Transcript

Geostatistical Time-lapse Seismic Inversion

Valter Miguel Dos Santos Proença

Thesis to obtain the Master of Science Degree in

Petroleum Engineering

Supervisor: Professor Leonardo Azevedo Guerra Raposo Pereira.

Examination Committee

Chairperson: Professor Amílcar de Oliveira Soares

Supervisor: Professor Leonardo Azevedo Guerra Raposo Pereira

Member of the Committee: Professor Maria João Correia Colunas Pereira

November 2016

i

Abstract

With the decrease of the oil price and of new discoveries, it is not only important to use new exploration

and production methods and technologies but improve and optimize the monitoring of hydrocarbon

reservoirs during production. This leads to a more efficient decision making process, while reducing risks

and increasing oil recovery factors. Although a relatively new area, nowadays there are already some

methodologies that invert time-lapse seismic data, based on a Deterministic or Bayesian inversion

framework. This thesis introduces a new iterative geostatistical time-lapse seismic inversion methodology.

The objective of this work is the development and implementation of a new methodology able to reproduce

the changes of the elastic properties during production time integrating time-lapse seismic reflection data

from the different acquisition times. This new approach was tested and implemented in a realistic synthetic

dataset Stanford VI-E. The results show a good reproduction of the changes in the elastic properties of

interest. It was also concluded, during the comparison of the both methods used in this work, that the

horizontal time-slices of elastic differences computed from the conventional global elastic inversion do not

reproduce the spatial continuity of the elastic differences values. Otherwise the proposed workflow,

reproduces the spatial continuity pattern of these changes with improved accuracy.

Keywords: Reservoir characterization, geostatistical time-lapse inversion, elastic changes, 4D seismic

data, geostatistical modeling

ii

.

iii

Resumo

Com a diminuição do preço do petróleo e de novas descobertas, é importante não só usar novos

métodos e tecnologias de exploração e produção, mas melhorar e otimizar o monitoramento de

reservatórios de hidrocarbonetos durante a produção. Isto era conduzir a um processo de tomada de

decisão mais eficiente, reduzindo os riscos e aumentando os fatores de recuperação dos hidrocarbonetos.

Embora estejamos presentes numa área relativamente nova, hoje em dia já existem algumas metodologias

que invertem dados sísmicos 4D, com base em metodologias de inversão determinísticas e Bayesianas.

Esta tese apresenta uma nova metodologia de inversão de sísmica interativa 4D geostatistica. O objetivo

deste trabalho é o desenvolvimento e implementação de um novo procedimento capaz de reproduzir as

alterações das propriedades elásticas durante o tempo de produção integrando dados de reflexão sísmica

4D a partir dos diferentes tempos de aquisição. Esta nova abordagem foi testada e implementada num

conjunto de dados sintéticos realistas Stanford VI-E. Os resultados mostram uma boa reprodução das

alterações das propriedades elásticas de interesse. Conclui-se ainda, durante a comparação de ambos os

métodos utilizados no âmbito deste trabalho, que as fatias de tempo horizontais das diferenças elásticas

calculadas a partir da inversão elástica convencional global, não reproduzem a continuidade espacial dos

valores das diferenças elásticas reais. No caso contrario, o proposto método desta tee, reproduz o padrão

de continuidade espacial dessas mudanças com maior precisão.

Palavras-chave: caracterização do reservatório, inversão 4D geostatistica, mudanças elásticas, sísmica

4D,modelação geostatistica.

iv

v

Table of Contents

Abstract ............................................................................................................................................... I

Resumo ............................................................................................................................................. III

Table of Contents ................................................................................................................................ V

List of Figures .................................................................................................................................... VI

List of Tables ...................................................................................................................................... X

Acronyms .......................................................................................................................................... XII

Chapter 1: Introduction ........................................................................................................................ 1

1.1 Motivation ............................................................................................................................. 1

1.2 Objectives ............................................................................................................................. 2

1.3 Structure of the thesis ........................................................................................................... 3

Chapter 2: Geostatistical Models for Earth Science Applications.......................................................... 5

2.1 Spatial continuity analysis .......................................................................................................... 5

2.2 Variograms ................................................................................................................................ 6

2.3 Direct Sequential Simulation and co-Simulation (DSS and Co-DSS) ........................................ 10

2.4. Direct Sequential co-simulation with joint probability distributions ............................................ 13

Chapter 3: Integration of Seismic data In Reservoir Modeling Characterization .................................. 15

3.1 Seismic inversion .................................................................................................................... 15

3.2 Geostatistical Seismic Inversion .............................................................................................. 18

3.3 Time-Lapse Seismic ................................................................................................................ 19

3.4 Time-lapse seismic inversion ................................................................................................... 20

Chapter 4: Methodologies ................................................................................................................. 23

4.1 Geostatistical elastic inversion: Global Elastic Inversion ........................................................... 23

4.2 Geostatistical Time-lapse Inversion ......................................................................................... 24

Chapter 5: Case Study ...................................................................................................................... 29

5.1 Data Description ...................................................................................................................... 29

5.1.1 Structure and stratigraphy ................................................................................................. 32

5.1.2 Exploratory data analysis .................................................................................................. 34

vi

5.1.3 Parameterization .............................................................................................................. 38

5.2 Geostatistical Time-lapse Inversion ......................................................................................... 40

5.2.1 Synthetic seismic .............................................................................................................. 40

5.2.2 Acoustic and Elastic Impedance Models ........................................................................... 42

5.2.3 Histograms, Bi-plots and Statistics .................................................................................... 45

5.3 Comparison between geostatistical elastic inversion and geostatistical time-lapse inversion

methodologies ................................................................................................................................... 50

5.3.1 Differences between seismic ............................................................................................ 50

5.3.2 Acoustic and Elastic Differences ....................................................................................... 52

Chapter 6: Discussion, Conclusions and Future work ........................................................................ 55

References ....................................................................................................................................... 59

Appendix A ......................................................................................................................................... 1

vii

List of Figures

Figure 1-Seismic near-stack amplitude differences between the different times of acquisition during

production. Differences from (from the left to the right): base survey and (10 years of production, 10 and 25

of production and 25 and 30 years of production). .................................................................................... 1

Figure 2 - Graphic representation of a spherical model (Pereira, 2014). .............................................. 8

Figure 3 - Graphic representation of an exponential model (Pereira, 2014). ........................................ 9

Figure 4 - Graphic representation of a Gaussian model (Pereira, 2014). ............................................. 9

Figure 5 - Sampling of global probability distribution Fz (z), by intervals defined by the local mean and

variance of z(xu): The value � ��*corresponds to the local estimate z(xu)*. The simulated value z(xu)* is

drawn from the interval of Fz (z) defined by � (� ��*, �2(��)) ( adapted from Soares 2001). ............... 11

Figure 6 - Schematic of possible work flows in seismic inversion for reservoir properties leading to

multiple realizations of reservoir properties conditioned to seismic data, the rock - physics models, and the

model of spatial continuity imposed by geology.(a,b) Two variations of sequential work flow. (c)

Simultaneous Bayesian inversion work flow. Data come in at various stages, including seismic data for

inversion, well-log data, cores and thin sections when available, and geologic data from outcrops (Bosh et

al., 2010). .............................................................................................................................................. 16

Figure 7 - Schematic representation of the GEI workflow (Azevedo, 2013). ....................................... 24

Figure 8 - Graphic representation of the geostatistical time-lapse inversion. ...................................... 27

Figure 9 -Real models of acoustic impedance from each year of acquisition, T0, T10, T25 and T30

respectively. ........................................................................................................................................... 29

Figure 10- Real models of elastic impedance from each year of acquisition, T0, T10, T25 and T30

respectively. ........................................................................................................................................... 30

Figure 11 - Representation of the partial angle stacks form seismic reflection data, 0º, 10º, 20º, 30º and

40º respecively. ..................................................................................................................................... 31

Figure 12 – Representation of the stack wavelets correspondent to the 5 angle stacks. .................... 32

Figure 13 – Perspective view of the Stanford VI-E top structure: view from SW (left), view from SE (right).

The color indicates the depth to the top (Lee & Mukerji 2012). ................................................................ 33

Figure 14 - Histogram representing the number of acoustic impedance samples per class of the well-log

data o each year of acquisition. .............................................................................................................. 34

Figure 15 - Histogram representing the number of elastic impedance samples per class of the well-log

data o each year of acquisition. .............................................................................................................. 35

viii

Figure 16 - Well-log data bi- histogram of each acquisition time, expressing the relation between the two

properties of interest (AI and EI). ............................................................................................................ 36

Figure 17 - Chosen main direction of AI for modeling, with exponential model................................... 37

Figure 18 - Chosen main direction of EI for modeling, with exponential model................................... 37

Figure 19 - Illustration of the simulation mask, T1T2, T2T3 respectively. The red color represents the

places to be simulated. .......................................................................................................................... 39

Figure 20 - Inverted residual models from T10, T25, and T30 respectively. ....................................... 39

Figure 21 - Global correlation coefficient evolution between synthetic and real seismic reflection data for

the four available seismic vintages using geostatistic time-lapse inversion method. ................................ 40

Figure 22 - Comparison of all angle stacks between the real seismic (left), the synthetic seismic from

the best-fit models (Center) and the Synthetic seismic computed from the mean of the models from the last

iteration. The seismic results are related to the different times of acquisition divided by rows, being the first

related to the year 0 and the last to the year 30. ..................................................................................... 41

Figure 23 - Comparison of the near angle stack between the real acoustic impedance models (Right),

the simulated best-fit acoustic impedance models (Center) and the mean of the simulated acoustic models

computed from the last iteration. The first line is related to the year 0 and the other too years 10, 25, and 30

respectively. ........................................................................................................................................... 43

Figure 24 - Comparison of the near angle stack between the real elastic impedance models (Right), the

simulated best-fit elastic impedance models (Center) and the mean of the simulated elastic models

computed from the last iteration. The first line is related to the year 0 and the other too years 10, 25, and 30

respectively. ........................................................................................................................................... 44

Figure 25 - Acoustic impedance histograms from the well-log data of each year against the histograms

from the best-fit models of each acquisition year. ................................................................................... 46

Figure 26 - Elastic impedance histograms from the well-log data of each year against the histograms

from the best-fit models of each acquisition year. ................................................................................... 47

Figure 27 - Bi-histograms of well-log data against best-fit models of each year, expressing the relation

between the two properties of interest, AI and EI. ................................................................................... 48

Figure 28 - Comparison the near angle stacks between the real seismic (left column), the Synthetic

seismic computed from the mean of the models from the last iteration using Geostatistic Time-lapse

inversion (center column) and the synthetic seismic from the best-fit models using GEI (right column). The

seismic results are related to the different times of acquisition divided by lines, being the first related to the

year 0 and the last to the year 30. .......................................................................................................... 51

ix

Figure 29 - Acoustic impedance differences from: the real (left), geostatistical time lapse inversion

(middle) and global elastic inversion (right) for the T0-T10, T10-T25, and T25-T30 years respectively .... 52

Figure 30 - Elastic impedance differences from: the real (left), geostatistical time lapse inversion (middle)

and global elastic inversion (right) for the T0-T10, T10-T25, and T25-T30 years respectively .................. 53

x

List of Tables

Table 1 - Values of the modeled variograms. .................................................................................... 37

Table 2 - Values of the modeled variograms ..................................................................................... 38

Table 3 – Acoustic impedance main statistics of well-log data and best-fit models related to each year.

.............................................................................................................................................................. 49

Table 4 – Elastic impedance main statistics of well-log data and best-fit models related to each year 49

xi

xii

Acronyms

AI: Acoustic Impedance

Cdf: Cumulative distribution function

Co-DSS: Direct Sequential co- Simulation

DSS: Direct Sequential Simulation

EI: Acoustic Impedance

GEI: Global Elastic Inversion

Ip: Primary impedance

Is: Secondary impedance

Pdf: Probability distribution function

Vs: Secondary wave velocity

Vp: Primary wave velocity

1

Chapter 1: Introduction

1.1 Motivation

Time-lapse or 4D seismic reflection data involves acquisition, processing, and interpretation of repeated

seismic surveys in time over the same spatial region. From a hydrocarbon perspective, the objective of 4D

seismic reflection data is to determine the changes occurring in the reservoir as a result of hydrocarbon

production or injection of water or gas into the reservoir (Figure 1). A typical time-lapse different dataset

allow assessing the amplitude differences between the different seismic vintages and infer about the

changes in the elastic properties of the reservoir. Without considering noise present in the seismic reflection

data the differences should be close to zero, except where reservoir conditions changed. Using time-lapse

methodologies help better understanding about the fluid production, injection and the geometry of the active

zones of the reservoir. The production in these reservoirs causes changes in the subsurface elastic

properties and therefore changes in the seismic wave velocities. This changes can be monitored and

interpreted using time-lapse seismic reflection data. The interpretation of the reservoir pressure, saturation

and fluid contacts using these kinds of datasets can contribute for an improvement in production, optimizing

the placement of producers and injector wells.

Figure 1-Seismic near-stack amplitude differences between the different times of acquisition during production.

Differences from (from the left to the right): base survey and (10 years of production, 10 and 25 of production and 25

and 30 years of production).

2

Time-lapse studies normally use two or more migrated 3D seismic volumes obtained months or years

apart for the same spatial area. These can consist of stacked data volumes or stacks created from partial

offsets if AVO/AVA aspects are considered. To study and interpret the changes in the seismic volumes,

some different approaches have been done, however none within a geostatistical framework. Geostatistical

modeling and inversion techniques contribute for a better reservoir characterization since they allow the

assessment of the spatial uncertainty of the inferred property. There are several modeling techniques that

allow accessing different levels of detail in the characterization of the reservoir model, however and

regardless of the technique used, there will be always a degree of uncertainty associated with inverse

models that is important to evaluate (Azevedo, 2013). Inverting different vintages of seismic reflection data

separately is not enough since there is no conditioning of the properties changes between each acquisition

time. The main motivation of this work is to have a workable algorithm that with a unique geostatistical

approach, integrate the time-lapse seismic reflection data inverting all times simultaneously, ensuring the

reproduction of the elastic changes in reservoir during production.

1.2 Objectives

The main objective of this Master thesis is the development and implementation of a new iterative

geostatistical time-lapse seismic inversion methodology. The proposed algorithm was tested in the classic

Stanford VI-E reservoir (Lee & Mukerji, 2012). The main objectives of this work may be summarized by the

following:

- Implementation of a new Geostatistical time-lapse seismic inversion methodology.

- Compare the new algorithm results with those obtained by standard geostatistical elastic

inversion (GEI) of the different acquisition times to understand the importance of the joint

inversion.

- Compare the results of the new algorithm with the real synthetic data models of Acoustic

Impedance (AI) and Elastic Impedance (EI) to see the effectiveness of this new method.

- Understand the pros and cons of this new method compared to the ones already in use.

In the developing of this thesis, various computer tools were used the geostatistical modeling software

GeoMS ( CERENA / IST) for modeling the variogram data logs of the wells, the two geostatistical seismic

inversion methodologies were performed using code developed in Matlab® (Math works) and Petrel®

(Schlumberger) for data visualization and manipulation.

3

1.3 Structure of the thesis

This thesis is divided in six chapters:

• The first chapter, introduce the scientific question that led to the research summarized by this

thesis, as well as the objectives to accomplish.

• Chapter II introduces the theoretical background in which the geostatistical model is developed,

and the description of the direct sequential simulation and direct sequential co-simulation

algorithms that are part of the inversion procedures developed under the scope of this work.

• Chapter III gives an overview on the basics of the seismic inversion and the global geostatistic

elastic seismic inversion algorithm that is the base of the proposed methodology of this work. It

is also presented some ideas and studies developed using time-lapse seismic as well as some

historical background.

• Chapter IV describes and illustrates the methodologies of the two workflows tested,

geostatistical elastic inversion (GEI) and geostatistical time-lapse inversion (GTI).

• Chapter V presents the case study and the application of the two methodologies in a realistic

synthetic reservoir, Stanford VI-E, as well the results from those tests.

• Chapter VI show the conclusions taken from this work concerning the limitation and strong points

observed, as well as some discussion about further applications in the practical implementation

of the introduced methodology.

4

5

Chapter 2: Geostatistical Models for Earth Science Applications

2.1 Spatial continuity analysis

The study of the spatial continuity pattern of the property of interest aims to understand the dispersion

of the experimental data and the degree of anisotropy between the variables in question, in order to know

how a natural phenomenon behaves spatially.

There are some rules and combined practices that should be considered during the study of the spatial

continuity pattern of a given property, such as the calculation of experimental variograms and its modeling.

These rules can be compiled and subdivided into the following groups: Quality and Sampling type,

Conceptualization of Geostatistical Model and Data Analysis Statistics (Soares, 2006).

The quality and sampling type are very important since, sometimes the sampling data has anomalous

values, from asymmetric distributions, what can generate anomalous values in the variogram. Therefore it

is necessary to evaluate those sampling clusters to eliminate outliers that will change significantly the

variogram estimators, the means and the histogram, for example (Soares, 2006). After considering the

geostatistical model as basis, it should be considered the statistical analysis of the available data, which

includes the evaluation of the main characteristics of the regionalized variable by univariate and multivariate

data analysis. This statistical analysis can be done in three descriptive groups of spatial distribution:

univariate description, bivariate description (when possible) and spatial description (Soares, 2006)

The Univariate description consists of making individual analysis of the variables of interest to get an

idea of the dispersion of the attribute under consideration, understanding the main characteristics of the

data. The study is based on frequency tables, histograms and cumulative histograms, obtaining like this the

distribution location measures (minimum, maximum, average, median, quartiles, and percentiles),

measures of dispersion (range, variance, standard deviation) (Soares, 2006).

The bivariate description consists in the simultaneous study of the behavior of two variables. It allows to

establish relations between two variables, allowing to determine whether the differences between the

distributions are significant at the statistical level. Bivariate analysis uses essentially biplots, bi-histograms

and different regression analysis tools (Soares, 2006). In the bivariate analysis of this scientific work was

used a type of graphical representations called biplots, one most widely used type of chart on the correlation

between two variables defined in the same spatial positions.

6

In the study of the spatial distribution of a given variable it is attempted to visualize de variable behavior

in relation to their dispersion in space. It is also intended to make the planning of the basic parameters for

the calculation of variogram, by viewing the spatial arrangement of the experimental variable data and get

a first idea of the spatial dispersion characteristics of the same variable, like anisotropy and

anomalies,(Soares, 2006). There are some situations that need to be taken under account, essentially when

they are characterized by higher values of local variances that imply higher errors in the spatial inference

and especially with greater uncertainty, like for example:

• The higher the extreme values and agglomerates in the same area, the greater is the average

value and the smaller the value of the local variance;

• The higher the extreme values and there is no concentration in the same are, the greater the

average value and the local variance;

• The lower extreme values in this area, the lower the average value unlike the local variance that

tends to increase.

2.2 Variograms

In geostatistics, the variogram (or semi-variogram) describes the spatial relationship between points of

a given attribute Z (x). It is a spatial continuity instrument used to quantitatively represent the change in a

regionalized phenomenon in space, i.e. the variable of interest (Soares, 2006). It is an intrinsic function that

reflect the spatial structure of the phenomenon under study, measuring statistical relationships like the

covariance that exist between samples of successive values of distance (h) between each pair of

experimental data. The variogram is also an increasing function in which the values of the abscissa, vary

until a given value of h, a value known as the variogram range (Soares, 2006).Variograms that have a sill

are associated with stationary regionalized variables, otherwise they are considered non-stationary. Note

that the phenomenon under study can nevertheless be stationary in sub-regions of the area in study.

The estimator of the variogram (Equation 1) provides information on the spatial continuity of the

attributes, at different distances�Z(x + h)�, for various values of h, for the calculation of this estimator. The

centered covariance estimator (Equation 2) provides information of spatial average of the product between

different attributes at different distances�Z(x)Z(x + h))�. These estimators are the geostatistical measures

more often used which represent the correlation between samples(Zα and Zα + h), and can be calculated

by the following expressions (Soares, 2006):

7

γ(h) = 12 E[Z(x) − Z(x + h)]� (1)

C(h) = E�Z(x)Z(x + h) − E�Z(x) E�Z(x + h) (2)

The standard form of covariance, the correlogram (Equation 3), can be defined as the ratio between the

covariance and variogram:

γ(h) = C(0) − C(h) (3)

Variograms are generally anisotropic, but can also be isotropic, representing a different behavior of

spatial correlation in different directions or in a different distance h. This difference in behavior is

representative (in most cases) when we relate variograms of horizontal directions with variograms with

vertical directions. The anisotropy is dependent on the variable or natural system under study. For example,

a given geological layer, should have a higher spatial continuity in the horizontal direction when compared

with the spatial continuity of the vertical direction (Soares, 2006). The opposite phenomenon occurs when

computing experimental variogram from existing well data. In these cases, the vertical direction is more

spatially sampled then in the horizontal direction. As the module of difference between samples (|ℎ|)

increases, the average variation between pairs of samples tends to increase.

Once the variogram is an increasing function, as the module |ℎ| increases, the average variation between

pairs of samples tends to increase as well, until it reaches the range, distance from which it no longer occurs

that increase and the function begins to stabilize. Using the variogram in the main directions, it is possible

to calculate the average dimensions of the body along respective directions (Soares, 2006).

When created the experimental variogram from the set of samples in region of interest it is necessary to

adjust the data to a theoretical model, doing the data modeling through a general mathematical function

that represents the continuity of the spatial variable (Soares, 2006). This stage is very important as it

summarizes the structural characteristics of spatial phenomenon under study, including anisotropy and

degree of continuity of the variable under study.

The practice of geostatistical modeling is limited to the use of a restricted set of functions that covers the

spatial continuity pattern of most natural phenomenon. It is necessary to select the best suited to each case

but essentially those that meet certain conditions: positive definite models that can provide stable solutions.

The positive condition of the models restricts from the beginning, the choice of possible functions for

interpolating the experimental variogram samples. The theoretical models most used variograms are:

Spherical model, Exponential and Gaussian have, all of them, functions of two parameters, namely a Sill($),

which is the superior limit that the variogram values tend as the distance between sampling pares increase,

and the amplitude ( h = a ) of the variogram, that the distance from which the values of the variogram ( %(ℎ))

8

stop to grow and are equal to the sill, that is normally coincident to the variance of the attribute in study. The

amplitude measures the distance from which the values of that attribute begin to show no correlation

between them.

The spherical model (Equation 4) is one of the models most used in modeling experimental variograms.

Is characterized by a rapid growth at the origin and is expressed by the mathematical equation:

%(ℎ) = & C '1.5 ha − 0.5 *ha+,- for h ≤ aC for h > a (4)

Figure 2 - Graphic representation of a spherical model (Pereira, 2014).

Regarding the exponential model (Equation 5) and comparing with the previous model, the exponential

model has a growth near the origin and a greater spatial continuity for larger distances of h. This model

asymptotically reaches the sill and the amplitude is defined as the distance over which the model value

reaches the 95% sill (Soares, 2006), namely (3) = 4, 678 . The exponential model is expressed by the

following equation:

γ(h) = C 91 − e;,<= > (5)

9

Figure 3 - Graphic representation of an exponential model (Pereira, 2014).

The Gaussian model (Equation 6) is characterized by its parabolic behavior near the origin, with a slow

and growth favorable to the adjustment of regular phenomena, in contrast to the rapid growth of the previous

two models who report a more irregular behavior of the phenomena. The Gaussian model reflects a great

continuity of the variable under study, with the general expression:

γ(h) = C ?1 − @*;,ABCB +D (6)

Figure 4 - Graphic representation of a Gaussian model (Pereira, 2014).

The amplitude value corresponds to the distance for which the model reaches the 95% of the sill,γ(a) =0,95C; like the exponential model.

10

2.3 Direct Sequential Simulation and co-Simulation (DSS and Co-DSS)

In the geostatistical inversion methodology proposed under the scope of this work it is used the direct

sequential simulation (DSS).This stochastic sequential simulation algorithm uses the variable in its original

domain not requiring any transformation. This approach (Figure 5) is based on the principle introduced by

Journel (1994) for the Gaussian Sequential Simulation (GSS). However, it was with the Soares (2001)

optimization that this algorithm allowed to successfully reproduce the variogram and the histogram of a

continuous variable (Soares, 2001) using the average and local estimated variances by simple kriging for

re-sampling of the distribution law and not defining the local distribution laws (Similar to the GSS

method).Nevertheless, the essential advantage of this algorithm is allowing simulation and co-simulation,

without the need for any transformation of the original variables (Soares, 2001).

In practice, there is a re-sampling of the global cumulative distribution function (CDF) globalFz(z), to

obtain a new auxiliary probability distribution function,F′z(z) defined by intervals centered on the estimation

by simple kriging,I(JK)*, and with a proportional amplitude to the variance of kriging, σ� sk (Soares, 2001).

This new function, F′z(z) can be constructed by the definition of a Local Gaussian CDF, � (� (��)*, ��MN(�O))

, generated by the Guassian transform (Equation 7) in Fz interval (P), centered on the simple kriging estivate

value with an interval range proportional to the kriging variance, ��MN (Soares,2001).

1n Q z(xR)S

RTU = [z(xV)]∗ (7)

1n QXz(xR) − [z(xV)∗]Y� = σMN�(xV)S

RTU (8)

y(x) = φ�z(x)�, with G�y(x)� = F`�z(x)� (9)

11

Figure 5 - Sampling of global probability distribution Fz (z), by intervals defined by the local mean and variance of

z(xu): The value � (��)*corresponds to the local estimate z(xu)*. The simulated value z(xu)* is drawn from the

interval of Fz (z) defined by � (� (��)*, ��ab(�O)) ( adapted from Soares 2001).

The DSS simulation algorithm can be described and summarized by the following sequence of steps

(Soares, 2001; Azevedo, 2013):

1. Definition of a random path that covers the entire simulation grid and visit all the nodes, ��, with �=1,...,Ns where N is the total number of nodes that make the simulation grid;

2. Estimation of the local average parameters, P(��)* and the variance, ��MN(�O), estimated values by

simple kriging and conditioned by the original experimental data simulated previously;

3. Define the interval of Fz (z) to be sampled, as previously explained (Figure 5);

4. Generation of a Pa(��) value from the cdf of Fz (z);

a. Generate a value u from the uniform distribution between [0,1]

b. Generate a value �s from � (� (��)*, ��MN(�O))

c. Return the simulated value Pa(��) = c;U(�a) 5. Loop until all the N nodes of the simulated grid have been simulated.

The co-simulation of two variables is done when those express a dependency between them, and in

addition to the individual distributions and variograms, their simulation must express that correlation.

Basically two different approaches exist for the co-simulation, the one utilized in this thesis uses the

12

application of the Bayes rule, where the simulation of each variable it is processed in sequence of simulation

of conditional distributions (Gomez-Hernandez & Journel, 1993, Goovaerts, 1997a).

The Bayes principle is applied in the co-simulation procedure to calculate the distribution of the local

conditional probability at a given location �. Assuming that two dependent variables: ZU(x) , as the first to

be simulated individually by DSS due to its greater spatial continuity, and the second variable, Z�(x) , that

is simulated conditioned by the previous simulated values of ZU(x). In the location xd the values of Z�(xd)

are generated from the law of conditional distribution:

ef(g�(�d)|g�(�C) = P�(�C), gU(�h) = PU a(�h), i = 1, j) (10)

Where zUM(xR) are the values previously simulated of ZU(x) and z�(x=) are the experimental and the

values already simulated of Z�(x) within a likelihood adjacent to the location xd. After the simulation of gU(�), the same DSS algorithm is now applied to Z�(x), using the images previously simulated of ZU(x) as

secondary and with a spatial correlation between the both variables. The values of Z�(x), are generated in

any spatial location of �d, from the laws of conditional distribution to the values previously simulated.

In direct sequential co-simulation (co-dss), it is necessary to estimate the local co-kriging [z�(xV)] CSK

(equation 11) and the co-krigring variance (equation 12) of the property to be simulated. This can be done

by using simple co-kriging estimate (Soares, 2001).

[z�(xd)∗]klm = Q λop

=TU[z�(xo) − m�] + λrXPaU(xd) − mUY + s� (11)

σklm�(xd) = Var�Z�(xV)∗ − Z�(xV) (12)

Where λ= and λr are the weights and mU and m� are the means for each variable respectively.

The Co-DSS can be described by the following sequence of steps (Soares, 2001; Azevedo 2013):

1. Definition of a random path in the simulation grid �d to be co-simulated;

2. Estimation of the parameters of local mean P(�O) ∗ and variance σklm�(xd) ,values estimated by co-

kriging variance( σklm�(xd) ) conditioned to the original and simulated previously values;

3. Definition of the interval ef(g) to be sampled, as previously explained (Figure 5);

4. Calculation of a valor gu2(x0)* , from the cdf of F`(P) ;

a. Generate of a valor v from a uniform distribution w(0,1);

13

b. Generate a valor �� from � (�(�d)*, σ�xMN(�d)) ;

c. Return the simulated value g� a(�d) = c;U(�a) 5. Loop until all the N nodes of the simulated grid have been simulated.

The co-simulated models allow to reproduction of the values of the experimental data ((P� (�C)) in their

locations, the marginal probability distribution of ef(P) and the pattern spatial continuity imposed by a

variogram model. However the simulated models that result from the co-DSS only allow the representation

of linear correlations between the primary and secondary variables, independently from the joint probability

distribution observed in the experimental data (Horta & Soares, 2010).

2.4. Direct Sequential co-simulation with joint probability distributions

Because of the limitations regarding the assumption of linear correlation between primary and secondary

variables, the traditional Co-DSS methodology, was modified to include non-linear correlations. This new

approximation, introduced by Horta & Soares (2010) and named direct co-simulation with joint probability

distributions ensures the reproduction of the experimental joint probability distribution between the primary

and secondary variables on the simulated models (Horta & Soares 2010).

As any stochastic sequential simulation, the procedure of the co-DSS with joint distributions is based in

the application of the Bayes principal, in a successive sequence of steps. This principle is in the base of the

estimation of the local conditional probability distributions, sequentially in every node of the simulation grid

(Azevedo, 2013). The simulation methodologies assume, in a model previously simulated, the secondary

variable,ZU(x) is initially calculated for all the simulation grid: followed by the primary variableZ�(x). The local

cumulative distribution function, from where the value of z�y (x) is calculated, is defined by the simple co-

located co-kriging estimate and by the co-kriging variance. When compared with the traditional co-DSS, the

main difference in this approximation is the way that the conditional cdf (F (Z�(x)|ZU(�)), is sampled. The

value of z�y (�), instead of being calculated from the global cumulative distribution function, like in the

traditional co-DSS, is calculated from the local conditional distribution, expressed by the next equation

(Equation 13) (Horta & Soares 2010):

F[Z�(u)|ZU(u) = ZU(u)∗] = prob�Z�(u) < z|ZU(u)∗ K (13)

The co-DSS algorithm with joint probability distributions, allows the reproduction of the values of the

experimental data in their locations in the simulated models, the original spatial continuity model and

reproduces the marginal distributions from the primary variable. This co-DSS algorithm distinguishes by the

14

ability to reproduce the joint distributions, estimated from the experimental data between the primary and

secondary variable,e (g�(�)|gU(�)) even when the original joint distribution is complex (Azevedo, 2013).

The procedure of the co-DSS algorithm can be synthetized by the following steps (Horta & Soares, 2010;

Azevedo, 2013)

1) Estimation of the global bi-distribution, e (g�(�)|gU(�)), from the experimental data;

2) Simulation of a secondary variable ZU(x), with the DSS algorithm for all the simulation grid;

3) Estimate the local mean and variance in, �O, with the estimation of the simple co-krigring co-

localized and correspondence co-krigring variance, conditioned by the original experimental data Z�(x=), by the data previously simulated Z�(x=)* end by the secondary variable value ZU(xV)*, inside

an adjacent likelihood u;

4) Calculation the cdf,F (Z�(x)|ZU(x)), with the simulated values of the secondary variable , PU�(u).

5) Simulation of the value P��(u) , from the conditional cdf, ,F (z�(x)|zU(x) = zUy(u)) , using the

approximation of the traditional DSS algorithm;

6) The models resulting of the stochastic simulation, from the co-DSS with joint probability distributions

application, ensure the spatial reproduction of the bi-distributions between the primary and the

secondary variables (Azevedo, 2013). With a representation of the vertical sections related with a

velocity model of P waves we can see the differences between the co-simulation of the models that

are the result of the DSS and the models that result from the use of DSS with joint probability

distributions and the last one, presents a bigger variability with a detailed approximation, compared

with the original co-DSS.

15

Chapter 3: Integration of Seismic data In Reservoir Modeling Characterization

3.1 Seismic inversion

Over the last few years, seismic inversion techniques had a major impact in the characterization of the

subsurface geology, helping the professionals to make better and more efficient decisions on the oil industry.

Seismic reflection data is used in reservoir characterization not only for obtaining a geometric description of

the main subsurface structures but also for estimating the spatial distribution of the subsurface properties

such as lithology and fluids. However, transforming seismic data into reservoir properties is an inverse

problem with a non-unique solution even for noise-free data, due to the limited frequency of recorded

seismic waves, forward-modeling simplifications needed to obtain solutions in a reasonable time,

uncertainties in well-to-seismic ties (depth-to-time conversion), in estimating a representative wavelet and

in the links between reservoir and elastic properties. (Bosch et al., 2010). The main objective of a study

associated with the use of seismic inversion techniques is the inference of the subsurface elastic properties

from available seismic reflection data. The process of predicting the behavior of a particular physical system

under investigation is normally designed as forward modelling. The relationship between the observed data

and the earth models can be express by the next equation (Tarantola, 2005; Azevedo et al., 2012):

���� = �(�) + � (14)

Considering the case of seismic inversion problems, ���a represents the recorded seismic reflection data,

F is the convolution model and m is the model parameter space for the properties to invert that depend on

the objective of the inversion, which can be acoustic impedance and/or elastic impedance, density or the P

and S wave velocities. The forward model (F) related to the previous equation can be defined by the next

equation:

� = � ∗ � (15)

where A corresponds to the recorded seismic amplitudes, r to the reflection coefficient of the subsurface

that are convolve with a wavelet w. Theses variables are dependent on the elastic properties ( density and

P and S wave velocities) of the subsurface geology ( Azevedo,2013).

Therefore the seismic method has some limitations that make the seismic inversion problems non-linear

and with multiple solutions, the limited bandwidth and resolution of the seismic reflection data, noise,

measurement errors, numerical approximations and physical assumptions about the involved forward

16

models (Tarantola 2005;Bosch et al. 2010; Tompkins et al. 2011). Because of these limitations, the inverted

elastic models, that are the result of the seismic inversion process, are only a possibility between n models

of the earth that are unknown, which equally satisfy the same observed seismic data (Azevedo, 2013), this

means that exist multiple acceptable solutions for the convolution of the reflection series with the wavelet

combined with the observed seismic (Francis, 2005). Therefore if the correlation between the recorded

seismic and the synthetic seismic, calculated from the best inverted models, is weak, then we can conclude

that the correlation between the correspondent real elastic and synthetic models will be weak too (Azevedo,

2013).

Inverted elastic Earth models are nowadays routinely used in reservoir characterization studies. Because

of this reason it is extremely important to understand the different seismic inverse methodologies and the

assumptions associated with each, assumptions about prior probability distributions and about spatial

continuity patterns have a significant impact in the exploration of the model parameter space and

consequently in uncertainty assessment of the best-fit inverse models. Seismic inversion problems can

therefore be divided in two main approaches, the deterministic, also designated as optimization techniques,

and stochastic (Bosch et al. 2010), this last one is also divided into two different methods, Bayesian and

iterative geostatistical inversion methodologies (Figure 6).

Figure 6 - Schematic of possible work flows in seismic inversion for reservoir properties leading to multiple

realizations of reservoir properties conditioned to seismic data, the rock - physics models, and the model of spatial

continuity imposed by geology.(a,b) Two variations of sequential work flow. (c) Simultaneous Bayesian inversion work

flow. Data come in at various stages, including seismic data for inversion, well-log data, cores and thin sections when

available, and geologic data from outcrops (Bosh et al., 2010).

17

Deterministic methods are the most used for post-stack and pre-stack seismic inversion are band-limited,

or integration of the seismic trace (R.O., 1979), Colored Inversion (Lancaster & Whitcombe, 2000) and

sparse-spike and model-based (Russell & H., Introduction to seismic inversion methods , 1988). From this

group of procedures, the sparse-spike and model-based methodologies are the most used deterministic

inversion techniques among the geophysical community (Bosch et al, 2010). Sparse-spike techniques

deconvolve the seismic trace under sparseness assumptions of the reflectivity series, were the reflectivity’s

that are obtained have their impedances computed, including the missing low frequencies, usually from the

well data, seismic-velocity analysis, or a kriging estimate of the low-frequency trend. In the model-based

methods (Russel & Toksoz, 1991), were the inversion algorithm perturbs the initial given model until some

minimization criteria are satisfied and the objective function, or the function to be minimized, is usually a

difference between the observed and modeled data (Bosch et al, 2010).

Despite the deterministic inversion methods are cheaper and faster in terms of computational work, they

give us just a smooth representation of the Earth’s subsurface with much less spatial variability when

compared with the real subsurface geology. If we are aiming to understand more global and less detailed

information they can be very suitable.

Stochastic methods can be, compared with the deterministic ones, more complex slower to process and

more expensive, but they assess the uncertainty, giving us a number of equiprobable solutions that

represent the different possible scenarios that can exist. Among the stochastic methods, Bayesian

approaches ensure the propagations of the uncertainty from the prior probability distributions, estimated

from experimental data (e.g. well-log data) to the probability distributions of the model parameters space

(Grana et al. 2012). Any inversion problem can be posed as a Bayesian inference problem, were the prior

knowledge is updated, accounting for observations (e.g. Tarantola, 1987, 2005; Duijndam, 1988a, 1988b;

Ulrych et al., 2001).

The assumptions about the prior distributions and the spatial continuity pattern have big impact in the

assessment of the uncertainty. Among these assumptions we have the linearization of the forward-model

operator, Gaussian distribution for the prior probability distributions and for the errors present within the

observed seismic data. (e.g. Buland and More 2003; Buland and El Ouair 2006).

The computational time is much lower when compared with the approaches based on genetic algorithms

or simulated annealing and the inverse solution is mathematical tractable. However, the reduction of the

computational burden has a direct impact in the exploration of the uncertainty space even in the present of

exact prior information (Scales and Tenorio 2001; Tarantola 2005; Bosh et al., 2010). Grana and Della

Rossa (2010) developed and inversion framework for the Bayesian linearized inversion, and in order to

overcome the Gaussian assumption about the prior probability distribution, they use Gaussian-mixture

models, were they define the prior probability distributions as Gaussian-mixtures, and a linearized forward

model, as well as the inferred posterior probability distribution.

18

Geostatistical approaches define the inverse solution as a probability density function on the model

parameters space (Bosch et al., 2010). Normally the inverse solution is achieved by sampling the model

parameter space by Monte Carlo or by using geostatistical sequential simulation combined with global

optimization algorithms. Genetic algorithms (e.g. Mallic 1995; Mallick 1999; Boschetti, Dentith, and List1996;

Soares et al., 2007) and simulated annealing (Sen and Stoffa 1991; Ma 2002) fall within this class of inverse

methodologies.

3.2 Geostatistical Seismic Inversion

The methodology proposed under the scope of this work is based on a family of geostatistical inversion

methodologies that use the principle of cross-over genetic algorithm (Azevedo et al., 2013b). This iterative

geostatistical methodology is based on two key main ideas: the use, at the end of each iteration, of global

optimizer based on a genetic algorithm with the cross-over principle; and the perturbation of the inverted

models with stochastic sequential simulation, the Direct Sequential Simulation (Soares, 2001). The affinity

criterion, that measures the convergence of the inverse procedure and guide the genetic optimization

algorithm, is based on the correlation coefficient between synthetic and real seismic volumes, which is

achieved by the maximization or minimization of an objective function that measures the mismatch between

inverted and real seismic. Seismic inversion as explain in the precious section is an inverse, nonlinear and

with multiple solutions problem that can be summarized by the Equation 14 (Tarantola, 2005), and in

geostatistical seismic inversion case, one wish to calculate the parameters (elastic impedances) which give

rise to the known solution, real seismic. The objective is to estimate a subsurface Earth model that after

being forward modeled, produces synthetic seismic data that shows good correlation with the recorded data.

Depending on the limitations associated with the chosen inversion methodology resultant inverse models,

can be acoustic and/or elastic impedance, for post-stack seismic data, or density, P-wave and S-wave

models if a more elaborated inversion algorithm is being used to invert pre-stack seismic reflection data

(Francis, 2006).

Geostatistical seismic inversion methodologies may be divided in Trace-by-trace geostatistical seismic

inversion or global geostatistical seismic inversion methodology depending how the model parameter space

is perturbed. The first class of methods perturb the inversion grid sequentially in a trace-by-trace basis,

while global approaches generate at once realizations for the entire inversion grid. Contrary to trace-by-

trace approaches, global approaches can keep low signal-to-noise areas with a poor match during the entire

inversion procedure.

19

3.3 Time-Lapse Seismic

Time-lapse seismic or 4D seismic refers to the acquisition of 2D/3D seismic reflections in different times

over the same spatial area. The differences between the different seismic acquisitions acquired under the

same conditions, provide the information about the variation of the elastic properties under a production

scenario. This differences serve to compare changes in the amplitude volumes so that the pressure of fluid

flow, water saturation and other dynamic properties can be studied. The fact that the reservoirs present are

becoming deeper and with more complex geology, creates a need to extend their life of production to the

maximum, using better and more efficient recuperation methods. The method to study 4D seismic involves

acquisition, processing and interpretation of the repeated seismic data to comprehend the changes in the

reservoir during production. This knowledge as very real budgetary implication as increasing the recovery

factor, even for a 1%, will have a significant impact on the revenue of the companies.

Originally, seismic time-lapse monitoring was strictly a qualitative subject, and changes observed visually

were related in a heuristic way to the reservoir product. As the seismic technology matured, and greater

accuracy was assigned to the differences observed, there was an increasing effort to incorporate the results

into more quantitative studies. Initially, output from reservoir simulators was used to provide input to

Gassmann fluid-substitution schemes to compare with seismic observations. Then, some pressure effect

on the rock frame were included. The comparison between predicted seismic changes and those observed

was sometimes used to update the original reservoir model, just as history simulation and its history-

matching capability with the data provided by seismic time-lapse monitoring, guiding the simulator (or the

engineer) in the inter-well regions and further constraining the initial model. These efforts are in some cases

related to work on geo-mechanical modeling of reservoirs for the inclusion of deformation in simulation.

There are three main reservoir properties that are affected by the hydrocarbon production, saturation (gas,

oil and water), pressure and temperature. These parameters directly affect the following seismic

parameters, p-wave, s-wave and density. For time lapse to be effective, changes in the reservoir properties

must cause a detectable change in the seismic parameters. With this concept seismic surveys are recorded

over the time span in a hydrocarbon producing field to measure elastic changes and transferring the

knowledge into dynamic properties such as saturation, pressure and temperature. A time lapse study relies

on a key factor to a successful 4D experiment, repeatability. Repeatability is the indicator of how identical

base and monitor surveys are acquired and processed in the same conditions using the same parameters,

without sufficient repeatability, subtle 4D changes cannot be detected. This rule can be divided in two

categories which are related with acquisition (Shooting directions, acquisition systems, geometry/positioning

and seasonal) and processing (algorithm, software and hardware). For good time-lapse surveys, we should

have two above mentioned component as identical as possible, if seismic surveys are repeatable then they

should have almost identical traces at respective locations.

20

Observations made on seismic time-lapse studies frequently include changes in amplitude and changes

in time. Changes in amplitude can often be used to directly monitor fluid migration because the reflection

character changes by replacing oil/water with gas for example. Other changes in reservoir properties must

always be considered, such as effective pressure acting on the rock frame, and it is not always possible to

separate these effects using stacked data alone. The use of offset stacks or elastic impedance volumes

helps to reduce this uncertainty, separating the changes that seem to be caused by fluid substitution from

those caused by pressure change, and a seismic petro-physical model in required in the interpretation. The

change in seismic velocity between separate monitoring experiments will also result in a change of two-way

travel time to reflectors that lie beneath the producing reservoir. This velocity-induced “pull-up” may be

monitored and provides an indication of the spatial location of reservoir changes. Because of this effect,

interpreters should take great care in the use of direct difference volumes that are obtained simple by

subtraction of seismic volumes obtained at different times, the seismic reflection data should be processed

in order to address this problems.

3.4 Time-lapse seismic inversion

Time-lapse seismic data is key information that can constrain reservoir models, to assess changes of

the subsurface properties due to changes in hydrocarbon saturation and pressure. This can be done by

comparing the base seismic data (acquired pre-production) with the monitor surveys acquired over the time

(Landro et al., 2003; Furre et al., 2006). When the repeatability of the seismic data is good, time-lapse

seismic data can be used to determine changes in the subsurface that are related to reservoir production.

Time-lapse seismic data can be inverted for dynamic properties or integrated in dynamic reservoir models

and it has recently been used as an additional data set for history matching in reservoir applications

(Aanonsen et al., 2003) and integrated workflows with production and seismic data .In this type of approach,

time-lapse synthetic seismic data are computed based on dynamic reservoir models, and the difference

between synthetic and real seismic is included in the objective function in the history matching workflow.

Traditional approach for the 4D inversion is to invert base and monitor survey independently and subtract

the elastic results to infer the change from production effect. This approach results in the 4D anomalies hard

to interpretation. Differences interpreted from independent inversions may not be related with production

but only with differences of the different seismic vintages. They may be artifacts related with convergence

to different solutions in the model space that are not necessarily consistent with the real subsurface elastic

changes. Inverting all vintages simultaneously, combined from example in a single objective function, allows

inferring reliable and consistent models for the expected differences in the elastic behavior. As the

technologies related with 4D seismic inversion mature, there is an increased emphasis on quantitative 4D

interpretation workflows, involving 4D elastic inversion followed by rock physics inversion to estimate

21

changes in fluid saturation and reservoir pressure. Obtaining reliable estimates of time-lapse changes in

elastic properties using 4D seismic inversion is a challenging task which has received considerable attention

in the last years. A couple of 4D inversion approaches have been proposed (Sarkar et al., 2003): workflows

where base and monitor surveys are inverted separately and then subtracted to calculate changes in elastic

attributes, sequential inversion schemes where inversion results for a base survey are used to define an

initial model for inverting a monitor survey (e.g., Lafet et al.,2005), direct inversion of amplitudes differences

for changes in elastic parameters (e.g, Buland and El Ouair, 2006) and global inversion methods where all

vintages are inverted simultaneously (e.g., El Ouair and Stronen, 2006). Studies and experiences shows

that coupling the inversion of base and monitor surveys is important to obtain quantitative estimates of

impedance changes and reduce the non-uniqueness of the inversion process. Among the stochastic

methods, one of the most recent methods is the Bayesian inversion of time-lapse seismic data for the

estimation of static reservoir properties and dynamic property changes were they propose a methodology

to jointly estimate rock properties (Porosity) and dynamic property changes (pressure and saturation

changes) form time-lapse seismic data. This methodology is based on a full Bayesian approach to seismic

inversion and it is done in two steps: First the estimation of conditional probability of elastic properties and

their relative changes, then the estimation of posterior probability of rock properties and dynamic property

changes (Dario Grana and Tapan Mukerji, 2015). The result is a set of point-wise probability distributions

that are used to predict the most probable reservoir models and evaluate the associated uncertainty.

The proposed methodology of this work is based on a family of geostatistical inversion methodologies

that use the principle of cross-over of genetic algorithms (Azevedo et al., 2013b).

22

23

Chapter 4: Methodologies

4.1 Geostatistical elastic inversion: Global Elastic Inversion

This chapter introduces the global elastic inversion (GEI) (Figure 7), since it is the basis for the proposed

approach, followed by the proposed methodology developed under the scope of this thesis that will be

presented in the next section.

The Global Elastic Inversion (GEI) was developed and implemented by Nunes et al. (2012), and consists

of an inversion methodology of Ns partial stacks for acoustic and elastic impedance. This method of

inversion used the same principles of GSI: it uses Direct Sequential Simulation DSS and Co-DSS with joint

probability distributions in order to generate acoustic and elastic impedance models ensuring the

reproduction of the relationship between the original properties as inferred from the existing well data (Nunes

et al., 2012)

This methodology can me summarized in the following sequence of steps (Nunes el al., 2012; Azevedo,

2013):

I. Simulation of a set of Ns models of acoustic impedance with DSS algorithm, conditioned

exclusively by the well data related to this property;

II. Co-simulation of a set of Ns models of elastic impedance, related to the individual acoustic

impedance simulated models, using DSS with a joint probability distributions. Each model of elastic

impedance is conditioned by the EI data from the well jointly with the AI model previously simulated;

III. From the Ns model of AI and EI Ns of synthetic partial angles-stacks are computed using the Fatti

et al. (1994) approximation. Ns synthetic partial angle stacks are generated by convolution with the

corresponding angle-dependent wavelet.

IV. Comparison between each seismic trace form all the synthetic partial angle stacks with the

corresponding real seismic trace, in terms of correlation coefficient.

V. From the initial ensemble of pairs of AI and EI models, the pair of impedance traces that jointly

produce a synthetic seismic trace with the highest correlation coefficient, when compared with the

corresponding observed seismic trace, are stored in two auxiliary volumes of best acoustic and

elastic impedance models, along with the respective local correlation coefficient. The best volumes,

together with the local correlation volumes, are then used as secondary variables in the co-

simulation process for the generation of a new set of impedance models during the next iteration.

24

After iteration one the ensemble of AI models is simulated conditioned to the available acoustic

impedance well-log data, the best AI and the local correlation volumes from the previous iteration.

In the same way EI models generated after iteration one, are co-simulated using previously

simulated AI model, as auxiliary model, conditioned to the available elastic impedance well-log data

and the best EI and local correlation volumes from the previous iteration.

VI. Return to the step I. till the global correlation coefficient between the synthetic and the recorded

partial angle stacks, for all available n angle stacks, is above a certain threshold.

Figure 7 - Schematic representation of the GEI workflow (Azevedo, 2013).

4.2 Geostatistical Time-lapse Inversion

The proposed methodology of this work is also based on the group of iterative geostatistical inversion

methodologies that use the principle of cross-over of genetic algorithms, as it was referred previously. This

approach can be considered an extension of the previous explained methodology (GEI), once the properties

of interest to invert are the same (acoustic and elastic impedances). The main goal of the proposed method,

is to access the dynamic behavior of the elastic properties in the reservoir, therefore the main advantage is

the use of a procedure to invert simultaneously different seismic reflection vintages directly to the properties

25

elastic differences. Due to this reason, we need to deal with the elastic changes that will correspond to our

hard data, traduced by the computation between the differences of the properties, associated to each

acquisition time during the production of the reservoir. The final inverted models are constituted by the

residuals representing a set of equiprobable scenarios of the elastic response that may allow us to access

and quantify the spatial uncertainty associated. Once the elastic changes are not verified in all the study

area as we can see through the seismic differences illustrated in Figure 1, it were used masks within the

inversion grid, in order to simulate exclusively in the elastic changes areas, avoiding the simulation of

residual values in zones where the changes were not noticed.

The general workflow of the proposed methodology is represented in the Figure 8 and it can be briefly

described by the following sequence of steps:

1) Calculate from the existing hard-data differences between well-log measurements related to each

different seismic vintage for each property to invert. These data will be used as conditioning in the

simulation procedure;

2) Stochastic sequential simulation of a set of Ns acoustic and elastic impedance models using DSS

and co-DSS with joint probability distributions respectively. Both properties are conditioned to the

available well log data and, in the case of elastic impedance, will also be conditioned to the previous

acoustic impedance model. For the acquisition time (T0) corresponding to the beginning of the

production, the well log data used are the original values of the elastic properties, once there are

no historic in elastic changes. To note that, the residues of each time are simulated in a masked

grid, conditioned by the differences between the seismic of each year: the zones which don’t occur

amplitude changes in the seismic data, are not simulated and are filled with the data from the

acoustic and elastic impedance models from the first year, in order to have the full grid to calculate

the synthetic seismic data.;

3) After the sum between the pair of the initial inverted models (T0) and the residuals from each time

of production (T10, T25 and T30), calculate synthetic seismic data volumes for those pairs of

realizations of each time generated in the previous step. The synthetic seismic data is obtained by

the convolution between the angle-dependent wavelets and the reflectivity coefficients computed

by Fatti´s approximation for each angle gather;

4) Compare the synthetic seismic volumes, in a trace-by-trace basis, with the corresponding

recorded seismic data in terms of correlation coefficient where the final result will be a best local

correlation cube that store the traces that produced the highest correlation coefficient. This process

of seismic comparison is applied to all times simultaneously obtaining, at the end of each iteration,

eight local correlation cubes: four volumes to each properties, each pair associated to each

acquisition time. From the best traces selected simultaneously and stored in the best correlation

coefficient volumes it will be computed the correspondent best elastic volumes for acoustic and

26

elastic impedances residuals. The best elastic residuals volumes along with the local correlation

coefficient volumes, after being subtracted by the T0 properties models, are used as secondary

variables in the co-simulation process for generating a new set of impedance residual models in the

next iteration. The new ensemble of impedance residual models, for the next iterations, will be

conditioned by the available well-log data (and each impedance model at each realization, in the

elastic case) but also by the best volumes (best elastic and local correlation cubes) computed in

the previous iteration. The iterative geostatistical inversion procedure finishes when global

correlation coefficient between the synthetic average and the recorded partial angle stacks, for all

the available n angle stacks, is above a certain threshold;

27

Figure 8 - Graphic representation of the geostatistical time-lapse inversion.

0 5 10 15 20 25 30 35-6

-4

-2

0

2

4

6

8

10x 10

5

10º

20º

30º

40º

����� = (�4 − ��) ����� = (�� − ��) ����� = (�� − ��) ����� = (�4 − ��) ����� = (�� − ��) ����� = (�� − ��)

AI T0 EIT0

∆AI ResT1 ∆AI ResT2

∆AI ResT3

∆EI ResT1

∆EI ResT2

∆EI ResT3

Real T0 Synthetic T0

Synthetic T1

Synthetic T2

Synthetic T3

Real T1

Real T2

RealT3

Best CC T1

Best CC T0

Best CC T2

Best CC T3

Best CC T1

Best CC T0

Best CC T2

Best CC T3

Best AI T0

Best AI T1

Best AI T2

Best AI T3

Best EI T0

Best EI T1

Best EI T2

Best EI T3

Calculation of the differences between well-log data of each

Stochastic Simulation of AI and ∆AI Stochastic Co-Simulation of EI and ∆EI

Fatti’s approximation Forward modelling

Seismic

comparison

per angle

Sim

ula

te u

ntil th

e n

um

be

r o

f re

aliz

ation

s is r

ea

ch

ed.

Selection of best local correlation cubes for both properties

Selection of best AI and EI cubes

28

29

Chapter 5: Case Study

5.1 Data Description

The synthetic dataset used in this thesis for testing the proposed methodology, introduced in the previous

chapter, was the Stanford VI-E (Lee and Mukerji, 2012) is composed by a three-layer prograding fluvial

channel with a structure that consists of an asymmetric anticline with axis N15ºE. It intends to reproduce a

realistic siliciclastic environment with meandriform channels. It is composed by a grid with the dimensions

of 150×200×120 in I, j and k directions, respectively. Time-lapse seismic data was computed from the

available elastic properties (density, Vp and Vs) at the beginning and each tenth year until 30 years of

production and partially stacked seismic data with 5 mean partial angle stacks ( 0º, 10º, 30º, and 40º). Also

part of the dataset are 31 wells with respective well log data of elastic impedance values. The original three-

dimensional volumes of AI and EI that were used exclusively to compare with the Seismic inversion results.

For testing purposes the original tridimensional grid that was reduced to 150×200×66 in the directions I, j

and k. The Figure 9 shows the original Acoustic impedance models for the 4 time steps considered, i.e.,

during the production of 30 years.

Figure 9 -Real models of acoustic impedance from each year of acquisition, T0, T10, T25 and T30 respectively.

T0

T30

T10

T25

30

Figure 10 shows the evolution of EI variable along the 30 years of production. The changes in the true

impedance values are small because the S-wave velocity does not propagate through liquid media being

the only change associated with the density of the fluid filling the pore space.

Figure 10- Real models of elastic impedance from each year of acquisition, T0, T10, T25 and T30 respectively.

Although it is difficult to notice some significant changes between different years of seismic acquisition,

those can be seen in the residuals computed by subtraction between each pair of acquisition times (Figure

1). Figure 11 presents the seismic reflection data associated to each acquisition time since the beginning

of the production. They are divided in 5 partial angle stacks: near-stack (0º), near-midstack (10º), mid-stack

(20 º), far-midstack (30º) and far-stack (40º). Within each time, it can also be noticed some amplitude

changes with the increasing of the offset distances as we can verify in the channel especially in the further

partial angle stacks.

T0 T10

T25 T30

31

Figure 11 - Representation of the partial angle stacks form seismic reflection data, 0º, 10º, 20º, 30º and 40º

respecively.

T0 T10 T25 T30 N

ea

r-sta

ck

Ne

ar-

mid

sta

ck

Mid

-sta

ck

Fa

r-m

id

sta

ck

Fa

r-sta

ck

32

The figure below (Figure 12) represents the wavelet used in this work. The angle-dependent wavelets

were used to generate the true real partial angle stacks for the 30 years of production.

Figure 12 – Representation of the stack wavelets correspondent to the 5 angle stacks.

5.1.1 Structure and stratigraphy

The Stanford VI-E reservoir is composed by a classical structural oil trap, an anticline, that is an

asymmetric anticline with axis N15ºE (Figure.13) and it has a different dip on each flank generally the dip

decreases slowly towards the northern part of the structure.

33

Figure 13 – Perspective view of the Stanford VI-E top structure: view from SW (left), view from SE (right).

The color indicates the depth to the top (Lee & Mukerji 2012).

The reservoir used in this study is 3.75 Km wide ( East-West) and 5 Km long (North-South), with a

shallowest top depth of 2.5 Km and deepest top depth of 2.7 Km. It haves 200 m of thickness and consists

of three layers with thicknesses of 80 m, 40 m and 80 m respectively.

The stratigraphy of the reservoir represents a prograding fluvial channel system and it can be divided in

3 layers, where deltaic deposits in the layer 3 were deposited first and followed by meandering channels in

layer 2 and sinuous channels in layer 1. This sequence of clastic deposits represents a progradation of a

fluvial system into the basin located in this case toward the north of the reservoir.

The layer 1 of this data set consists in sinuous channels where two facies are represented: Shale

deposits, overbank fine sands along the inner edges of the meandric channels where the average channel

thickness is 20 meters, average width 600 meters and boundary thickness is 1.5 meters. The second layer

also consists of meandering channels that have the same deposits as the first layer with average channel

thickness is 300 meters, average width 26 meters and boundary thickness is 1.5 meters. The last and third

layer of the reservoir consist of deltaic deposits and are represented by two facies: Shale deposits and

channel sand deposits with a channel thicknesses that varies between 7 and 40 meters and channel widths

from 70 till 400 meters.

34

5.1.2 Exploratory data analysis

In the methodologies tested under the scope of this thesis, there are three characteristics that need to

be present in each realization of AI and EI: the representation of the values of the experimental data at its

location, the reproduction of the global distributions functions as inferred from the experimental hard data

(Figures 14 and 15) and the spatial continuity patterns traduced by the variogram model.

Figure 14 - Histograms of acoustic impedance related to each well-log data per year of acquisition.

As we can see, the four histograms representing the frequency classes of AI values (Figure 14)

correspondent to the pre-production and post-production scenarios (after 0 , 10, 25 and 30 years of

production) They have a bimodal distribution what is expected due to the fluvial environment: Sands and

Shales. These two facies with different geological properties, present different classes of values as the

parameter of interest is define by the next equation:

�� = �� = �v ×   (16)

35

In the equation 16, Ip means primary impedance, AI acoustic impedance, Vp velocities of primary waves,

and   it’s the density. It is expected that we have higher values of AI in less saturated deposits like in shales

(7800-8600 KPa.m/s), and lower values in sands (5200-6000 KPa.m/s), that are the reservoir rock.

Figure 15 - Histograms of elastic impedance related to each well-log data per year of acquisition.

For EI case(Figure 15) we still have bimodal distribution of frequency values but with a narrower range,

with classes of values from 2400 kPa.m/s and 5400 kPa.m/s and the two peaks situated between 3000-

3240 kPa.m/s and 4200-4560 kPa.m/s. As expected the values are much lower than the acoustic

impedance, because we know that the secondary seismic waves are slower than the primary waves, we

can see also that the values of Elastic Impedance don’t change much along the 30 years of oil production.

�u = ¡� = � ×   (17)

The more information exists about the petro-elastic properties of the subsurface the better the reservoir

modeling and characterization. Then it is necessary to study and understand the relationship between the

properties of interest, the study of bi-distribution or joint distributions (Figure 16). To get a better and more

36

realistic reservoir characterization it’s important to ensure the reproduction of the relations between the

parameters in study, in the best pairs of inverted models in the GEI and Geostatistical Time-lapse Inversion.

Figure 16 - Well-log data bi- histogram of each acquisition time, expressing the relation between the two

properties of interest (AI and EI).

Understanding the spatial behavior of the phenomenon under study is a key step in the exploratory data

analysis of our data. The properties of interest were modeled with a spherical variogram model (Figures 17

and 18; Tables 1 and 2) for each time-lapse considered (0, 10, 25 and 30 years) in three main directions

North-South (0.0), East-West (90.0) and in the vertical direction (0.90). The vertical direction presents a

greater spatial continuity than the other two horizontal directions due to the low number of wells and the

large distance between them. As such, to model the spatial continuity of each property, it was used

omnidirectional horizontal variograms and variograms with direction perpendicular to these with 180º of

tolerance, 5 of lag distance and 1000 of cut distance. Unlike the horizontal directions, the vertical

experimental variograms were easier to calculate due to high vertical resolution and the number of samples

supplied by the data logs of the wells. Spherical models have been used with one structures small amplitude

ranges and none of the variogram models arrived to the sill due to the great heterogeneity of area of study,

37

so in terms of spatial distribution the properties are non-stationary. These variogram parameters were used

on the procedure of both sequential simulation algorithms. To make the modelization of the variograms

possible it was utilized the software “geoMS” (Geostatistical Modelling Software).

Figure 17 - Chosen main direction of AI for modeling, with exponential model.

Table 1 - Values of the modeled variograms.

Figure 18 - Chosen main direction of EI for modeling, with exponential model.

IpBase

(Sill=1313688,625)

(0:0)

(90:0)

(0:90)

C1 1177097 1177097 1258572

a1 22 22 20

38

Table 2 - Values of the modeled variograms

IsBase

(Sill=253782,125)

(0:0)

(90:0)

(0:90)

C1 248144 248144 241233

a1 39 39 25

5.1.3 Parameterization

The methodology tested in this work was the Geostatistical Time-lapse Seismic Inversion, where the

differences between the elastic properties were estimated through the seismic and well measurements

differences presented in the time-lapse data during the years of production. The dataset available to this

application example is composed by the available well-log data from the properties to invert (Acoustic

Impedance and Elastic Impedance); the joint-probability distributions functions between both properties of

each year considered; the different angle-dependent wavelets, as well as the seismic differences per angle

stack between each year. As we point to the study of differences of the elastic properties, it is important to

invert and simulate only the regions that changes of recorded time-lapse seismic data when compared with

previous available vintage (figure 19) . As introduced in the methodology section for the initial time, AI and

EI are simulated in the original domain, while for the remaining years the residuals of both properties

between consecutive vintages are the simulated properties (Figure 20). It is important to note that for the

inversion of time T1, there is no mask (T0T1) because all, or almost all the grid suffered some kind of change

(Vp, density and Vs), fact that is understandable due to beginning of hydrocarbon production.

The seismic inversion methodology was run with the following parameterization: 6 iteration, with 32

simulations for each time, so in each iteration there are 256 simulated scenarios of Elastic impedance and

Acoustic impedance related to each acquisition time.

39

Figure 19 - Illustration of the simulation mask, T1T2, T2T3 respectively. The red color represents the places to be

simulated.

Once the residual models are simulated, as shown in the chapter IV, before the creation of synthetic

seismic, they are added to the generated model of T0 of the same simulation to bring the simulated

properties back to the original domain and to fill in the cells that were not simulated, i.e., areas where there

are no differences in terms of seismic reflection data.

Figure 20 - Inverted residual models from T10, T25, and T30 respectively.

40

5.2 Geostatistical Time-lapse Inversion

This section presents the results from the geostatistical elastic time-lapse inversion, with the goal to

compare those, with the real models of the properties of interest (AI and EI), the synthetic seismic from the

best models as well as the synthetic seismic from the average of the models from the last iteration with the

real seismic. It will also be presented the distributions functions, traduced by the histograms and bi-

histograms from the well log data with the best fit models of the properties to invert. All these comparisons

are done for each year considered from two horizontal sections.

5.2.1 Synthetic seismic

In this work used 5 partial angle stacks were used as recorded seismic time-lapse data. Because the

differences between them are not significant, and for the sake of simplicity, it will only be shown the

nearstack volume. Figure 21 shows the evolution of the global correlation coefficient between the synthetic

and the real seismic reflection data for all the four vintages available and Figure 22 shows the real seismic

followed by the synthetic seismic from the best fit models and the synthetic seismic data computed from the

average of the 32 models of the last iteration of each year of acquisition. The synthetic seismic from the

best models help us to understand how good the match between the real seismic was especially if the

producing areas where the well data density is higher. The synthetic seismic from the mean intend to

conserve the variability from the models of the last iteration having a smoothness of the possible scenarios

simulated given less importance to the extreme values what results in a better spatial continuity.

Figure 21 - Global correlation coefficient evolution between synthetic and real seismic reflection data for

the four available seismic vintages using geostatistic time-lapse inversion method.

41

Figure 22 - Comparison of all angle stacks between the real seismic (left), the synthetic seismic from the best-fit

models (Center) and the Synthetic seismic computed from the mean of the models from the last iteration. The seismic

results are related to the different times of acquisition divided by rows, being the first related to the year 0 and the last

to the year 30.

42

5.2.2 Acoustic and Elastic Impedance Models

The acoustic and elastic impedances are the properties of interest inverted in this work, which were

simulated and co-simulated respectively, Figure 23 and Figure 24 show the comparison between the real

models of acoustic impedance and elastic impedance, respectively, with the best-fit inverted models of the

last iteration for these two properties. It is clear the presence of the fluvial channel of the left and the shapes

of the channels in the middle, a little less evident in the elastic impedance models.

43

Figure 23 - Comparison of the near angle stack between the real acoustic impedance models (Right), the

simulated best-fit acoustic impedance models (Center) and the mean of the simulated acoustic models computed

from the last iteration. The first line is related to the year 0 and the other too years 10, 25, and 30 respectively.

44

Figure 24 - Comparison of the near angle stack between the real elastic impedance models (Right), the simulated

best-fit elastic impedance models (Center) and the mean of the simulated elastic models computed from the last

iteration. The first line is related to the year 0 and the other too years 10, 25, and 30 respectively.

45

5.2.3 Histograms, Bi-plots and Statistics

This section introduces the results of this work regarding the statistical analysis from the best-fit models

illustrated previously. It comprises the probability distribution functions and the joint probability distribution

functions traduced by the histograms and the cross-plots between both properties, respectively, for each

acquisition time. Also the main statistics presented in the Table 3 and Table 4. These elements are very

important to retrieve from the well-log data as they allow us to evaluate, statistically, the global behavior of

the inverted model in to obtain a more reliable reproduction of the sub-surface geology. The joint and

marginal distributions in the bi-histograms are also very important to honor in the inverse models once, in

reservoir characterization, the extreme values of the properties need to be well reproduced. The Figures

25 and 26 represents the histograms from the well-log data and the best-fit models of acoustic and elastic

impedance respectively, and in the Figure 27 the cross-plots between the AI and EI of the well-log data and

the best-fit models.

46

Figure 25 - Acoustic impedance histograms from the well-log data of each year against the histograms from the

best-fit models of each acquisition year.

47

Figure 26 - Elastic impedance histograms from the well-log data of each year against the histograms from the

best-fit models of each acquisition year.

48

Figure 27 - Bi-histograms of well-log data against best-fit models of each year, expressing the relation between

the two properties of interest, AI and EI.

49

Table 3 – Acoustic impedance main statistics of well-log data and best-fit models related to each year.

Acoustic Impedance

Main

statistics T0 T10 T20 T30

Well-log data

Mean 7103,43 7069,09 7069,07 7071,93

Variance 131709,44 130889,44 130682,45 129718,39

Std.Deviation 1147,65 1144,07 1143,16 1138,94

Minimum 4517,99 4544,71 4544,43 4544,04

Maximum 9190,18 9124,72 9124,72 9124,72

Best-fit model

Mean 7083,73 7057,82 7060,83 7063,59

Variance 561870,49 562662,59 56157,31 562708,51

Std.Deviation 749,85 750,57 749,77 750,14

Minimum 4517,99 4482,02 4483,58 4470,25

Maximum 9190,18 9346,71 9345,11 9328,54

Table 4 – Elastic impedance main statistics of well-log data and best-fit models related to each year

Elastic Impedance

Main

statistics T0 T10 T20 T30

Well-log

data

Mean 3698,03 3698,12 3698,14 3698,99

Variance 255197,28 254368,98 254206,21 253411,77

Std. Deviation

505,17 504,35 504,19 503,40

Minimum 1925,09 1923,97 1923,97 1923,97

Maximum 5042,87 4552,68 4552,68 4952,68

Best-fit

model

Mean 3461,91 3463,51 3461,92 3463,05

Variance 126058,18 126378,44 126372,48 126395,53

Std. Deviation

355,47 355,49 355,48 355,52

Minimum 1925,09 1923,98 1923,81 1923,97

Maximum 5042,87 5042,87 5042,87 5042,87

50

5.3 Comparison between geostatistical elastic inversion and geostatistical time-lapse

inversion methodologies

This section, compares the seismic reflections and elastic models from the two geostatistical

methodologies used in this work. It aims show the efficiency of the proposed geostatistical time-lapse

seismic inversion methodology

5.3.1 Differences between seismic

In this section it is presented the comparison between seismic reflection data, and as explained before,

the 5 partial angle stacks used, did not show significant differences between then, therefore, after computing

synthetic seismic of the best-fit models in the geostatistical elastic inversion procedure and the synthetic

seismic from the mean of the best-fit models using the geostatistical time-lapse inversion, it was only used

the near stack to do the comparisons. As previously stated, using the synthetic seismic of the mean of the

best models helped us to grant a higher spatial continuity. Figure 28 illustrate the differences between real

seismic data and the synthetic seismic data from both methodologies.

51

Figure 28 - Comparison the near angle stacks between the real seismic (left column), the Synthetic seismic

computed from the mean of the models from the last iteration using Geostatistic Time-lapse inversion (center

column) and the synthetic seismic from the best-fit models using GEI (right column). The seismic results are

related to the different times of acquisition divided by lines, being the first related to the year 0 and the last to the

year 30.

52

5.3.2 Acoustic and Elastic Differences

In this section it’s presented the main goal of this work, to evaluate if the real changes in elastic properties

are being assessed and propagated during the dynamic changes in the reservoir. Therefore in the Figures

29 and 30 are shown the real differences between the properties of interest (AI and EI), the mean of the

best models using geostatistic time-lapse inversion method and the differences between the best models

using the geostatistical elastic method. This differences are done subtracting the previous models by the

models of the next year of acquisition, so it is expected to see lower values in production zones. The main

objective by comparing this differences, is to see, which applied method can reproduce better the spatial

location of the maximum and the minimum values.

Figure 29 - Acoustic impedance differences from: the real (left), geostatistical time lapse inversion (middle) and

global elastic inversion (right) for the T0-T10, T10-T25, and T25-T30 years respectively

53

The Figure 30 represents the differences between elastic impedance models of each acquisition year.

Like in the previous figure, is used the mean of the best models inverted by the method of geostatistic time-

lapse seismic inversion against the best geostatistical elastic inversion models, and the real differences

from the real data that is supposed to address. This differences are much difficult to see when compared

with the acoustic impedance showed previously, even so, the main goal of this comparisons can be

achieved, the reproduction of the real changes on the inverted models.

Figure 30 - Elastic impedance differences from: the real (left), geostatistical time lapse inversion (middle) and

global elastic inversion (right) for the T0-T10, T10-T25, and T25-T30 years respectively

54

55

Chapter 6: Discussion, Conclusions and Future work

In this last chapter are discussed the results obtained using the geostatistical time-lapse inversion

compared to the real ones, and also the main results between the application of the proposed and the

geostatistical elastic inversion approaches. In the last part, are presented some final remarks of this work,

along with some suggestions associated with the future work in order to improve the results and the

efficiency of the proposed methodology.

The geostatistical time-lapse inversion methodology presented a reasonable global correlation

coefficient evolution (Figure 21) between the real seismic data and the synthetic seismic data of the inverted

elastic models. Although the convergence of the methodology was not high as we expected, this fact can

be explained regarding the main goal of this work, once we are inverting several seismic reflection data

acquisitions simultaneously, implying the integration of more set parameters within the objective function.

Nevertheless we obtained good representations of the synthetic seismic data (Figure 22) showing that, even

presenting some noise in the data, the main amplitude variations of the seismic are well illustrated by the

channels of the best-fit model. Computing the synthetic seismic data of the mean of the elastic models, lead

to a decreasing of the noise and consequently a better approximation to the real seismic data.

Attending to the obtained acoustic impedance models presented in Figures 23 it is possible to see that

the channel in the right side is very well retrieved comparing with the real one. The other channel in the

middle, is not so well reproduced like the previous one, fact that can be explain because of two reasons:

the low thickness of the channel, what reduces the seismic reflectivity area and the less well log data

available in that zone. Although these evidences, we can see its boundaries traduced by the lower acoustic

impedance values defining the main trend of the channel. The properties values on the zones between the

channels are not so closer to the real models, however the most important zones are the channels areas,

which correspond to the hydrocarbon accumulations when we are dealing with siliciclastic environments.

The same results are extended to the elastic impedance models (Figure 24), where we can still see the

channel in the right side but the channel in the middle is more difficult to distinguish, once this property is

harder to retrieve than the acoustic impedance in seismic inversion methodologies.

Playing a key role in what seismic inversion methodologies concern, the statistic elements and the

probability distribution functions are very important to be well reproduced in order to obtain more reliable

inverse models. As we can see from the histogram displayed in Figures 25 and 26 the main populations are

well reproduced comparing with those retrieved from well-log data for each property associated with each

acquisition year. One of the reasons that the population corresponding to the lower AI values are not so well

reproduced in the best-fit models histograms is that the well-log data presents less number of samples then

the best-fit models that comprise the values from all the simulation grid. Attending to the best-fit elastic

impedance histograms, it is important to highlight that the central class of values (around 3500 kPa.m/s) is

56

composed by the greatest number of samples, as in the case of well-log data histograms. The

representation of the joint probability distribution functions, traduced by the bi-plots (Figure 27), is also

satisfactory showing the trends of the relationship between the acoustic impedance against the elastic

impedance properties when compared with those retrieved from the available well-log data. In Tables 3 and

4, are presented the statistic concerning the well-log data and best-fit models related to each year. In general

the statistics of the inverse models are well reproduced compering with the well-log data, especially the

minimum and the maximum values of the data. For the other statistics to illustrate a better similarity, the

global convergence of the proposed methodology should present a higher value and the seismic data should

also be better inverted in all the simulation grid.

Moreover and in order to evaluate the performance of the proposed methodology, were also run a global

elastic inversion for each year separately conditioned by the same data set but directly inverting to the

properties themselves and not the elastic changes. The results obtained from both methodologies were

compared between them regarding the seismic reflection data and the properties differences between each

acquisition year.

Although the synthetic seismic is similar for both methods (Figure 28), the boundaries of the represented

channels are much better reproduced using the proposed methodology when compared with GEI. Also the

non-channel zones have more amplitude values close to the real seismic, as the GEI synthetic seismic have

more values close to zero and therefore different from the real one. Concerning the Figures 29 and 30, the

horizontal time-slices of elastic differences computed from the global elastic inversion present no spatial

continuity once the elastic difference values are not in the right locations in the simulation grid. Otherwise

the proposed method, even also do not show a good spatial continuity, we can verify that the range of elastic

changes values is honored, appearing in the right locations and within the property boundaries. The zones

where no elastic changes are noticed, are much better reproduced in the residual models of the proposed

methodology, giving us more reliable models regarding the study of the spatial uncertainty of elastic

differences. This better reproduction of elastic changes during all the years of production, can be explained

by the fact that the proposed methodology is able to simultaneously integrate the different seismic reflection

data, allowing the link between the differences from the previous and the next year of the acquisition time,

and also the fact of inverting directly to the elastic changes. Another great advantage when compared to

other seismic inversion algorithms (deterministic and Bayesian), is the possibility to presenting scenarios as

solutions, allowing a better assessment and quantification of the spatial uncertainty related with the geology

subsurface. All this advantages come with some limitations especially the computational costs and time

consuming once we are doing several seismic inversions simultaneously. Another fact that can greatly limit

the proposed methodology is associated to be a recent and in developing area and that need to have more

seismic acquisitions during the reservoir production stage. The fact that this methodology strongly depends

on the available 4D seismic data acquisition drives to a higher exploration costs and consequently the need

of a better improvement and knowledge regarding a study area. For these reasons, the development of this

57

proposed methodology would need more research and time investment to improve the optimization of the

workflow and the constraints in this approach.

As future work, we can go through this problem, integrating the petro-physical properties to build reliable

rock physic models, when we have more available wells-log data, in order to accurately estimate the elastic

changes of the properties and their dynamic behavior along the oil field life time.

59

References

Aanonsen, S. (2003). Effect of scale dependent data correlations in an integrated history matching loop

combining production data and 4D seismic data.

Avseth, P., Mukerji, T., & Mavko, G. (2010). Quantitative seismic interpretation: Applying rock physics tools

to reduce interpretation risk. Cambridge university press.

Avseth, Per, Mukerji, T., & Mayko, G. (2005). Quatitative Seimic Interpertation.

Azevedo, L. (2013). Geostatistical methods for integrating seismic reflection data into subsurface earth

models. Lisboa.

Bosch, M., Mukerji, T., & F.Gonzalez, E. (2010). Seismic inversion for reservoir properties combining

statistical rock physics and geostatistics : a review.

Buland, A., & El Ouair, Y. (2006). Bayesian time-laspe inversion.

Caetano, H. (2009). Integration of seismic information in seismic inversion.

Duijndam, A. (1988). Bayesian Estimation in Seimic Inversion. Part I : Principles Geophysical Prospecting.

Fatti, J., Smith, G., Vail, P., Strauss, P., & Levitt, P. (1994). Detection of gas in sandstone reservoirs using

AVO analysis: A 3-D seismic case history using the Geostack technicque. Geophysics, 59(9), 1362-

1376.

Francis, A. (2006). Understanding Stochastic Inversion : Part 1."First Break24 (November): 79-84.

Furre, A., Bakken, E., Kloy, T., & Nordby, L. (2006). Time-lapse seismic project integrating geophysics and

reservoir engineering.

Goovaerts, P. (1997). Geostatistics for Natural Resources and Evaluation. Geostatistics for Natural

Resources Evaluation. Oxford University Press.

Grana, D., Mukerji, T., Dvorkin, J., & Mavko, G. (n.d.). Stochastic Inversion of Faceis from Seismic Data

Based on Sequential Simulations and Probability Perturbation Method.

Horta, A., & Soares, A. (2010). Direct Sequential Co-simulation with Joint Probability Distributions.

Mathematical Geosciences.

Lafet, Y., Duboz, P., Deschizeaux, B., Lefeuvre, F., & Hubans, C. (2005). 4-D Stratigraphic Inversion of the

Girassol Field-Towards a Mora Quantitative Approach.

60

Lancaster , S., & Whitcombe, D. (2000). Fast track coloured inversion.

Landro, M., Veire, H., Duffaut, K., & Najjar, N. (2003). Discrimination between pressure and fluid saturation

changes from marine multicomponent time-lapse seismic data.

Lee, J., & Mukerji, T. (2012). The Stanford VI-E reservoir: A synthetic data set for joint seismic-EM time-

lapse monitoring algorithms.

Nunes, R., Schwedersky, G., Dillon, L., Guerreiro, L., Caetano, H., Maciel, C., & Leon, F. (2012).

Geostatistical Inversion of Prestack Seimic Data. .

Pereira, P. (2014). Inversão Geoestatística de Dados Sísmicos 3D Multi-distribuições Locais. Tese de

Mestrado em Engenharia Geologica, Universidade de Aveiro, Departamento de Geociencias ,

Aveiro.

R.O., L. (1979). Sysnthetic sonic logs-a process for startigraphic interpetation Geophysics.

Russel, B., & Hampson, D. (1991). Comparison of Poststack Seismic Inversion Methods.

Russel, B., & Toksoz, M. (1991). Comparison of poststack seismic inversion methods.

Russell, & H., B. (1988). Introduction to seismic inversion methods .

Sarkar, S., Gouveia, W., & Johnston, D. (2003). On the inversion of time-lapse seismic data.

Soares, A., Diet, J., & Guerreiro, L. (2007). Stochastic Inversion with a Global Perturbation Method.

Petroluem Geostatistics. Cascais.

Tarantola, A. (2005). Inverse Problem Theory.

Tompkins, Michael , J., Juan, L., Fernández, M., Alumbaugh, L., & Mukerji, T. (2011). Scalable Uncertainty

Estimation for NonLinear Inverse Problems Using Parameter Reductionn, Constraint Mapping, and

Geometric Sampling : Marine Controlled-Source. .

Ulrych, T. J., Sacchi, M. D., & Woodbury, A. (2001). A Bayes tour of inversion: A tutorial. Geophysics, 66(1),

55-69.

61

Appendix A

Parameterization for the first acquisition time (T10 acoustic impedance).

Ip10

(Sill=1305586,25)

(0:0)

(90:0)

(0:90)

C1 969501 969501 1258572

a1 16 16 20

Parameterization for the first acquisition time (T10 elastic impedance).

Is10

(Sill=253137,609)

(0:0)

(90:0)

(0:90)

C1 212183 212183 241233

a1 38 38 25

Parameterization for the first acquisition time (T25 acoustic impedance).

Ip25

(Sill=1303517,125)

(0:0)

(90:0)

(0:90)

C1 973570 973570 1264892

a1 18 18 23

Parameterization for the first acquisition time (T25 elastic impedance).

Is25

(Sill=252974,781)

(0:0)

(90:0)

(0:90)

C1 233628 233628 241153

a1 40 40 20

Parameterization for the first acquisition time (T30 acoustic impedance).

Parameterization for the first acquisition time (T30 elastic impedance).

Is30

(Sill=252181,797)

(0:0)

(90:0)

(0:90)

C1 238116 238116 245058

a1 46 46 20

Ip30

(Sill=1293886)

(0:0)

(90:0)

(0:90)

C1 1292078 1292078 1242814

a1 29 29 20


Top Related