturation, and must be defined by a set of shows an example micro-model (an artificial
functions. The capillary pressure curve is a sum- etched-glass pore space network) in which fluid
mary of fluid-fluid interactions, and for any element phase distributions can be visualised. Even for
of rock gives the average phase pressures for all the this comparatively simple pore space, the num-
fluid-fluid contacts within the porous medium at a ber and nature of the fluid-fluid and fluid-solid
given saturation. For an individual pore, Pc can be interfaces is bewildering. What determines
related to measurable geometries (curvatures) and whether gas, oil or water will invade the next
forces (interfacial tension), and defined theoreti- available pore as the pressure in one phase
cally – but for a real porous medium it is an average changes?
property. Figure 4.7 shows some example One response – the modelling approach – is
measured Pc curves, based on mercury intrusion that good answers to this problem are found in
experiments (Neasham 1977). mathematical modelling of pore networks (e.g.
The slope of the Pc curve is related to the pore McDougall and Sorbie 1995; Blunt 1997; Øren
size distribution. More uniform pore-size and Bakke 2003; Behbahani and Blunt 2005).

1000
Small pores

100
0.5mD

Medium pores
Pc (psi)

50mD
10

1000mD
1 Large pores

Non-wetting phase
invades largest
0.1
1 0 pores first
PV occupied

Fig. 4.7 Example capillary pressure functions: capillary drainage curves based on mercury intrusion experiments
measuring the non-wetting phase pressure required to invade a certain pore volume (PV)
4.2 Multi-phase Flow 123

Fuller discussions of these methods are found


in, for example, Barker and Thibeau (1997),
Ekran and Aasen (2000), Pickup et al. (2005).
Reservoir simulators generally perform
dynamic multi-phase flow simulations – that is,
the pressures and saturations are allowed to vary
with position and time in the simulation grid. The
Kyte and Berry (1975) upscaling method is the
most well-known dynamic two-phase upscaling
method, but there have been many alternatives
proposed, such as Stone’s (1991) method and
Todd and Longstaff (1972) for miscible gas.
The strength of the dynamic methods is that
they attempt to capture the ‘true’ flow behaviour
for a given set of boundary conditions. Their
principle weaknesses are that they can be diffi-
cult and time-consuming to calculate and can be
plagued by numerical errors.
In contrast, the steady-state methods are easier
Fig. 4.8 Example micro-model, where fluid distributions to calculate and understand and represent ideal
are visualised within an artificial laboratory pore-space multi-phase flow behaviour. There are three
(Statoil archive image of micromodel experiment steady-state end-member assumptions:
conducted at Heriot Watt University)
• Viscous limit (VL): The assumption that the
flow is steady state at a given, constant frac-
tional flow. Capillary pressure is assumed to
Another response – the laboratory approach – is be zero.
that you need to measure the multiphase flow • Capillary equilibrium (CE): The assumption
behaviour in real rock samples at true reservoir that the saturations are completely controlled
conditions (pressures and temperatures). In real- by capillary pressure. Applied pressure
ity, you need both measurements and modelling gradients are assumed to be zero or negligible.
to obtain a good appreciation of the “rules” • Gravity-Capillary equilibrium (GCE): Similar
governing multiphase flow. Our concern here is to CE, except that in addition the saturations
to understand how to handle and upscale these are also controlled by the effect of gravity on
functions within the reservoir model. the fluid density difference. Note that GCE is
similar to the vertical equilibrium (VE)
assumption also applied in reservoir simulation
4.2.2 Two-Phase Steady-State (Coats et al. 1971), except that VE assumes
Upscaling Methods negligible capillary pressure.
The viscous limit assumption is similar to a
Multiphase flow upscaling, involves the process steady-state core flood experiment which is
of calculating the large-scale multiphase flows sometimes used in core analysis of multi-phase
given a known distribution of the small-scale flow (referred to as special core analysis, or
petrophysical properties and flow functions. SCAL). Here, a known and constant fraction of
There are many methods for doing this, but it is oil and water is injected into the sample (let us
useful to differentiate two: say 20 % oil and 80 % water) and the permeabil-
1. Dynamic methods ity for each phase is calculated from the pressure
2. Steady-state methods drop and flow rate for that phase. The procedure
124 4 Upscaling Flow Properties

Pc curve 7) Repeat for the


next pressure, etc.
1
1) Pick a
pressure

Pc / krel
kro
2,3,4) Find
the krel for krw
each phase

0
0 S 1
w
5) Solve the Darcy 6) Calculate effective
equations for each phase relative permeabilities

Fig. 4.9 Illustration of the capillary equilibrium steady-state upscaling method

is then repeated for a different fractional flow,


and so on. It is assumed that capillary pressure
and gravity have no effect. The method can be
assumed to apply for a Darcy-flow-dominated
two-phase flow system. The method is consid-
ered to be valid at larger length-scales, where
capillary forces can generally be neglected (e.g.
for model grid cell sizes greater than about 1 m
vertically).
For the capillary equilibrium steady state
assumption it is the Darcy flow effects that are
neglected and all fluxes are deemed to be con-
trolled by the capillary pressure curve. For a
given pressure, the saturation is known from the
Pc curve and the local phase permeability is then
determined from the relative permeability
curves. The calculation is then repeated for
each chosen decrement of pressure until the sat-
uration range is covered (Fig. 4.9). The method is
considered to be valid at smaller length-scales, Fig. 4.10 SEM image of laminae in an aeolian sandstone
where capillary forces are likely to dominate (Image courtesy of British Gas)
(e.g. at length-scales less than about 0.2 m).
There is also a rate-dependence for viscous and (Fig. 4.10), and therefore capillary forces are
capillary forces – higher flow rates favour vis- likely to be important at this length-scale.
cous forces while lower flow rates favour capil- The gravity-capillary equilibrium method
lary forces. Note that layering in sedimentary uses the same principle as the CE method
rock media is often at the mm to cm scale except that vertical pressure gradient is also
4.2 Multi-phase Flow 125

Upscaled Relative Permeability


1.0

Relative Permeability
0.8

krox
0.6
krwx
0.4
kroz krwz
0.2

0
Example Rock Curves 0 0.2 0.4 0.6 0.8 1.0
2.0
Pc Rock 1 Sw
1.8
krw Rock 1
1.6 kro Rock 1
krel or Pc (bars)

1.4 Pc Rock 2
krw Rock 2
1.2
kro Rock 2
1.0
0.8
0.6
0.4
0.2
Upscale
0.0
0.0 0.2 0.4 0.6 0.8 1.0
Sw

Fig. 4.11 Capillary equilibrium steady-state upscaling method applied to a simple layered model

applied resulting in a vertical trend in the 4. Find the phase permeabilities, e.g. ko1 ¼ k1 *
saturation at any chosen pressure reference. kro1;
The GCE solution should tend towards the 5. Calculate the upscaled permeability for each
CE solution as the length-scale becomes direction and for each phase using the arith-
increasingly small. metic and harmonic averages;
All three steady-state methods involve a series 6. Invert back to upscaled relative permeability,
of independent single-phase flow calculations and e.g. kro1 ¼ ko1/kupscaled (once again the arith-
therefore can employ a standard single-phase metic and harmonic averages are used to
pressure solver algorithm. The methods can there- obtain the upscaled absolute permeability);
fore be rapidly executed on standard computers. 7. Repeat for next value pressure, Pc2.
The capillary equilibrium method can be eas- Note that the upscaled curves are highly aniso-
ily calculated for a simple case, as illustrated in tropic, and in fact sometimes lie outside the range
Figure 4.11 by an example set input functions for of the input curves. This is because of the effects
a regular layered model. Upscaled relative per- of capillary forces – specifically capillary trapping
meability curves for this simple case can be cal- when flowing across layers. Capillary forces result
culated analytically (using a spreadsheet or in preferential imbibition of water (the wetting
calculator). The method uses the following phase) into the lower permeability layers, making
steps (refer to Fig. 4.9): flow of oil (the non-wetting phase) into these low
1. Chose a value for pressure, Pc1; permeability layers even more difficult.
2. Find the corresponding saturation value, Sw1; These somewhat non-intuitive effects of capil-
3. Determine the relative permeability for oil lary pressure in laminated rocks can be demon-
and water for each rock type, kro1, krw1, kro2, strated experimentally (Fig. 4.12). In the case of
krw2; two-phase flow across layers in a water-wet
126 4 Upscaling Flow Properties

laminated rock – a cross-bedded aeolian sandstone surfactant chemicals) or by increasing the flow
(Fig. 4.10), oil becomes trapped in the high perme- rate, and thereby the viscous forces. Alterna-
ability layers upstream of low permeability layers tively, a modified flow strategy favouring flow
due to water imbibition into the lower permeability along the rock layers (parallel to bedding) would
layers (Huang et al. 1995). result in less capillary trapping and a more effi-
The chosen flow rate for this experiment was cient waterflood. In the more general case, where
at typical reservoir waterflood rate (around the rock has variable wettability the effects of
0.1 m/day). This trapped oil can be mobilised capillary/viscous interactions become more com-
either by reducing the capillary pressure (e.g. plex (e.g. McDougall and Sorbie 1995; Huang
by modifying the interfacial tension by use of et al. 1996; Behbahani and Blunt 2005).

Exercise 4.1 (a) Calculate the upscaled horizontal and


Permeability upscaling for a simple layered vertical single-phase permeability using
model. averaging.
The simple repetitive layered model (b) Calculate selected values for the
shown in Fig. 4.11 can be used to illus- upscaled two-phase relative permeabil-
trate single and multi-phase permeability ity curves, assuming steady-state capil-
upscaling using the “back of the enve- lary equilibrium conditions. Use the
lope” maths. We assume the two layers flow functions shown in Fig. 4.11, as
are regular and of equal thickness. Rock tabulated below, with water saturation,
type 1 has a permeability of 100md and Sw; relative permeability to water, krw;
rock type 2 has a permeability of relative permeability to oil, krow; cap-
1,000md. illary pressure, Pc (bars). Choose Pc
values of 0.05 and 0.3 (shown in bold).

Table for rock type 1 (100 mD) Table for rock type 2 (1,000 mD)
Sw krw krow Pc Sw krw krow Pc
0.2092 0.0 0.9 10.0 0.05432 0.0 0.9 10.0
0.209791 0.0000004 0.897752 2.744926 0.055066 0.0 0.897752 0.976926
0.212154 0.000009 0.888792 0.938248 0.058048 0.000016 0.888792 0.333925
0.215108 0.000038 0.877668 0.590923 0.059 0.000023 0.887 0.3
0.221016 0.000151 0.855673 0.372172 0.061777 0.000065 0.877668 0.210311
0.226 0.0003 0.839 0.3 0.069234 0.000259 0.855673 0.132457
0.238740 0.000943 0.791683 0.201984 0.091604 0.001618 0.791683 0.071887
0.256464 0.002413 0.730655 0.147628 0.113974 0.004141 0.730655 0.052541
0.26828 0.003771 0.691590 0.127213 0.119 0.005 0.718 0.05
0.32736 0.015084 0.515190 0.080120 0.128888 0.006471 0.691590 0.045275
0.38644 0.033939 0.368967 0.061135 0.203456 0.025884 0.515190 0.028515
0.44552 0.060336 0.250969 0.050461 0.278024 0.058239 0.368967 0.021758
0.449 0.062 0.245 0.05 0.352592 0.103536 0.250969 0.017959
0.5046 0.094275 0.159099 0.043483 0.42716 0.161775 0.159099 0.015476
0.56368 0.135756 0.091074 0.038504 0.501728 0.232956 0.091074 0.013704
0.62276 0.184779 0.044366 0.034742 0.576296 0.317079 0.044366 0.012365
0.68184 0.241344 0.016099 0.031781 0.650864 0.414144 0.016100 0.011311
0.74092 0.305451 0.002846 0.029380 0.725432 0.524151 0.002846 0.010456
0.8 0.3771 0.000000 0.027386 0.8 0.6471 0.000000 0.009747
4.2 Multi-phase Flow 127

Fig. 4.12 Summary of a 0.6 10000


waterflood experiment
Capillary trapping of oil
across a laminated water-

Low-k layer
upstream of low-k layer
wet rock sample (Redrawn

Remaining Oil Saturation


from Huang et al. 1995,
#1995, Society of

Permeability (mD)
Petroleum Engineers Inc., 0.4
reproduced with
permission of SPE. Further
reproduction prohibited 1000
without permission)

0.2

Flow direction
0.0 100
0 10 20
Length along sample (cm)

4.2.3 Heterogeneity and Fluid Forces To treat this issue more formally, we use scal-
ing group theory (Rapoport 1955; Li and Lake
It is important to relate these multi-phase fluid 1995; Li et al. 1996; Dengen et al. 1997) to under-
flow processes to the heterogeneity being stand the balance of forces. The viscous/capillary
modelled. This is a fairly complex issue, and ratio and the gravity/capillary ratio are two of a
fundamental to what reservoir model design is number of dimensionless scaling group ratios that
all about. As a way in to this topic we use the can be determined to represent the balance of fluid
balance of forces concept to give us a framework forces. For example, for an oil-water system we
for understanding which scales most affect a can define the following force ratios:
particular flow process. For example, we know
that capillary forces are likely to be important for Viscous ux Δxμo
rocks with strong permeability variations at the ¼ ð4:11Þ
Capillary kx ðdPc =dSÞ
small scale (less than 20 cm scale is a good rule
of thumb).
Figure 4.13 shows a simple sketch of the end- Gravity Δρ g Δz
¼ ð4:12Þ
members of the fluid force system. We have three Capillary ðdPc =dSÞ
end members: gravity-, viscous- and capillary-
dominated. Reality will lie somewhere within the where,
triangle, but appreciation of the end-member Δx, Δz are system dimensions,
systems is useful to understand the expected ux is fluid velocity,
flow-heterogeneity interactions. Note, that for μo is the oil viscosity,
the same rock system the flow behaviour will kx is the permeability in the x direction,
be completely different for a gravity-dominated, (dPc/dS) is the slope of the capillary pressure
viscous-dominated or capillary-dominated flow function,
regime. The least intuitive is the capillary- Δρ is the fluid density difference and g is the
dominated case where water (for a water-wet constant due to gravity.
system) imbibes preferentially into the lower The viscous/capillary ratio is essentially a
permeability layers. ratio of Darcy’s law with a capillary pressure
128 4 Upscaling Flow Properties

Gravity dominated

Reality ?
Viscous dominated Capillary dominated

Fig. 4.13 The fluid forces triangle with sketches to illustrate how a water-flood would behave for a layered rock
(yellow ¼ high permeability layers)

gradient term, while the gravity/capillary ratio is


the buoyancy term against the capillary pressure
gradient. Δx and Δz represent the physical length
scales – essentially the size of the model in the x
and z directions. There are several different
forms of derivation of these ratios depending on
the physical assumptions and the mathematical 150psi
Dx
approach, but the form given above should allow
the practitioner to gain an appreciation of the 1 km
factors involved. It is important that a consistent
Fig. 4.14 Sketch of pressure drawdown between an
set of units are used to ensure the ratios remain injection and production well pair for water-flooding an
dimensionless. oil reservoir
For example, a calculation to determine when
capillary/heterogeneity interactions are impor-
tant can be made by studying the ratio of capil- ratio for different layer contrasts and heterogene-
lary to viscous forces. Figure 4.14 shows a ity length-scales (Ringrose et al. 1996).
reference well pair assuming 1 km well spacing If the layering in a reservoir occurs at the
and a 150 psi pressure drawdown at the produc- >10 m scale then viscous forces tend to dominate
ing well. We are interested in the balance of (or the Viscous/Capillary ratio must be very low
forces and a rock unit within the reservoir, for capillary forces to be significant at this scale).
represented by alternating permeability layers However, if the layers are in the mm-to-cm range
with a spacing of Δx. Figure 4.15 shows the then capillary forces are much more likely to be
result of the analysis of the viscous/capillary important (or the Viscous/Capillary ratio must be
4.3 Multi-scale Geological Modelling Concepts 129

Fig. 4.15 Example Tidy up text elements


calculation of the viscous/ 10000

(c. 150psi drawdown between wells 1km apart)


capillary ratio for a layered

Normalised to a viscous force of 1kPa/m


system as a function of Viscous–dominated
length scale, for selected

Viscous /Capillary Ratio


permeability contrasts 100
(Redrawn from Ringrose
et al. 1996, #1996, Society
of Petroleum Engineers
Inc., reproduced with
permission of SPE. Further 1
reproduction prohibited
without permission)

0.01
Capillary-dominated

0.0001
0.01 1 100
Heterogeneity lengthscale (m)
[Distance between layers of contrasting permeability]

very high to override capillary effects). Note, how- recognised for some time. Haldorsen and Lake
ever, that the pressure gradients will vary as a (1984) and Haldorsen (1986) proposed four
function of spatial position and time and so in conceptual scales associated with averaging
fact the Viscous/Capillary ratio will vary – viscous properties in porous rock media:
forces will be high close to the wells and lower in • Microscopic (pore-scale);
the inter-well region. • Macroscopic (representative elementary volume
An important and related concept is the capil- above the pore scale);
lary number, most commonly defined as: • Megascopic (the scale of geological heteroge-
neity and or reservoir grid blocks);
μq • Gigascopic (the regional or total reservoir
Ca ¼ ð4:13Þ
γ scale).
Weber (1986) showed how common sedimen-
where μ is the viscosity, q is the flow rate and γ is tary structures including lamination, clay drapes
the interfacial tension. and cross-bedding affect reservoir flow
This is a simpler ratio of the viscous force to properties and Weber and van Geuns (1990)
the surface tension at the fluid-fluid interface. proposed a framework for constructing
Capillary numbers around 104 or lower are geologically-based reservoir models for different
generally deemed to be capillary-dominated. depositional environments. Corbett et al. (1992)
and Ringrose et al. (1993) argued that multi-scale
modelling of water-oil flows in sandstones
4.3 Multi-scale Geological should be based on a hierarchy of sedimentary
Modelling Concepts architectures, with smaller scale heterogeneities
being especially important for capillary-
4.3.1 Geology and Scale dominated flow processes (see Sect. 2.3.2.2 for
an introduction to hierarchy). Campbell (1967)
The importance of multiple scales of heterogene- established a basic hierarchy of sedimentary
ity for petroleum reservoir engineering has been features related to fairly universal processes of
130 4 Upscaling Flow Properties

0 m 2

b
0 m 10
0 cm 20

0 cm 2

Fig. 4.16 Field outcrop sketches illustrating multi-scale (d) Fault deformation fabric around a normal fault
reservoir architecture through an inter-bedded sandstone and silty clay sequence
(a) Sandstone and siltstone lamina-sets from a weakly- (Redrawn from Ringrose et al. 2008, The Geological
bioturbated heterolithic sandstone Society, London, Special Publications 309 # Geological
(b) Sandy and muddy bed-sets in a tidal deltaic lithofacies Society of London [2008])
(c) Prograding sedimentary sequences from a channelized
tidal delta

deposition, namely lamina, laminasets, beds and scale, lithofacies-scale and sequence-stratigraphic
bedsets. Miall (1985) showed how the range of scale elements can be identified. In addition to the
sedimentary bedforms can be defined by a series importance of correctly describing the sedimen-
of bounding surfaces from a 1st order surface tary length scales, structural (Fig. 4.16d) and dia-
bounding the laminaset to 4th (and higher) genetic processes act to modify the primary
order surfaces bounding, for example, composite depositional fabric.
point-bars in fluvial systems. At the most elemental level we are interested
Figure 4.16 illustrates the geological hierarchy in the pore scale (Fig. 4.17) – the rock pores that
for a heterolithic sandstone reservoir. Lamina- contain fluids and determine the multi-phase flow
4.3 Multi-scale Geological Modelling Concepts 131

Fig. 4.17 The pore scale –


example thin section of
pores in a sandstone
reservoir (Statoil image
archive, # Statoil ASA,
reproduced with
permission)

behaviour. Numerical modelling at the pore scale Pixed-based modelling approaches (e.g. SGS,
has been widely used to better understand perme- SIS) can be applied at pretty-much any scale,
ability, relative permeability and capillary pres- whereas object-based modelling approaches
sure behaviour for representative pore systems will tend to have very clear associations with
(e.g. Bryant and Blunt 1992; Bryant et al. 1993; pre-defined length scales. In both cases the
McDougall and Sorbie 1995; Bakke and Øren model grid resolution needs to be fine enough
1997; Øren and Bakke 2003). Most laboratory to explicitly capture the heterogeneity being
analysis of rock samples is devoted to measuring represented in the model. Process-based
pore-scale properties – resistivity, acoustic modelling methods (e.g. Rubin 1987; Wen et al.
velocity, porosity, permeability, and relative per- 1998; Ringrose et al. 2003) are particularly
meability. Pore-scale modelling allows these appropriate for capturing the effects of small-
measured flow properties to be related to funda- scale geological architecture within a multi-
mental rock properties such as grain size, grain scale modelling framework.
sorting and mineralogy. However, the applica- In the following sections we look at some key
tion of pore-scale measurements and models in questions the reservoir modelling practitioner
larger-scale reservoir models requires a frame- will need to address in building multi-scale res-
work for assigning pore-scale properties to the ervoir models:
geological concept. We do this by assigning flow 1. How many scales to model and upscale?
properties to lamina-scale, lithofacies-scale or 2. Which scales to focus on?
stratigraphic-scale models. This can be done 3. How to best construct model grids?
quite loosely, with weak assumptions, or system- 4. Which heterogeneities matter most?
atically within a multi-scale upscaling hierarchy.
Statistical methods for representing the spatial
architecture of geological systems were covered 4.3.2 How Many Scales to Model
in Chap. 2. What concerns us here is how we and Upscale?
integrate geological models within a multi-scale
hierarchy. This may require a re-evaluation of Despite the inherent complexities of sedimentary
the scales of models needed to address different systems, dominant scales and scale transitions can
scale transitions. be identified (Fig. 4.18). These dominant scales
132 4 Upscaling Flow Properties

Fig. 4.18 Examples of geologically-based reservoir sim- (dimensions 80 m  1 km  3 km);


ulation models at four scales (d) Reservoir simulation grid for part of the Heidrun field
(a) Model of pore space used as the basis for multi-phase illustrating grid cells displaced by faults in true structural
pore network models (50 μm cube); position (dimensions 200 m  3 km  5 km)
(b) Model of lamina-sets within a tidal bedding facies (Statoil image archives, # Statoil ASA, reproduced with
(dimensions 0.05 m  0.3 m  0.3 m); permission)
(c) Facies architecture model from a sector of the Heidrun
field showing patterns of tidal channel and bars

are based both on the nature of rock heterogeneity 2. Lithofacies to geomodel. Where a larger-scale
and the principles which govern macroscopic flow geological concept (e.g. a sequence strati-
properties. In this discussion, we assume four graphic model, a structural model or a diage-
scales – pore, lithofacies, geomodel and reservoir. netic model) postulates the spatial arrangement
This gives us three scale transitions: of lithofacies elements. Here, the geomodel is
1. Pore to lithofacies. Where a set of pore-scale taken to mean a geologically-based model of
models is applied to models of lithofacies the reservoir, typically resolved at the sequence
architecture to infer representative or typical or zone scale.
flow behaviour for that architectural element. 3. Geomodel to reservoir simulator. This stage
The lithofacies is a basic concept in the may often only be required due to computational
description of sedimentary rocks and limitations, but may also be important to ensure
presumes an entity that can be recognised good transformation of a geological model into
routinely. The lamina is the smallest sedimen- 3-dimensional grid optimised for flow simula-
tary unit, at which fairly constant grain depo- tion (e.g. within the constraints of finite-
sition processes can be associated with a difference multiphase flow simulation). This
macroscopic porous medium. The lithofacies third step is routinely taken by practitioners,
comprises some recognisable association of whereas steps 1 and 2 tend to be neglected.
laminae and lamina sets. In certain cases, Features related to structural deformation
where variation between laminae is small, (faults, fractures and folds) occur at a wide
pore-scale models could be applied directly range of scales (Walsh et al. 1991; Yielding
to the lamina-set or bed-set scales. et al. 1992) and do not naturally fall into a
4.3 Multi-scale Geological Modelling Concepts 133

step-wise upscaling scheme. Structural features modelling of the petrophysical properties within
are typically incorporated at the geomodel scale. the seismically resolved element (see Chap. 2).
However, effects of smaller scale faults may Upscaling methods impose further limitations
also be incorporated as effective properties (as on the value and utility of models within a multi-
transmissibility multipliers) using upscaling scale framework. In conventional upscaling –
approaches. The incorporation of fault transmis- from a geological model to a reservoir simulation
sibility into reservoir simulators is considered grid – there are various approaches used. These
thoroughly by Manzocchi et al (2002). Conduc- cover a range which can be classed in terms of
tive fractures may also affect sandstone the degree of simplification/complexity:
reservoirs, and are often the dominant factor in 1. Averaging of well data directly into the flow
carbonate reservoirs. Approaches for multi-scale simulation grid: This approach essentially
modelling of fractured reservoirs have also been ignores explicit upscaling and neglects all
developed (e.g. Bourbiaux et al. 2002) and will aspects of smaller scale structure and flows.
be developed further in Chap. 6. The approach is fast and simple and may be
Historical focus over the last few decades has useful for quick assessment of expected reser-
been on including increasingly more detail into voir flows and mass balance. It may also be
the geomodel, with only one upscaling step being adequate for very homogeneous and highly
explicitly performed. Full-field geomodels are permeable rock sequences.
typically in the size range of 1–10 million cells 2. Single-phase upscaling only in Δz: This com-
with horizontal cell sizes of 50–100 m and vertical monly applied approach assumes a simulation
cell sizes of order 1–10 m. Multi-scale modelling grid designed with the same Δx, the Δy as the
allows for better flow unit characterization and geological grid. The approach is often used
improved performance predictions (e.g. Pickup where complex structural architecture
et al. 2000; Scheiling et al. 2002). There are also provides very tight constraints to design of
examples where a large number of grid cells are the flow modelling grid. Upscaling essentially
applied to sector or near-well models reducing cell comprises use of averaging methods but
sizes to the dm-scale. Upscaling of the near-well ensures a degree of representation of thin
region requires methods to specifically address layering or barriers. Also, where seismic data
radial flow geometry (e.g. Durlofsky et al. 2000). gives a good basis for the geological model in
Recent focus on explicit small-scale the horizontal dimensions, vertical upscaling
lithofacies modelling includes the use of million of fine-scale layering to the reservoir simula-
cell models with mm to cm size cells (e.g. tor scale is typically required.
Ringrose et al. 2003; Nordahl et al 2005). 3. Single-phase upscaling in Δx Δy and Δz: With
Numerical pore-scale modelling employs a simi- this approach multi-scale effective flow
lar number of network nodes at the pore scale properties are explicitly estimated and the
(e.g. Øren and Bakke 2003). Model resolution is upscaling tools are widely available (diagonal
always limited by the available computing tensor or full-tensor methods). Multiphase
power, and although continued efficiencies and flow effects are however neglected.
memory gains are expected in the future, the use 4. Multi-phase upscaling in Δx Δy and Δz: This
of available numerical discretisation at several approach represents an attempt to calculate
scales within a hierarchy is preferred to efforts effective multiphase flow properties in larger
to apply the highest possible resolution at one of scale models. The approach has been used
the scales (typically the geomodel). There is also rather too seldom due to demands of time
an argument that advances in seismic imaging and resources. However, the development of
coupled with computing power will enable direct steady-state solutions to multiphase flow
geological modelling at the seismic resolution upscaling problems (Smith 1991; Ekran and
scale. However, even when this is possible, Aasen 2000; Pickup and Stephen 2000) has
seismic-based lithology prediction (using seis- led to wider use in field studies (e.g. Pickup
mic inversion) will require smaller-scale et al 2000; Kløv et al 2003).
134 4 Upscaling Flow Properties

These four degrees of upscaling complexity geological scale, as in the example shown in
help define the number and dimensions of models Fig. 4.18.
required. The number of scales modelled is typi-
cally related to the complexity and precision of
answer sought. Improved oil recovery (IOR) 4.3.3 Which Scales to Focus On?
strategies and reservoir drainage optimisation (The REV)
studies are often the reason for starting a multi-
scale approach. A minimum requirement for any Geological systems present us with variability at
reservoir model is that the assumptions used for nearly every scale (Fig. 4.19). To some extent
smaller scale processes (pore scale, lithofacies they are fractal (Turcotte 1992), showing similar
scale) are explicitly stated. variability at all scales. However, geological
For example, a typical set of assumptions systems are more accurately described as multi-
commonly used might be: fractal – showing some scale-independent
We assume that two special core analysis similarities – but dominated by process-controlled
measurements represent all pore-scale physical scale-dependent features (e.g. Ringrose 1994).
flow processes and that all effects of geological However you describe them, geological systems
architecture are adequately summarised by the are complex, and we need an approach for
arithmetic average of the well data.
simplifying that complexity and focussing on the
Assumptions like these are rarely stated important features and length-scales.
(although often implicitly assumed). More ide- The Representative Elementary Volume
ally, some form of explicit modelling at each (REV) concept (Bear 1972) provides the essen-
scale should be performed using 3D multiphase tial framework for understanding measurement
upscaling methods. At a minimum, it is scales and geological variability. This concept is
recommended to explicitly define pore-scale fundamental to the analysis of flow in permeable
and geological-scale models, and to determine a media – without a representative pore space we
rationale for associating the pore-scale with the cannot measure a representative flow property

Fig. 4.19 Multi-scale variability in a heterolithic (tidal bedsets (hammer for scale); Sequence-stratigraphic
delta) sandstone system: Laminaset scale: Core photo- scale: Sand-dominated para-sequence between mudstone
graph with measured permeability (Red indicates >1 units (Photos A. Martinius/Statoil # Statoil ASA,
Darcy); Bedset scale: interbedded sandy and muddy reproduced with permission)
4.3 Multi-scale Geological Modelling Concepts 135

Fig. 4.20 The


Representative Elementary Pore
Volume (REV) concept,
after Bear 1972
Mainly pores

Property
The REV
A representative
Mainly grains
average of
grains and pores
Grain

Sample Volume

Fig. 4.21 The pore-scale REV illustrated for an example thin section (The whole image is assumed to be the pore-scale
REV) (Photo K. Nordahl/Statoil # Statoil ASA, reproduced with permission)

nor treat the medium as a continuum in terms of develop a multi-scale approach to the REV con-
the physics of flow. The original concept cept. It is not at first clear how many averaging
(Fig. 4.20) refers to the scale at which pore- length-scales exist in a rock medium, or indeed if
scale fluctuations in flow properties approach a a REV can be established at the scale necessary
constant value both as a function of changing for reservoir flow simulation. Despite the
scale and position in the porous medium, such challenges, some degree of representativity of
that a statistically valid macroscopic flow prop- estimated flow properties is necessary for flow
erty can be defined, as illustrated in Fig. 4.21. modelling within geological media, and a multi-
The pore-scale REV is thus an essential scale REV framework is required.
assumption for all reservoir flow properties. Several workers (e.g. Jackson et al. 2003;
However, rock media have several such scales Nordahl et al. 2005) have shown that an REV
where smaller-scale variations approach a more can be established at the lithofacies scale – e.g. at
constant value. It is therefore necessary to around a length-scale of 0.3 m for tidal
136 4 Upscaling Flow Properties

Fig. 4.22 The lithofacies REV illustrated for an example heterolithic sandstone (Photo K. Nordahl/Statoil # Statoil
ASA, reproduced with permission)

heterolithic bedding (Fig. 4.22). In fact, the con- of the two datasets differ – the core plugs at a
cept of representivity is inherent in the definition smaller scale record high degree of variability
of a lithofacies, a recognisable and mappable while the wireline data provides a more averaged
subdivision of a stratigraphic unit. The same result at a larger scale. Both sets of data can be
logic follows at larger geological scales, such as integrated into a petrophysical model at the
the parasequence, the facies association or the lithofacies REV. Nordahl and Ringrose (2008)
sequence stratigraphic unit. Recognisable and extended this concept to propose a multi-scale
continuous geological units are identified and REV framework (Fig. 4.24), whereby the natural
defined by the sedimentologist, and the reservoir averaging length scales of the geological system
modeller then seeks to use these units to define can be compared with the various measurement
the reservoir modelling elements (cf Chap. 2.4). length scales.
As a general observation, core plug data is Whatever the true nature of rock variability, it
often not sampled at the REV scale and therefore is a common mistake to assume that the averaging
tends to show a wide scatter in measured values, inherent in any measurement method (e.g. electri-
while wire-line log data is often closer to a cal logs or seismic wave inversion) relates directly
natural REV in the reservoir system. The true to the averaging scales in the rock medium. For
REV – if it can be established – is determined example, samples from core are often at an inap-
by the geology and not the measurement device. propriate scale for determining representativity
However, wire-line log data usually needs labo- (Corbett and Jensen 1992; Nordahl et al. 2005).
ratory core data for calibration, which presents us At larger scales, inversion of reservoir properties
with a dilemma – how should we integrate dif- from seismic can be difficult or erroneous due to
ferent scales of measurement? thin-bed tuning effects. Instead of assuming that
Nordahl et al. (2005) performed a detailed any particular measurement gives us an appropri-
assessment of the REV for porosity and perme- ate average, it is much better to relate the mea-
ability in a heterolithic sandstone reservoir unit surement to the inherent averaging length scales in
(Fig. 4.23). This example illustrates how appar- the rock system.
ently conflicting datasets from core plug and So how do we handle the REV concept in
wireline measurements can in fact be reconciled practice? The key issue is to find the length-
within the REV concept. The average and spread scale (determined by the geology) where the
4.3 Multi-scale Geological Modelling Concepts 137

Fig. 4.23 Assessment of the lithofacies REV, from 90th percentile. The solid line is the median and the black
Nordahl et al. (2005). Comparison of porosity (a) and dots are the outliers. The values at the REV are measured
horizontal permeability (b) estimated or measured from on the bedding model at a representative scale (With the
different sources and sample volumes. The lower and distribution based on ten realisations) (Redrawn from
upper limits of the box indicate the 25th and the 75th Nordahl et al. 2005, Petrol Geoscience, v. 11 # Geologi-
percentile while the whiskers represent the 10th and the cal Society of London [2005])

Thin section Probe Core Logging Welltests


& SEM Perm. plugs tools & seismic

Sequence REV
Permeability

Lithofacies REV

Lamina REV

10-12 10-11 10-10 10-9 10-8 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 104
Measurement Volume [m3] (log scale)

Fig. 4.24 Sketch illustrating multiple scales of REV within a geological framework and the relationship to scales of
measurement (Adapted from Nordahl and Ringrose 2008)

measurement or model gives a representative 4.3.4 Handling Variance as a Function


average of the smaller-scale natural variations of Scale
(Fig. 4.25). At the pore-scale this volume is
typically around a few mm3. For heteroge- Typical practice in petroleum reservoir studies is
neous rock systems the REV is of the order to assume that an average measured property for
of m3. The challenge is to find the representa- any rock unit is valid and that small-scale
tive volumes for the reservoir system in the variability can be ignored. Put more simply, we
subsurface. often assume that the average log-property
138 4 Upscaling Flow Properties

Fig. 4.25 Rock sculpture by Andrew Goldsworthy (NW Highlands of Scotland) elegantly capturing the concept of the
REV

response for a well through a reservoir interval is Statistics for ln(k) are shown as the population
the ‘right average.’ A statistician will know that distributions are approximately log normal. It is
an arbitrary sample is rarely an accurate repre- well known that the sample variance should
sentation of the truth. Valid statistical treatment reduce as sample scale is increased. Therefore,
of sample data is an extensive subject treated the reduction in variance between datasets (c)
thoroughly in textbooks on statistics in the and (d) – core data to reservoir model – is
Earth Sciences – e.g. Size (1987), Davis (2003), expected. It is, however, a common mistake in
Isaaks and Srivastava (1989) and Jensen et al. multi-scale reservoir modelling for an inappropri-
(2000). ate variance to be applied in a larger scale model,
The challenges involved in correctly inferring e.g. if core plug variance was used directly to
permeability from well data are illustrated here represent the upscaled geomodel variance.
using an example well dataset (Fig. 4.26, Comparison of datasets (a) and (b) reveals
Table 4.2). This 30 m cored-well interval is from another form of variance that is commonly
a tidal deltaic reservoir unit with heterolithic ignored. The probe permeameter grid (2 mm
lithofacies and moderate to highly variable spaced data over a 10 cm  10 cm core area)
petrophysical properties (the same well dataset is shows a variance of 0.38 [ln(k)]. The core plug
discussed in detail by Nordahl et al. (2005)). dataset for the corresponding lithofacies interval
Table 4.2 compares the permeability statistics (estuarine bar), has σ2 ln(k) ¼ 0.99, which
for different types of data from this well: represents variance at the lithofacies scale. How-
(a) High resolution probe permeameter data; ever, blocking of the probe permeameter data at
(b) Core plug data; the core plug scale shows a variance reduction
(c) A continuous wireline-log based estimator of factor of 0.79 up to the core plug scale (column
permeability for the whole interval; 2 in Table 4.2). Thus, in this dataset (where high
(d) A blocked permeability log as might be typi- resolution measurements are available) we know
cally used in reservoir modelling. that a significant degree of variance is missing
4.3 Multi-scale Geological Modelling Concepts 139

Probe permeameter data Permeability, ln(k) (mD)


-4 -2 0 2 4 6 8
2590
0.4
Frequency

0.2
2595

0
Estuarine bar
10

100

1000

10000
2600
k (md)
Depth
(m)

2605

Estuarine bar core plugs


Whole interval core plugs
Wireline k-estimate 2610

Blocked wireline k

2615

2620

Fig. 4.26 Example dataset from a tidal deltaic flow unit Geological Society, London, Special Publications 309
illustrating treatment of permeability data used in reser- # Geological Society of London [2008])
voir modelling (Redrawn from Ringrose et al. 2008, The

Table 4.2 Variance analysis of example permeability dataset


Estuarine bar lithofacies Whole interval (flow unit)
(b) Probe data (e) (f)
upscaled to plug (c) Core plug (d) Core Wireline-k Blocked
(a) Probe-k data scale data plug data estimate well data
Scale of data 10  10 cm; 2  2 cm squares c.15–30 cm c.15–30 cm 15 cm 2m
2 mm spaced of 2 mm-spaced spaced core spaced plugs digital log blocking
data data plugs
N¼ 2,584 25 11 85 204 16
Mean ln(k) 7.14 7.14 6.39 1.73 2.32 2.17
σ2 ln(k) 0.38 0.30 0.99 8.44 5.94 4.80
Variance – 0.79 – – – 0.81
adjustment
factor, f
140 4 Upscaling Flow Properties

from the datasets conventionally used in reser- needed to handle and upscale the data in order to
voir modelling. derive an appropriate average. Assuming that we
Improved treatment of variance in reservoir have datasets which can be related to the REV’s in
modelling is clearly needed and presents us with the rock system, we can then use the same multi-
a significant challenge. The statistical basis for scale framework to guide the modelling length
treating population variance as a function of scales. Reservoir model grid-cell dimensions
sample support volume is well established with should ideally be determined by the REV
the concept of Dispersion Variance (Isaaks and lengthscales. Explicit spatial variations in the
Srivastava 1989), where: model (at scales larger than the grid cell) are
then focussed on representing property variations
σ2 ða; cÞ ¼ σ2 ða; bÞ þ σ2 ðb; cÞ that cannot be captured by averages. To put this
Total Variance Variance concept in its simplest form consider the follow-
variance within blocks between blocks ing modelling steps and assumptions:
ð4:14Þ 1. From pore scale to lithofacies scale: Pore-scale
models (or measurements) are made at the
where a, b and c represent different sample pore-scale REV and then spatial variation
supports (in this case, a ¼ point values, b ¼ at the lithofacies scale is modelled (using
block values and c ¼ total model domain). deterministic/probabilistic methods) to estimate
The variance adjustment factor, f, is defined as rock properties at the lithofacies-scale REV.
the ratio of block variance to point variance and 2. From lithofacies scale to geomodel scale.
can be used to estimate the correct variance to be Lithofacies-scale models (or measurements)
applied to a blocked dataset. For the example are made at the lithofacies-scale REV and then
dataset (Table 4.2, Fig. 4.26) the variance adjust- spatial variation at the geological architecture
ment factor is around 0.8 for both scale adjust- scale is modelled (using deterministic/probabi-
ment steps. listic methods) to estimate reservoir properties
With additive properties, such as porosity, at the scale of the geological-unit REV (equiva-
treatment of variance in multi-scale datasets is lent to geological model elements).
relatively straightforward. However, it is much 3. From geomodel to full-field reservoir simula-
more of a challenge with permeability data as tor. Representative geological model
flow boundary conditions are an essential aspect elements are modelled at the full-field reser-
of estimating an upscaled permeability value voir simulator scale to estimate dynamic flow
(see Chap. 3). Multi-scale geological modelling behaviour based on reservoir properties that
is an attempt to represent smaller scale structure have been correctly upscaled and are (arguably)
and variability as an upscaled block permeability representative.
value. In this process, the principles guiding There is no doubt that multi-scale modelling
appropriate flow upscaling are essential. How- within a multi-scale REV framework is a chal-
ever, improved treatment of variance is also crit- lenging process, but it is nevertheless much pre-
ical. There is, for example, little point rigorously ferred to ‘throwing in’ some weakly-correlated
upscaling a core plug sample dataset if it is random noise into an arbitrary reservoir grid and
known that that dataset is a poor representation hoping for a reasonable outcome. The essence of
of the true population variance. good reservoir model design is that it is based on
The best approach to this rather complex prob- some sound geological concepts, an appreciation
lem, is to review the available data within a multi- of flow physics, and a multi-scale approach to
scale REV framework (Fig. 4.24). If the dataset is determining statistically representative properties.
sampled at a scale close to the corresponding Every reservoir system is somewhat unique,
REV, then it can be considered as fairly reliable so the best way to apply this approach method is
and representative data. If however, the dataset is try it out on real cases. Some of these are
clearly not sampled at the REV (and is in fact illustrated in the following sections, but consider
recording a highly variable property) then care is trying Exercise 4.2 for your own case study.
4.3 Multi-scale Geological Modelling Concepts 141

Exercise 4.2 variability and low variability (the REV) –


Find the REVs for your reservoir? similar to Fig. 4.21 – using the sketch below.
Use your own knowledge a particular Note that the horizontal axis is given as
geological reservoir system or outcrop to a vertical length scale (dz, across bedding)
sketch on the most likely scales of high to make volume estimation easier.

Scales of measurement

1000 Lithofacies Geological Sequence


Pore/lamina
scale scale
Permeability (md)

scale

100

10

0.0001 0.001 0.01 0.1 1 10 100


Vertical Lengthscale [m]

4.3.5 Construction of Geomodel adequate 3D grids for realistic fault architectures,


and Simulator Grids and significant manual work is necessary.
Upscaling procedures for regular Cartesian
The choice of grid and grid-cell dimensions is grids are well established, but the same operation
clearly important. Upscaled permeability, the in realistically complex grids is much more
balance of fluid forces, and reservoir property challenging.
variance are all intimately connected with the The construction of 3D grids suitable for res-
model length-scale. The construction of three ervoir simulation is also non-trivial and requires
dimensional geological models from seismic significant manual editing. There are several
and well data remains a relatively time consum- reasons for this:
ing task requiring considerable manual work • The grid resolution in the geologic model and
both in construction of the structural framework the simulation models are different, leading to
and, not least, in construction of the grid for missing cells or miss fitting cells in the simu-
property modelling (Fig. 4.27). lation model. The consequences are overesti-
Problems especially arise due to complex mation of pore volumes, possibly wrong
fault block geometries including reverse faults communication across faults, and difficult
and Y-faults (Y-shaped intersections in the verti- numerical calculations due to a number
cal plane). Difficulties relate partly to the small or “artificial” grid cells.
mapping of horizons into the fault planes for • The handling of Y-shaped faults using corner
construction of consistent fault throws across point grid geometries (now widely used in
faults. Currently, most commercial gridding soft- black oil simulators) is difficult. Similarly,
ware is not capable of automatically producing the use of vertically stair-stepped faults
142 4 Upscaling Flow Properties

Fig. 4.27 Example


reservoir model grid
(Heidrun Field fault
segments, colour coded by
reservoir segment) (Statoil
image archives, # Statoil
ASA, reproduced with
permission)

improves the grid quality and flexibility, but • Flow simulation accuracy depends on the grid
does not solve the whole problem. When quality, and the commonly used numerical
using grids with stair-step faults special atten- discretisation schemes in commercial
tion must be paid to estimation of fault seal simulators have acceptable accuracy only for
and fault transmissibility. There is generally ‘near’ orthogonal grids. Orthogonal grids do
insufficient information in the grid itself for not comply easily with complex fault
these calculations, and the calculation of fault structures, and most often compromises are
transmissibility must be calculated based on made between honouring geology and
information from the conceptual geological keeping “near orthogonal” grids.
model. Figure 4.28 illustrates how some of these
• The handling of dipping reverse faults using problems have been addressed in oilfield studies
stair-step geometry in a corner point grid (Ringrose et al. 2008). After detailed manual grid
requires a higher total number of layers than construction including stair-step faults to handle
required for an un-faulted model. Y-faults, smaller faults are added directly into
• Regions with fault spacing smaller than the the flow simulation grid. However, some
simulation grid spacing give problems for gridding problems cannot be fully resolved
appropriate calculation of fault throw and using the constraints of corner point simulation
zone to zone communication. Gridding grids and optimal, consistent and automated grid
implies that smaller-scale geomodel faults generation based on realistic geomodels is a chal-
are merged and a cumulated fault throw is lenge. The use of unstructured grids reduces
used in the simulation model. This is not gen- some of the gridding problems, but robust, reli-
erally possible with currently available able and cost efficient numerical flow solution
gridding tools, and an effective fault transmis- methods for these unstructured grids are not gen-
sibility, including non-neighbour connections, erally available. Improved and consistent
must be calculated based on information from solutions for construction of structured grids
the geomodel, i.e. using the actual geometry and associated transmissibilities have been pro-
containing all the merged faults. posed (e.g. Manzocchi et al 2002; Tchelepi et al.
4.3 Multi-scale Geological Modelling Concepts 143

Fig. 4.28 Illustration of


the transfer of a structural
geological model to a
reservoir simulation grid
(Redrawn from Ringrose
et al. 2008, The Geological
Society, London, Special
Publications 309 #
Geological Society of
London [2008])

2005) and flow simulation on faulted grids significant. This is a clear argument in favour
remains a challenge. of explicit multi-scale reservoir modelling.
Furthermore, in the studies where the effects
of structural heterogeneity were assessed, both
4.3.6 Which Heterogeneities Matter? structural and sedimentary features were found to
be significant. That is to say, structural features
There are a number of published studies in and uncertainties cannot be neglected and are
which the importance of different multi-scale fully coupled with stratigraphic factors.
geological factors on reservoir performance Another approach to this question is to con-
have been assessed. Table 4.3 summarizes the sider how the fluid forces will interact with the
findings of a selection of such studies in which a heterogeneity in terms of the REV (Fig. 4.29).
formalised experimental design with statistical Pore and lamina-scale variations have the stron-
analysis of significance has been employed. The gest effect on capillary-dominated fluid pro-
table shows only the main factors identified in cesses while the sequence stratigraphic (or
these studies (for full details refer to sources). facies association) scale most affects flow pro-
What is clear from this work is that several cesses in the viscous-dominated regime. Gravity
scales of heterogeneity are important for each operates at all scales, but gravity-fluid effects are
reservoir type. While one can conclude that most important at the larger scales, where signif-
stratigraphic sequence position is the most icant fluid segregation occurs. That is, when both
important factor in a shallow marine deposi- capillary forces and applied pressure gradients
tional setting or that vertical permeability is fail to compete effectively against gravity
the most important factors in tidal deltaic stabilisation of the fluids involved.
setting, each case study shows that both larger Several projects have demonstrated the eco-
and smaller-scale factors are generally nomic value of multi-scale modelling in the
144 4 Upscaling Flow Properties

Table 4.3 Summary of selected studies comparing multi-scale factors on petroleum reservoir performance
Shallow Faulted Shallow Fault
Marinea Marineb Fluvialc Tidal Deltaicd modellinge
Sequence model V V V
Sand fraction S S V S n/a
Sandbody geometry S S n/a
Vertical permeability S S V n/a
Small-scale heterogeneity S S n/a
Fault pattern n/a S n/a n/a S
Fault seal n/a S n/a n/a S
V Most significant factor, S Significant factor, n/a not assessed
a
Kjønsvik et al. (1994)
b
England and Townsend (1998)
c
Jones et al. (1993)
d
Brandsæter et al. (2001a)
e
Lescoffit and Townsend (2005)

Gravity-dominated
Fluid
segregation
Permeability

Capillary trapping Sequence


and Sor REV
Lithofacies
Lamina REV
REV
Viscous
fingering and
channeling
Capillary-dominated Viscous-dominated

10-12 10-11 10-10 10-9 10-8 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 101 102 103 104
Measurement Volume [m3] (log scale)

Fig. 4.29 Sketch illustrating the expected dominant fluid forces with respect to the important heterogeneity length-
scales

context of oilfield developments. An ambitious alternating-gas (WAG) injection on the


study of the structurally complex Gullfaks field Veslefrikk Field (Kløv et al 2003), and depres-
(Jacobsen et al 2000) demonstrated that 25 surization on the Statfjord field (Theting et al
million-cell geological grid (incorporating struc- 2005). These studies typically show of the order
tural and stratigraphic architecture) could be of 10–20 % difference in oilfield recovery factors
upscaled for flow simulation and resulted in a when advanced multi-scale effects are
significantly improved history match. Both strat- implemented, compared with conventional
igraphic barriers and faults were key factors in single-scale reservoir simulation studies. For
achieving improved pressure matches to historic example, Figure 4.31 shows the effect of one-
well data. This model was also used for assess- step and two-step upscaling for the gas injection
ment of IOR using CO2 flooding. case study (illustrated in Fig. 4.30). The coarse-
Multi-scale upscaling has also been used to grid case without upscaling gives a forecasting
assess complex reservoir displacement pro- error of over 10 % when compared to the fine-
cesses, including gas injection in thin-bedded grid reference case, while the coarse-grid case
reservoirs (Fig. 4.30) (Pickup et al 2000; with two-step upscaling gives a result very close
Brandsæter et al. 2001b, 2005), water- to the fine-grid reference case.
4.4 The Way Forward 145

Fig. 4.30 Gas injection patterns in a thin-bedded tidal reservoir modelled using a multi-scale method and
incorporating the effects of faults in the reservoir simulation model (From a study by Brandsæter et al 2001b)

Fig. 4.31 Effect of multi- 1.0 2500


scale upscaling on Fine Grid
estimates of oil rate and Coarse Grid (no upscaling)

Gas Oil Ratio (Sm3/Sm3)


GOR for the gas injection 0.8 Coarse Grid (one-step upscaling) 2000
(fraction of target rate)

case study shown in Coarse Grid (two-step upscaling)


Fig. 4.30 (Redrawn from
Pickup et al. 2000, #2000, 0.6 1500
Oil Rate

Society of Petroleum
Engineers Inc., reproduced
with permission of SPE. 0.4 1000
Further reproduction
prohibited without
permission) 0.2 500

0 0
0 0.2 0.4
Volume Gas Injected (fraction of Free GIP)

cases. The modelling methods have achieved


4.4 The Way Forward sufficient speed and reliability for routine imple-
mentation (generally using steady-state methods
4.4.1 Potential and Pitfalls on near-orthogonal corner-point grid systems).
However, a number of challenges remain which
Multi-scale reservoir modelling has moved from require further developments of methods and
a conceptual phase, with method development on modelling tools. In particular:
idealised problems, into a practical phase, with • Multi-scale modelling within a realistic struc-
more routine implementation on real reservoir tural geological grid is still a major challenge;
Handling Model Uncertainty
5

Abstract
The preceding chapters have highlighted a number of ways in which a
reservoir model can go right or wrong.
Nothing, however, compares in magnitude with the mishandling of
uncertainty. An incorrect saturation model, for example, can easily give a
volumetric error of 10 % and perhaps even 50 %. A flawed geological
concept could be much worse. Mishandling of uncertainty, however, can
result in the whole modelling and simulation effort becoming worthless.
The cause of this is occasionally misuse of software, more commonly it
is due to the limitations of our datasets, but primarily it is our behaviour
and our design choices which are at fault.
Our aim is to place our models within a framework that can overcome
data limitations and personal bias and give us a useful way of quantifying
forecast uncertainty.

P. Ringrose and M. Bentley, Reservoir Model Design, DOI 10.1007/978-94-007-5497-3_5, 151


# Springer Science+Business Media B.V. 2015
152 5 Handling Model Uncertainty

Did you expect to see the trees?

5.1 The Issue decision, one which has perhaps already been
made. In this case we are indeed simply
5.1.1 Modelling for Comfort ‘modelling for comfort’, a low value activity,
rather than taking the opportunity to use
In Chap. 1 we identified the tendency for modelling to identify a significant business risk.
modelling studies to become a panacea for deci-
sion making – modelling for comfort rather than
analytical rigour. It is certainly often the case that 5.1.2 Modelling to Illustrate
reservoir modelling is used to hide uncertainty Uncertainty
rather than illustrate it. We have a natural ten-
dency to determine a best guess – the anchoring Useful modelling can be expressed as ‘reason-
heuristic of Kahneman and Tverky (1974) – and able forecasting.’ A convenient metaphor for this
the management process in many companies is our ability to predict the image on a picture
often inadvertently encourages the guesswork. from a small number of sample points.
However, in a situation of dramatic under- We illustrate this, graphically, using sampled
sampling the guess is often wrong and influenced selections (Fig. 5.1) from a landscape photograph
unconsciously by behavioural biases of the (the chapter cover image). A routine modelling
individuals or teams involved (summarised in workflow would lead us to analyse and charac-
Kahneman 2011). Best-guess models therefore terise each sample point: the process of reservoir
tend to be misleading and their role is reduced characterisation. Data-led modelling with no
to one of providing comfort to support a business underlying concept and no application of trends
5.1 The Issue 153

Cloud (light) Cloud (light)


Rock (light)

Cloud (dark)

Rock (dark)
Rock (light)

Grass (heterogeneous)

House Grass (massive)

Grass (dark)

Fig. 5.1 An undersampled picture – our task is to determine the image

could produce the stochastic result shown in proportions, and in this sense all ‘match’ the
Fig. 5.2. This representation is statistically con- data, although with strongly contrasting textures.
sistent with the underlying data, and would pass a Assuming we then proceeded to add “colours”
simple QC test comparing frequency of for petrophysical properties (Chap. 3) and re-
occurrences in the data and in the model, yet scale the image for flow simulation (Chap. 4),
the result is clearly meaningless. these images would produce strongly contrasting
Application of logical deterministic trends to fluid-flow forecasts.
the modelling process, as described in Chap. 2, Using these different images as possible alter-
would make a better representation, one which native realisations could be one way of exploring
would at least fit an underlying landscape con- uncertainty, but we argue this would be a poor
cept: the sky is more likely to be at the top, the route to follow. Reference to the actual image
grass at the bottom (Fig. 5.3). Furthermore, there (Fig. 5.5) reveals a familiar theme:
is an anisotropy ratio we can use so that we can
predict better spatial correlation laterally (the sky data 6¼ model 6¼ truth
is more likely to extend across much of the
image, rather than up and down). If the texture Even though most aspects of the image were
from this trend-based approach is deemed unrep- sampled, and the applied deterministic trends
resentative of landscapes, an object-based alter- were reasonable, there are significant errors in
native may be preferred (Fig. 5.4). Grass is the representation – object modelling of the sky
accordingly arranged in clusters, broadly ellipti- was inappropriate, hierarchical organisation was
cal, as are sky colours (clouds) and the rocky missed, and even some aspects of the
areas are arranged into ‘hills’, anchored around characterisation (grass vs. rocks) were over-
the data points they were observed in. A rough simplified. There are also some modelling
representation is beginning to take shape. elements missing, most noticeably: there were
The model representations in Figs. 5.2, 5.3, no trees. Rearranging the data and detailed anal-
and 5.4 each adhere to the same element ysis of the original samples does not reveal the
154 5 Handling Model Uncertainty

Fig. 5.2 Stochastic model representation of the data in Fig. 5.1, assuming stationarity

Fig. 5.3 Overlay of deterministic trends on the stochastic model in Fig. 5.2, overcoming stationarity
Fig. 5.4 Object model alternative to Fig. 5.3, maintaining deterministic trends and embracing a loose alignment of
lozenge shapes

Fig. 5.5 Reality: the data set was unable to detect key missing elements, therefore these elements are also absent from
the simple probabilistic model, even with a useful deterministic trend imposed
156 5 Handling Model Uncertainty

missing elements. On reflection, we can see that alternative approaches to uncertainty handling,
the aim of reproducing the statistical content of and lead to a general recommendation for
the sample dataset brings with it a major flaw in scenario-based approaches, along the way also
all the models. distinguishing different flavours of ‘scenario’.
Could the missing elements in Fig. 5.5 have Scenario-based modelling became a popular
been foreseen, given that they were absent in the means of managing sub-surface uncertainty
data sample? We would argue yes, to a large during the 1990s, although opinions differ widely
extent. From the data set it is possible to establish on the nature of the ‘scenarios’ – particularly
a concept of hilly countryside in a temperate with reference to the relative roles of determin-
climate – the ‘expert judgement’ of Kahneman ism and probability. In the context of reservoir
and Klein (2009). Having established this, there modelling, a scenario is defined here as a possi-
are in fact certain aspects which are consistent ble real-world outcome and is one of several
with the concept but not actually seen by the possible outcomes (Bentley and Smith 2008).
sample data. However, these can be anticipated. The idea of alternative, discrete scenarios
Ask yourself: followed on logically from the emergence of
• Could there be more than one type of house? integrated reservoir modelling tools (e.g.
Yes. Cosentino 2001; Towler 2002), which
• Could there be a small village? Yes. emphasised the use of 3D static reservoir
• Is there a structure to the clouds? Yes. modelling, ideally fed from 3D seismic data
• Are the hills logically arranged, ones with and leading to 3D dynamic reservoir simulation,
greater contrast in the foreground? Yes generally on a full-field scale.
• Could there be trees? Appreciating the numerous uncertainties
Taking the issue of trees specifically, these are involved in constructing such field models, the
highly likely to be present, given the underlying desire for multiple modelling naturally arises.
concept (grass and hills in a temperate climate). Although not universal (see discussion in
They are also likely to be under-sampled. Dubrule and Damsleth 2001), the application of
The parallels with reservoir modelling are multiple stochastic modelling techniques is now
hopefully clear: we need to use concepts to widespread, with the alternative models
honour the data but work beyond it to include described variously as ‘runs’, ‘cases’,
missing elements. If these elements are important ‘realisations’ or ‘scenarios’.
to the field development (e.g. open natural The different terminologies are more than
fractures, discontinuous but high permeability semantic. The notion of multiple modelling has
layers, cemented areas, sealing sub-seismic been explored differently by different workers,
faults, thin shales) then the presence or absence the essential variable being the balance between
of these features becomes the important uncer- deterministic and probabilistic inputs. Using
tainty. We should always ask ourselves: “could “multiple realisations” may sound more routed
there be trees?” in statistical theory than using some alternative
“model runs” – but is it? These concepts are best
related to differing approaches to the application
of geostatistical algorithms, and to differing
5.2 Differing Approaches ideas on the role of the probabilistic component
(Fig. 5.6).
Abandoning the route of modelling for comfort The contrasting approaches to uncertainty
and embarking on the harder but more interesting handling broadly fall into three groups:
and ultimately more useful route of modelling Rationalist approaches, in which a preferred
to illustrate uncertainty, we need a workflow model is chosen as a base case (Fig. 5.7).
(see Caers 2011, for a summary of statisti- The model is either run as a technical best
cal methodologies). This chapter will review guess, or with a range of uncertainty added
5.2 Differing Approaches 157

Fig. 5.6 Alternative Best Guess


approaches to uncertainty Anchored on a preferred ‘base case’
handling

BG

MS MD
Multiple Stochastic Multiple Deterministic
Models selected by building Models designed manually based on
‘equiprobable’ realisations from a base discrete alternative concepts
case model

a BG b
BG

MS MD MS MD

Concept Concept

% - Base case outcome + %

% - Base case outcome + % low case base case high case

Fig. 5.7 Base case–dominated, rationalist approaches (Redrawn from Bentley and Smith 2008, The Geological
Society, London, Special Publications 309 # Geological Society of London [2008])

to that guess. This may be either a percentage choice of the boundary conditions for the
factor in terms of the model output (e.g. simulations, such as assumed correlation
20 % of the base case volumes in-place) or lengths. Yarus and Chambers (1994) give sev-
separate low and high cases flanking the base eral examples of this approach, and the
case. This approach can be viewed as ‘tradi- options and choices are reviewed by Caers
tional’ determinism. (2011).
Multiple stochastic approaches, in which a large Multiple deterministic approaches, which avoid
number of models are probabilistically making a single best-guess or choosing a
generated by geostatistical simulation preferred base-case model (Fig. 5.9). In this
(Fig. 5.8). The deterministic input lies in the approach a smaller number of models
158 5 Handling Model Uncertainty

Fig. 5.8 Multiple


stochastic approaches
(Redrawn from Bentley
and Smith 2008, The
Geological Society,
London, Special
Publications 309 #
Geological Society
of London [2008])

Fig. 5.9 Multiple-deterministic, ‘scenario-based’ approach


5.3 Anchoring 159

are built, each one reflecting a complete It is often stated that for mature fields, a
real-world outcome following an explicitly- simple, rationalist approach may suffice because
defined reservoir concept. Geostatistical sim- uncertainty has reduced through the field life
ulation may be applied in the building of the cycle. This is a fallacy. Although, the magnitude
3D model but the selection of the model of the initial development uncertainties tends to
realisations is made manually (or mathemati- decrease with time, we generally find that as the
cally) rather than by statistical simulation field life cycle progresses new, more subtle,
(e.g. van de Leemput et al. 1996). uncertainties arise and these now drive the deci-
Each of the above approaches have been sion making. For example, in the landscape
referred to as ‘scenario modelling’ by different image in Fig. 5.5, 100 samples would signifi-
reservoir modellers. The argument we develop cantly improve the ability to describe the
here is that although all three approaches have image, but this is still insufficient to specify the
some application in subsurface modelling, location of an unsampled house. The impact of
multiple-deterministic scenario-building is the uncertainties in terms of their ability to erode
preferred route in most circumstances. value may, in fact, be as great near the end of
In order to make this case, we need to recall the field life as at the beginning.
the underlying philosophy of uncertainty Despite this, rationalist, base-case modelling
handling and give a definition for ‘scenario remains common across the industry. In a review
modelling’. of 90 modelling studies conducted by the authors
and colleagues across many companies, field
modelling was based on a single, best-guess
model in 36 % of the cases (Smith et al. 2005).
5.3 Anchoring This was the case, despite a bias in the sampling
from the authors’ own studies, which tended to
5.3.1 The Limits of Rationalism be scenario-based. Excluding the cases where the
model design was made by the authors, the pro-
The rationalist approach, described above as the portion of base case-only models rose to 60 %.
‘best-guess’ method, is effectively simple
forecasting – and puts faith in the ability of an
individual or team to make a reasonably precise 5.3.2 Anchoring and the Limits
judgement. If presented as the best judgement of Geostatistics
of a group of professional experts then this
appears reasonable. The weak point is that the The process of selecting a best guess in spite of
best guess is only reliable when the system wide uncertainty is referred to as ‘anchoring’,
being described is well ordered and well under- and is a well-understood cognitive behaviour
stood, to the point of being highly predictable (Kahneman and Tverky 1974). Once anchored,
(Mintzberg 1990). It must be assumed that the adjustment away from the initial best guess is
enough data is available from past activities to too limited as the outcome is overly influenced
predict a future outcome with confidence, and by the anchor point.
this applies equally to production forecasting, This often also occurs in statistical approaches
exploration risking, volumetrics or well to uncertainty handling, as these tend to be
prognoses. Although this is rarely the case in anchored in the available data and may therefore
the subsurface, except perhaps for fields with a make the same rational starting assumption as the
large (100+) number of regularly spaced wells, simple forecast, although adding ranges around a
there is a strong tendency for individuals or ‘most probable’ prediction (see examples in
teams (or managers) to desire a best guess, Chellingsworth et al. 2011).
and to subsequently place too much confidence Geostatistical simulation allows definition
in that guess (Baddeley et al. 2004). of ranges for variables, followed by rigorous
160 5 Handling Model Uncertainty

sampling and combination of parameters to yield the building of multiple, deterministically-driven


a range of results, which can be interpreted models of plausible development outcomes
probabilistically. If the input data can be Each scenario is a complete and internally
specified accurately, and if the combination pro- consistent static/dynamic subsurface realisation
cess maintains a realistic relationship between all with an associated plan tailored to optimise its
variables, the outcome may be reasonable. In development. In an individual subsurface sce-
practice, however, input data is imperfectly nario, there is clear linkage between technical
defined and the ‘reasonableness’ of the detail in a model, and an ultimate commercial
automated combination of variables is hard to outcome; a change in any element of the model
verify. Statistical rigour is applied to data sets prompts a quantitative change in the outcome
which are not necessarily statistically significant and the dependency between all parameters in
and an apparently exhaustive analysis may the chain (between the changed element and the
have been conducted on insufficient data. outcome) is unbroken.
The validity of the outcome may also be This contrasts with many probabilistic
weakened by centre-weighting of the input data simulations, in which model design parameters
to variable-by-variable best-guesses, which are statistically sampled and cross-multiplied,
creates an inevitability that the ‘most likely’ and in which dependencies between variables
probabilistic outcome will be close to the initial are either lost, or collapsed into correlation
best guess (Wynn and Stephens 2013). The coefficients.
geostatistical simulation itself is thus ‘anchored’. The scenario approach therefore places a
It is therefore argued that the application of strong emphasis on deterministic representation
geostatistical simulation does not in itself com- of a subsurface concept: geological, geophys-
pensate for a natural tendency towards a rational- ical, petrophysical and dynamic. Without a
ist best guess – it often tends to simply reflect it. clearly defined concept of the subsurface –
The crucial step is to select a workflow which clear in the sense that a geoscientist could rep-
removes the opportunity for anchoring on a resent it as a simple sketch – the modelling
best guess; and this is what scenario modelling, cannot progress meaningfully. We have used
as defined here, attempts to achieve. the mantra: if you can sketch it, you can model
it. Geostatistical simulation may be a key tool
required to build an individual scenario but
5.4 Scenarios Defined the design of the scenarios is determined
directly by the modeller. Multiple models are
The definition of ‘scenario’ adopted here follows based on multiple, deterministic designs. This
that described by van der Heijden (1996), who distinguishes the workflows for scenario
discussed the use of scenarios in the context of modelling, as defined here, from multiple sto-
corporate strategic planning and defined chastic modelling which tends to be based on
scenarios as a set of reasonably plausible, but statistical sampling from a single initial design.
structurally different futures. Note that multiple stochastic modelling is a
Alternative scenarios are not incrementally powerful tool for understanding reservoir
different models based on slight changes in con- model ranges and outcomes; it is simply not
tinuous input data (as with multiple probabilistic sufficient to fully explore subsurface uncer-
models), but models which are structurally dis- tainties from poorly sampled reservoirs.
tinct based on some defined design criteria. Scenario-based approaches place an emphasis
Translated to oil and gas field development, a on listing and ranking of uncertainties, from
‘scenario’ is therefore a plausible development which a suite of scenarios will be built, with no
outcome, and the ‘scenario-based approach’ to attempt being made to select a best guess case
reservoir modelling is defined as: up-front.
5.6 Applications 161

uncertainty of that factor with discrete cases,


5.5 The Uncertainty List rather than simply input a data distribution for a
higher level parameter such as net-to-gross.
The key to success in scenario modelling lies in
deriving an appropriate list of key uncertainties,
a matter of experience and judgement. However,
there is a strong tendency to conceptualise key 5.6 Applications
uncertainties for at least the static reservoir
models in terms of the parameters of the STOIIP 5.6.1 Greenfield Case
equation, i.e. when asked to define the key
uncertainties in the field, modellers will often The application of scenario modelling has been
quote parameters such as ‘porosity’ or ‘net most successfully reported for cases involving
sand’ as key factors. If the model-build new or ‘greenfield’ reservoir studies.
progresses with these as the key uncertainties to One of the first published examples was that
alter, this will most likely be represented as a of van de Leemput et al. (1996), who described
range for a continuous variable, anchored around an application of scenario-based modelling in the
a best guess. context of an LNG field development. Once suf-
A better approach is to question why ‘poros- ficient proven volumes were established to sup-
ity’ or ‘net sand count’ are considered significant port the scheme, the commercial structure of the
uncertainties. It will either emerge that the uncer- project focussed attention on the issue of
tainty is not that significant or, if it is, then it the associated capital expenditure (CAPEX).
relates to some underlying root cause, such as CAPEX therefore became the prime quantitative
heterogeneous diagenesis, or some local facies outcome of the modelling exercise, driven
control which has not been extracted from the largely by well numbers and the requirements
data analysis. for, and timing of, gas compression facilities.
For example, in Fig. 5.10 a PDF of net-to- The model scenarios were driven by a
gross is shown. A superficial approach to model selection of principal uncertainties, summarised
uncertainty would involve taking the PDF, input- in Fig. 5.11. Six static and five dynamic
ting it to a geostatistical algorithm and allowing uncertainties were drawn up, based on the judge-
sampling of the spread to account for the uncer- ment of the project team and input from peers.
tainty. As the data in the figure illustrates, this Maintaining the uncertainty list became a
would be misleading, because the range is continuing process, iterating with new well data
reflecting mixed geological concepts. The real from appraisal drilling, and the changing views
need is to understand the facies distribution, and of the group.
isolate the facies-based factors (in this case the For the field development plan itself, the
proportion of different channel types), and then uncertainty list generated 22 discrete scenarios,
establish whether this ratio is known within rea- each of which was matched to the small amount
sonable bounds. If not known, the uncertainty of production data, then individually tailored to
can be represented by building contrasting, but optimise the development outcome over the life
realistic, depositional models (the basis for two of the LNG scheme. The outcomes, in term of
scenarios) in which these elements are specifi- impact on project cost (CAPEX), are shown in
cally contrasted. The uncertainty in the net-to- Fig. 5.11.
gross parameter within each scenario is a second- A key learning outcome from this exercise
order issue to the geological uncertainty. was that a list of 11 uncertainties was unneces-
In defining key uncertainties, the need is sarily long to generate the ultimate result,
therefore to chase the source of the uncertainty although convenient for satisfying concerns of
to the underlying causative factor – ‘root cause stakeholders. The effect of statistical dominance
analysis’ – and model the conceptual range of meant that the range was not driven by all
162 5 Handling Model Uncertainty

Fig. 5.10 Root-cause


analysis: defining the
underlying causative
uncertainty (Redrawn from
Bentley and Smith 2008, wide
The Geological Society, f uncertainty
London, Special range
Publications 309 #
Geological Society of
London [2008])

net-to-gross

narrow
f uncertainty
range
within-
case

11 uncertainties, but by 2 or 3 key uncertainties normally addressed by modelling: sand body


to which the scheme was particularly sensitive. geometries, relative permeabilities, aquifer size
Contrary to the expectations of geoscientists, etc., were certainly poorly understood, but could
gross rock volume on the structures was not a key be shown to have no significant impact on the
development issue, even though the fields were field development decision. In hindsight, the
large and each had only two or three well dominant issues were foreseeable without
penetrations at the time of the field development modelling.
plan (FDP) submission. The key issue was the In the light of the above, continued post-FDP
potential enhancements of well deliverability modelling became more focussed, with a smaller
offered by massive hydraulic fracturing – not a number of scenarios fleshing out the dominant
factor typically at the heart of reservoir issues only. Tertiary issues were effectively
modelling studies. The majority of the issues treated as constants.
5.6 Applications 163

Project cost $
Static
Structure

Reservoir Properties

Sand Connectivity

‘Thief’ Zones

Fault Compartments

Dynamic
Relative Permeabilities
Aquifer Behaviour
WellProductivity (Hydraulic Fractures,
Condensate Drop-out, WellType)
Fluid Composition
- changes +

Fig. 5.11 Application of deterministic scenarios to a green field case: forecasting costs (Redrawn from Bentley and
Smith 2008, The Geological Society, London, Special Publications 309 # Geological Society of London [2008])

The above example was conducted without 5.6.2 Brownfield Case


selecting a ‘base case’ model. A development
scheme was ultimately selected, but this was Two published examples are summarised here
based on a range of outcomes defined by the which illustrate the extension of scenario
subsurface team. modelling to mature, or ‘brownfield’, reservoir
Scenario modelling for greenfields has been cases.
conducted many times since the publication of The first concerns the case of the Sirikit Field
this example. In the experience of the authors, in Thailand (Bentley and Woodhead 1998). The
the early learnings described in the case above requirement was to review the field mid-life and
have held true, notably: evaluate the potential benefit of introducing
• Large numbers of scenarios are not required to water injection to the field. At that point the
capture the range of uncertainty; field had been on production for 15 years, with
• The main uncertainties can generally be 80 wells producing from a stacked interval of
drawn up through cross-discipline discussion partially-connected sands. The required outcome
prior to modelling – if not these can be was a quantification of the economic benefit of
established by running quick sensitivities; water injection, to which a scenario–based
• This list should be checked and iterated as the approach was to be applied.
modelling progresses; The uncertainty list is summarised in
• The dominant uncertainties on a development Fig. 5.12. The static uncertainties were used to
project do not always include the issue of generate the suite of static reservoir models for
seismically driven gross rock volume, even input to simulation. In contrast to the greenfield
at the pre-development phase; cases, where production data is limited, the
• It is not necessary to select a base case model. dynamic uncertainties were used as the history
164 5 Handling Model Uncertainty

70
Static
forecast Heterolithics No Channels
Cumulative oil (MMbbls)

60

50
High Connectivity Small Bodies
40
Low Connectivity Body Orientation
30
history time of study
20
Dynamic
10
Condensate/Gas Ratio
0
0 1 2 3 4 5 6 7 8 9
Relative Permeabilities
Time (years)

Aquifer Size

wide incremental uncertainty


Fig. 5.12 Application of deterministic scenarios to a brownfield case: forecasting production (Redrawn from Bentley
and Smith 2008, The Geological Society, London, Special Publications 309 # Geological Society of London [2008])

matching tools – the permissible parameter and all were plausible. A base case was not
ranges for those uncertainties being established chosen.
before the matching began. The outcome makes a strong statement about
A compiled production forecast for the ‘no the non-uniqueness of simulation model
further drilling case’ is shown in Fig. 5.12. The matches. If a base case model had been
difference between that spread of outcomes and rationalised based on preferred guesses, any of
the spread from a parallel set of outcomes which the seven scenarios could feasibly have been
included water injection, was used to quantify chosen – only by chance would the eventual
the value of the injection decision. Of interest median model have been selected.
here is the nature of that spread. Although all The Sirikit case also confirmed that multiple
models gave reasonable matches to history, the deterministic modelling was achievable in rea-
incremental difference between the forecasts was sonable study times – scaled sector models were
larger than that expected by the team. It was used to ease the handling of production data (see
hoped that some of the static uncertainties Bentley and Woodhead 1998). The workflow
would simply be ruled out by the matching pro- yielded a surprisingly wide range of model
cess. Ultimately none were, despite 80 wells and forecasts.
15 years of production history. Fig. 5.13 summarises an application of sce-
The outlier cases were reasonable model nario modelling to a producing field with 4D
representations of the subsurface, none of the seismic, which generated additional insights
scenarios was strongly preferred over any other, into the use of scenarios. The case is from the
5.7 Scenario Modelling – Benefits 165

outcome – a range of water-cut breakthrough


Time to water break through times – is illustrated in Fig. 5.13.
(year 0 = study time)
The Gannet B study offered some additional
Well 1 insights into mature field scenario modelling:
• Although the truism is offered that multiple
models can match production data (there is no
uniqueness to history matches), the converse
is not necessarily true – everything cannot
always be matched;
• The above is more likely to be true in smaller
fields, where physical field limitations con-
Well 2 strain possible scenarios;
• In the specific case of Gannet B, the principal
matching tool was 4D seismic data (not well
production data), and it was the seismic which
was the matching target for the multiple
model scenarios;
• A base case selection from the quantified range
in water breakthrough times would have been
1 2 3 4 highly misleading; “between 9 months and 4
year years” was the answer to the question based on
the available data. Making a median guess
Static Dynamic would have simply hidden the risk.
Faulted case Aquifer

Sandy case
5.7 Scenario Modelling – Benefits
Muddy case

High volume case


The scenario-based approach as defined here
offers specific advantages over base case
modelling and multiple probabilistic modelling:
Fig. 5.13 Application of deterministic scenarios to a Determinism: the dominance of the underlying
brownfield case: forecasting water breakthrough
conceptual reservoir model, which is deter-
(Redrawn from Bentley and Smith 2008, The Geological
Society, London, Special Publications 309 # Geological ministically applied via the model design.
Society of London [2008]) Although the models may use any required
level of geostatistical simulation to re-create
Gannet B Field in the Central North Sea (Bentley the desired reservoir concept, the
and Hartung 2001; Kloosterman et al. 2003). geostatistical algorithms are not used to select
The issue to model in the Gannet B case was the cases to be run, nor to quantify the uncer-
the risk and timing of potential water break- tainty ranges in the model outcomes.
through in one of the field’s two gas producers, Lack of anchoring: the approach is not built on
and placing value on the possible contingent the selection of a base case, or best guess.
activities post-breakthrough. As with the cases Qualitatively, the natural tendency to under-
above, the study started with a listing and quali- estimate uncertainties is less prone to occur if
tative ranking of principal uncertainties in a a best guess is not required – the focus lies
cross-discipline forum. Unlike the previous instead on an exploration of the range.
cases, it proved not to be possible to match all Dependence: direct dependence between para-
static reservoir models with history. The lowest meters is maintained through the modelling
volume realisation would not match. The model process; a contrast between two model
166 5 Handling Model Uncertainty

realisations is fed through directly to two quan- achieved with a smaller number of scenarios, but
titative scenarios, which allow the significance the full set was run simply because it was not
of the uncertainty to be evaluated. particularly time-consuming (the whole study ran
Transparency: although the models may be inter- over roughly 5 man weeks, including static and
nally complex, the workflow is simple, and dynamic modelling). The case illustrates
feeds directly off the uncertainty list, which the efficacy of multiple static/dynamic modelling
may be no more complex than a short list of in greenfields, even when the compilation of runs
the key issues which drive the decision at hand. is manual. Figure 5.14b shows the results of a more
If the key issues which could cause a project to recent study (Chellingsworth, et al. 2011) in
fail are identified on that list, the model process which 124 STOIIP-related cases were efficiently
will evaluate the outcome in the result range. analysed using a workflow-manager algorithm.
The focus is therefore not on the intricacies of This issue is more pressing for brownfield sites,
the model build (which can be reviewed by an although the cases described above from the Sirikit
expert, as required), but on the uncertainty list, and Gannet fields illustrate that workflows for
which is transparent to all interested parties. multiple model handling in mature fields can be
practical. This challenge is also being improved
further by the emergence of a new breed of auto-
5.8 Multiple Model Handling matic history matching tools which achieve model
results according to input guidelines which can be
It is generally assumed that more effort will be deterministically controlled.
required to manage multiple models than a single It is thus suggested that the running of multi-
model, particularly when brownfield sites require ple models is not a barrier to scenario modelling,
multiple history matching. However, this is not even in fields with long production histories.
necessarily the case – it all comes down to a Once the conceptual scenarios have been clearly
choice of workflow. defined, it often emerges that complex models
Multiple model handling in greenfield sites are not required. Fit-for-purpose models also
is not necessarily a time-consuming process. come with a significant time-saving.
Figure 5.14a illustrates results from a study Cross-company reviews by the authors
involving discrete development scenarios. indicate that model-building exercises which
These were manually constructed from are particularly lengthy are typically those
permutations of 6 underlying static models and where a very large, detailed, base-case model is
dynamic uncertainties in fluid distribution and under construction. History matching is often
composition. This was an exhaustive approach pursued to a level of precision disproportionate
in which all combinations of key uncertainties to the accuracy of the static reservoir model it
were assessed. The final result could have been is based on. By contrast, multiple modelling

100% 100%
‘P90’ ‘P90’

50% ‘P50’ 50% ‘P50’

324 static ‘P10’ 124 static & dynamic ‘P10’


models models
0 0
0 100 200 mmstb 0 500 1000 mmstb

Fig. 5.14 Multiple deterministic cases for STOIIP (left) and ultimate recovery (right)
5.9 Linking Deterministic Models with Probabilistic Reporting 167

exercises tend to be more focussed and, paradox- The type of design depends on the purpose of
ically, tend to be quicker to execute than the very the study and on the degree of interaction between
large, very detailed base-case model builds. the variables. A simple approach is the Plackett-
Burmann formulation. This design assumes that
there are no interactions between the variables
and that a relatively small number of experiments
5.9 Linking Deterministic Models
are sufficient to approximate the behaviour of the
with Probabilistic Reporting
system. More elaborate designs, for example D-
optimal or Box-Behnken (e.g. Alessio et al. 2005;
The next question is how to link multiple-
Cheong and Gupta 2005; Peng and Gupta 2005),
deterministic scenarios with a probabilistic
attempt to analyse different orders of interaction
framework? Ultimately we wish to know how
between the uncertainties and require a signifi-
likely an outcome is. In reservoir modelling,
cantly greater number of experiments. The value
probability is most commonly summarised as
of elaboration in the design needs to be assessed –
the percentiles of the cumulative probability dis-
more is not always better – and depends on the
tribution – P90, P50, and P10, where P90 is the
model purpose, but the principles described below
value (e.g. reserves) which has a 90 % probabil-
apply generally.
ity of being exceeded, and P50 is the median of
A key aspect of experimental design is that the
the distribution. With multiple-deterministic
uncertainties can be expressed as end-members.
scenarios, as each scenario is qualitatively
The emphasis on making a base case or a best
defined, the link to statistical descriptions of the
guess for any variable is reduced, and can be
model outcome (e.g. P90, P50 and P10) can be
removed.
qualitative (e.g. a visual ranking of outcomes) or
The combination of Plackett-Burmann exper-
formalised in a more quantitative manner.
imental design with the scenario-based approach
An important development has been the merg-
is illustrated by the case below from a mature
ing of deterministically-defined scenario models
field re-development plan involving multiple-
with probabilistic reporting using a collection of
deterministic scenario-based reservoir modelling
approaches broadly described as ‘experimental
and simulation (Bentley and Smith 2008). The
design’. This methodology offers a way of
purpose of the modelling was to build a series of
generating probabilistic distributions of
history-matched models that could be used as
hydrocarbons in place or reserves from a limited
screening tools for a field development.
number of deterministic scenarios, and of relating
As with all scenario-based approaches, the
individual scenarios to specific positions on the
workflow started with a listing of the uncertainties
cumulative probability function (or ‘S’ curve). In
(Fig. 5.15), presumed in this case to be:
turn, this provides a rationale for selecting specific
models for screening development options.
Experimental design is a well-established tech-
nique in the physical and engineering sciences Structure
where it has been used for several decades (e.g.
Box and Hunter 1957). It has more recently Thin Beds
become popular in reservoir modelling and simula-
tion (e.g. Egeland et al. 1992; Yeten et al. 2005; Li Reservoir Quality
and Friedman 2005) and offers a methodology for
Contacts
planning experiments so as to extract the maximum
amount of information about a system using the Architecture
minimum number of experimental runs. In subsur-
face modelling, this can be achieved by making a Body Orientation
series of reservoir models which combine
uncertainties in ways specified by a theoretical
template or design. Fig. 5.15 Experimental design case: uncertainty list
168 5 Handling Model Uncertainty

Fig. 5.16 Alternative reservoir architectures (Images courtesy of Simon Smith) (Redrawn from Bentley and Smith
2008, The Geological Society, London, Special Publications 309 # Geological Society of London [2008])

Realisation Structure Quality Contacts Architecture Thin beds Orientation Response


1 -1 1 1 1 -1 1 1178
2 -1 -1 1 1 1 -1 380
3 -1 -1 -1 -1 -1 -1 109
4 1 -1 1 1 -1 1 1105
5 -1 -1 -1 1 1 1 402
6 1 -1 1 -1 -1 -1 1078
7 1 1 -1 1 1 -1 1176
8 1 -1 -1 -1 1 1 1090
9 -1 1 -1 -1 -1 1 870
0 -1 1 1 -1 1 -1 932
11 1 1 -1 1 -1 -1 1201
12 1 1 1 -1 1 1 1245
13 0 0 0 0 0 0 956
14 1 1 1 1 1 1 1656

Fig. 5.17 Plackett-Burmann matrix showing high/low combinations of model uncertainties and the resulting response
(resource volumes in Bscf)

1. Top reservoir structure; caused by poor qual- A model was built for each, with no preferred
ity seismic and ambiguous depth conversion. case.
This was modelled using alternative structural 4. Sand quality; this is an uncertainty simply
cases capturing plausible end-members. because of the limited number of wells and
2. Thin-beds; the contribution of intervals of was handled by defining alternative cases for
thin-bedded heterolithics was uncertain as facies proportions, the range guided by the
these intervals had not been produced or tested best and worst sand quality seen in wells.
in isolation. This uncertainty was modelled by 5. Reservoir orientation; modelled using alter-
generating alternative net-to-gross logs. native orientations of the palaeodip.
3. Reservoir architecture; uncertainty in the 6. Fluid contacts; modelled using plausible end-
interpretation of the depositional model was members for fluid contacts.
expressed using three conceptual models: tidal These six uncertainties were combined using
estuarine, proximal tidal-influenced delta and a 12-run Plackett-Burmann design. The way in
distal tidal-influenced delta models (Fig. 5.16). which the uncertainties were combined is shown
5.9 Linking Deterministic Models with Probabilistic Reporting 169

Distributions assigned to each term prior to running 10,000 simulations


Distributions
structure quality

-1.00 -0.50 -0.00 0.50 1.00 -1.00 -0.50 -0.00 0.50 1.00

thin-beds contacts

-1.00 -0.50 -0.00 0.50 1.00 -1.00 -0.50 -0.00 0.50 1.00
0.500 0.600
architecture orientation
0.375 0.450
0.250 0.300
0.125 0.150
0.000 0.000
-1.00 -0.50 -0.00 0.50 1.00 -1.00 -0.50 -0.00 0.50 1.00

Fig. 5.18 Parameter ranges and distribution shapes for each uncertainty

in the matrix in Fig. 5.17, in which the high case the underlying conceptual model, and requires
scenario is represented by +1, the low case by 1 the definition of a parameter distribution function
and a mid case by 0. In this case two additional (e.g. uniform, Gaussian, triangular). The distri-
runs were added, one using all the mid points and bution shapes selected for each uncertainty in this
one using all the low values. Neither of these two case are shown in Fig. 5.18. For variables where
cases is strictly necessary but can be useful to the value can be anywhere between the 1 and 1
help understand the relationship between the end members, a uniform distribution is appropri-
uncertainties and the ultimate modelled outcome. ate, for those with a central tendency a normal
The 14 models were built and the resource vol- distribution is preferred (simplified as a triangular
ume (the ‘response’) determined for each reservoir. distribution) and for some variables only discrete
A linear least-squares function was derived from alternative possibilities were chosen.
the results, capturing the relationship between the Once the design is set up, and assuming the
response and the individual uncertainties. The rela- independence of the chosen variables is still
tive impact of the individual uncertainties on the valid, the distributions can then be sampled by
resource volumes is captured by a co-efficient spe- standard Monte-Carlo analysis to generate a
cific to the impact of each uncertainty. probabilistic distribution. The existing suite of
The next step in the workflow is to consider models can then be mapped onto a probabilistic,
the likelihood of each uncertainty occurring in or S-curve, distribution (Fig. 5.19).
between the defined end-member cases, that is, in There are three distinct advantages to using
between the ‘1’ and the ‘1’. This relates back to this workflow. Firstly, it makes a link between
170 5 Handling Model Uncertainty

Moreover, having conducted an experimental


1.00
design, it may emerge that the P50 outcome is
0.75 significantly different from the previously
0.50 assumed initial ‘best guess.’ That is, this uncer-
0.25 tainty modelling approach can help compensate
for the biases that the user, or subsurface team,
0.00
started with.
1500 1600 1700 1800 1900
percentiles
• P90 1614, P50 1693, P10 1785 bcf
• P99 1503 and P1 1900 bcf 5.10 Scenarios and
Uncertainty-Handling
1500 GIIP main 1900

structure Scenario-based approaches offer an improvement


over base-case modelling, as results from the lat-
quality ter are anchored around best guess assumptions.
Best guesses are invariably misleading because
contacts data from the subsurface is generally insufficient
to be directly predictive. Scenarios are defined
architecture here as ‘multiple, deterministically-driven models
of plausible development outcomes’, and are pre-
thin-beds
ferred to multiple stochastic modelling alone, the
application of which is limited by the same data
orientation
insufficiency which limits base case modelling.
Each scenario is a plausible development future
Fig. 5.19 Probabilistic volumes from Monte Carlo sim-
based on a specific concept of the subsurface, the
ulation of the experimental design formulation development planning response to which can be
optimised.
probabilistic reporting and discrete multiple- The application of geostatistical techniques, and
deterministic models. This can be used to pro- conditional simulation algorithms in particular, is
vide a rationale for selecting models for simula- wholly supported as a means of completing a real-
tion. For example, P90, P50 and P10 models can istic subsurface model – usually by infilling a
be identified from this analysis and it may strongly deterministic model framework. Multiple
emerge that models reasonably close to these stochastic modelling can also be useful to explore
probability thresholds were built as part of the sensitivities around an individual deterministic sce-
initial experimental design. Alternatively, the nario. Deterministic design of each over-arching
comparison may show that new models need to scenario, however, is preferred because of transpar-
be built. This is easier to do now that the impact ency, relative simplicity and because each scenario
of the different uncertainties has been quantified, can be validated as a realistic subsurface outcome.
and is an improvement on an arbitrary assump- Scenario-based modelling is readily applica-
tion that a high case model, for example, ble to greenfield sites but, as the examples shown
represents the P10 case. Secondly, the workflow here confirm, is also practical for mature, brown-
focuses on the end-members and on capturing the field sites, where multiple history matching may
range of input variables, avoiding the need to be required at the simulation stage.
anchor erroneously on a best guess. Finally, the The key to success is the formulation of the
approach provides a way of quantifying the uncertainty list. If the issues which could cause
impact of the different uncertainties via tornado the business decision to fail are identified, then
diagrams or simple spider plots, which can in turn the modelling workflow will capture this and the
be used to steer further data gathering in a field. decision risk can be mitigated. If the issue is
References 171

Fig. 5.20 Did you


anticipate the trees?

missed, no amount of modelling of any kind can and herding. In: Curtis A, Wood R (eds) Geological
compensate. The list is therefore central, includ- prior information. Informing science and engineering,
The Geological Society special publications, 239. The
ing the identification of issues not explicit in the Geological Society, London, pp 15–27
current data set, but which can be anticipated with Bentley MR, Hartung M (2001) A 4D surprise at Gannet
thought. Remember, there may be trees B. EAGE annual technical conference, Amsterdam
(Fig. 5.20). (extended abstract)
Bentley M, Smith S (2008) Scenario-based reservoir
modelling: the need for more determinism and less
anchoring. In: Robinson A et al (eds) The future of
geological modelling in hydrocarbon development,
References The Geological Society special publications, 309.
The Geological Society, London, pp 145–159
Alessio L, Bourdon L, Coca S (2005) Experimental Bentley MR, Woodhead TJ (1998) Uncertainty handling
design as a framework for multiple realisation history through scenario-based reservoir modelling. SPE
matching: F6 further development studies. SPE 93164 paper 39717 presented at the SPE Asia Pacific confer-
presented at SPE Asia Pacific Oil and Gas conference ence on Integrated Modelling for Asset Management,
and exhibition, Jakarta, 5–7 Apr 2005 23–24 Mar 1998, Kuala Lumpur
Baddeley MC, Curtis A, Wood R (2004) An introduction Box GEP, Hunter JS (1957) Multifactor experimental
to prior information derived from probabilistic designs for exploring response surfaces. Ann Math
judgements: elicitation of knowledge, cognitive bias Stat 28:195–241
172 5 Handling Model Uncertainty

Caers J (2011) Modeling uncertainty in the earth sciences. Simulation symposium, 31 Jan–2 Feb, The
Wiley (published online) Woodlands, 2005
Chellingsworth L, Kane P (2013) Expectation analysis in Mintzberg H (1990) The design school, reconsidering
the assessment of volume ranges in appraisal and the basic premises of strategic management. Strateg
development – a case study (abstract). Presented at Manage J 6:171–195
the Geological Society conference on Capturing Peng CY, Gupta R (2005) Experimental design and
Uncertainty in Geomodels – Best Practices and analysis methods in multiple deterministic
Pitfalls, Aberdeen, 11–12 Dec 2013. The Geological modelling for quantifying hydrocarbon in place prob-
Society, London ability distribution curve. SPE paper 87002 presented
Chellingsworth L, Bentley M, Kane P, Milne K, at SPE Asia Pacific conference on Integrated
Rowbotham P (2011) Human limitations on hydrocar- Modelling for Asset Management, Kuala Lumpur,
bon resource estimates – why we make mistakes in 29–30 Mar 2004
data rooms. First Break 29(4):49–57 Smith S, Bentley MR, Southwood DA, Wynn TJ, Spence
Cheong YP, Gupta R (2005) Experimental design and A (2005) Why reservoir models so often disappoint –
analysis methods for assessing volumetric some lessons learned. Petroleum Studies Group
uncertainties. SPE J 10(3):324–335 meeting, Geological Society, London. Abstract.
Cosentino L (2001) Integrated reservoir studies. Editions Towler BF (2002) Fundamental principles of reservoir
Technip, Paris, 310 p engineering, vol 8, SPE textbook series. Henry L.
Dubrule O, Damsleth E (2001) Achievements and Doherty Memorial Fund of AIME, Society of Petro-
challenges in petroleum geostatistics. Petroleum leum Engineers, Richardson
Geosci 7:S53–S64 van de Leemput LEC, Bertram D, Bentley MR, Gelling R
Egeland T, Hatlebakk E, Holden L, Larsen EA (1992) (1996) Full-field reservoir modeling of Central
Designing better decisions. SPE paper 24275 Oman gas/condensate fields. SPE Reservoir Eng
presented at SPE European Petroleum Computer 11(4):252–259
conference, Stavanger, 25–27 May 1992. van der Heijden K (1996) Scenarios: the art of strategic
Kahneman D (2011) Thinking fast and slow. Farrar, conversation. Wiley, New York, 305pp
Straus and Giroux, New York, 499 p Wynn T, Stephens E (2013) Data constraints on reservoir
Kahneman D, Klein G (2009) Conditions for intuitive concepts and model design (abstract). Presented at the
expertise: a failure to disagree. Am Psycol 64:515–526 Geological Society conference on Capturing Uncer-
Kahneman D, Tverky A (1974) Judgement under uncer- tainty in Geomodels – Best Practices and Pitfalls,
tainty: heuristics and biases. Science 185:1124–1131 Aberdeen, 11–12 Dec 2013
Kloosterman HJ, Kelly RS, Stammeijer J, Hartung M, van Yarus JM, Chambers RL (1994) Stochastic modeling and
Waarde J, Chajecki C (2003) Successful application of geostatistics principals, methods, and case studies.
time-lapse seismic data in Shell Expro’s Gannet AAPG Comput Appl Geol 3:379
Fields, Central North Sea, UKCS. Petroleum Geosci Yeten B, Castellini A, Guyaguler B, Chen WH (2005) A
9:25–34 comparison study on experimental design and
Li B, Friedman F (2005) Novel multiple resolutions response surface methodologies. SPE 93347, SPE
design of experiment/response surface methodology Reservoir Simulation symposium, The Woodlands,
for uncertainty analysis of reservoir simulation 31 Jan–2 Feb 2005
forecasts. SPE paper 92853, SPE Reservoir

You might also like