Liu Thesis 2020

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 161

Copyright

by

Chuxi Liu

2020
The Thesis Committee for Chuxi Liu
Certifies that this is the approved version of the following Thesis:

Automatic History Matching with Data Integration for Unconventional


Reservoirs

APPROVED BY
SUPERVISING COMMITTEE:

Kamy Sepehrnoori, Supervisor

Wei Yu
Automatic History Matching with Data Integration for Unconventional
Reservoirs

by

Chuxi Liu

Thesis
Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Master of Science in Engineering

The University of Texas at Austin


December 2020
Dedication

To my beloved parents, Ziyun Liu and Aimin Ma for their unconditional love and

encouragement.
Acknowledgements

I would especially express my greatest gratitude toward my advisor, Dr. Kamy

Sepehrnoori, for his continuous supports and guidance for the entire period of my

master’s degree. His resourceful feedbacks and comments on the weekly meeting are

extremely helpful for me to envision the industry insights and also to build a more

rigorous thought process. It is my greatest pleasure and honor to work under this research

group, which is the leading research group in unconventional resource studies in the

entire United States.

Secondly, I am very thankful for Dr. Wei Yu. His professional expertise,

unconditional suggestions and rigorous trainings helped me gain invaluable experience

both in perspective of academia and industry. I am grateful for his utmost patience, trusts,

and sincere emotional support whenever I am facing a challenge. Most importantly, I am

grateful for his review and writing advices given towards the completion of this thesis.

My deepest credit also extends to previous graduate students, Marut Wantawin,

Silpakorn Dachanuwattana, and Sutthaporn Tripoppoom for building a solid foundation

on automatic history matching algorithm. I also appreciate the help and supports from our

research group students for skill sharing, especially Esmail Eltathan for his tutorial on the
fracture modelling and reservoir modelling and his knowledge regarding the numerical

optimization.

Finally, I want to acknowledge the great assistances from my UT PGE friends for

their care and helps, so that I can experience a wonderful life in Austin. In addition, I

want to appreciate the emotional and financial supports from my parents. Without them,

it would be difficult to obtain my achievements thus far.

v
Abstract

Automatic History Matching with Data Integration for Unconventional


Reservoirs

Chuxi Liu, M.S.E.

The University of Texas at Austin, 2020

Supervisor: Kamy Sepehrnoori

Given the dynamic production data of a reservoir, numerical optimization tools

such as history matching can minimize the global error and find an optimal reservoir

model that can approximate the fracture geometry and petrophysical parameters in the

subsurface. For unconventional reservoirs, the idea behind the automatic history

matching is well developed and the workflow is also applied to statistically generate an

ensemble of solutions that quantitatively characterizes associated uncertainties. However,

more uncertainties regarding fracture and reservoir properties could be further reduced by
using available information. Therefore, the objective of this study is to minimize

uncertainty when make realizations of shale reservoirs, by integration of data from

geology and geomechanics.

We utilized the developed automatic history matching (AHM) code and modified

the proxy engine, by substituting the neural network (NN) model with XGBoost

(XGBOOST) model. The XGBOOST is found to perform more efficiently and accurately

than NN, when the size of the available dataset for training is small. Furthermore, the
vi
AHM workflow is capable of modelling non-uniform half-length of hydraulic fractures in

the corner point gridding system and complex, realistic natural fracture distributions

using the fractal theory. Both of these functionalities partially fulfill some degrees of

reality, by mimicking the irregular half-length outputted from fracture modelling

software and naturally occurring patterns often found at cores. We applied this innovative

approach to actual shale gas and shale oil wells. We then found that by coupling

additional data into the AHM process, the fracture geometries and petrophysical

properties can be more accurately depicted. The obtained results are also highly

assimilating with the field experience from the engineers. In addition, by studying natural

fractures in the model, we found out that the connectivity between natural fractures and

wellbore/hydraulic fractures plays an important role in determining the well’s EUR

potential.

This study is beneficial because more reliable and robust results based on

geological/geomechanical information, along with non-deterministic realizations of

reservoir and fractures, can provide invaluable guidance towards well spacing planning,

EUR estimation and economic appraisal, and fracture design optimizations.

vii
Table of Contents
List of Tables ..................................................................................................................... xi

List of Figures ................................................................................................................... xii

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

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

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

1.3 Thesis Outline .......................................................................................................2

Chapter 2: Literature Review ...............................................................................................4

2.1 Fracture Modelling ...............................................................................................4

2.2 History Matching ..................................................................................................8

2.3 Corner Point Gridding ........................................................................................10

2.4 Natural Fractures.................................................................................................11

Chapter 3: Methodology ....................................................................................................12

3.1 Introduction .........................................................................................................12

3.2 Automatic History Matching (AHM) Workflow ................................................12

3.3 Input Module.......................................................................................................14

3.3.1 Field Inputs ..........................................................................................14

3.3.2 Data Treatments ...................................................................................14

3.3.3 Base Model Setup ................................................................................15

3.4 Iterative Optimization Module............................................................................16

3.4.1 Uncertainty Identifications...................................................................16

3.4.2 Smart Sampling....................................................................................18

3.4.3 EDFM with Hydraulic Fractures .........................................................19


viii
3.4.4 EDFM with Natural Fractures .............................................................22

3.4.5 Reservoir Simulator .............................................................................24

3.4.6 Artificial Intelligence (AI) ...................................................................25

3.4.7 Markov Chain Monte Carlo (MCMC) Algorithm ...............................28

3.5 Solution Module .................................................................................................30

3.5.1 Solution Filtering .................................................................................30

3.5.2 Representative Best Match ..................................................................31

3.5.3 Posterior Distribution ...........................................................................32

3.6 Post-processing Module ......................................................................................32

3.6.1 Probabilistic Production Forecast ........................................................32

3.6.2 Visualizations .......................................................................................33

Chapter 4: Application of Automatic History Matching to Corner-Point Gridding


System in Shale Reservoirs ..........................................................................................34

4.1 Introduction .........................................................................................................34

4.2 Reservoir Model .................................................................................................34

4.3 Hydraulic Fracture Generator in the Corner-Point Grid Model .........................41

4.4 Uncertain Reservoir and Fracture Parameters Identifications ............................47

4.5 Automatic History Matching ..............................................................................50

4.6 Probabilistic Production Forecasting ..................................................................73

4.7 Natural Fracture Sensitivity Study And Visualizations ......................................77

4.8 Conclusions.........................................................................................................86

Chapter 5: Development of Realistic Natural Fracture Module in AHM by Using


Fractal Theory ..............................................................................................................87

5.1 Introduction .........................................................................................................87

ix
5.2 Reservoir Model .................................................................................................87

5.3 Connectivity Analysis and Efficiency Optimization ..........................................91

5.4 Parameters Identifications...................................................................................93

5.5 History Matching ................................................................................................98

5.6 Probabilistic Production Forecasting ................................................................122

5.7 Matrix and Fracture Pressure Visualizations ....................................................125

5.8 Conclusions.......................................................................................................128

Chapter 6: Summary, Conclusions, and Recommendations for Future Work .................130

6.1 Summary and Conclusions ...............................................................................130

6.2 Recommendations for Future Work .................................................................131

Glossary ...........................................................................................................................133

Acronyms ................................................................................................................133

Nomenclature ..........................................................................................................134

References ........................................................................................................................137

x
List of Tables

Table 3.1: Summary of significant parameters that is supported by AHM workflow. ......17

Table 4.1: Summary of statistical information regarding natural fractures in this

study. .............................................................................................................35

Table 4.2: Summary of basic reservoir, fracture and gas desorption parameters used

in this study. ..................................................................................................36

Table 4.3: Summary of uncertain parameters, their ranges, fixed parameters, and

their deterministic values used for automatic history matching. ..................48

Table 5.1: Summary of basic reservoir, fracture and gas desorption parameters used

in this study. ..................................................................................................89

Table 5.2: Summary of uncertain parameters, their ranges, fixed parameters, and

their deterministic values used for automatic history matching. ..................95

xi
List of Figures

Figure 2.1: Illustration of the concept for Embedded Discrete Fracture Model: (a)

Physical domain; (b) Numerical domain. (Xu et al. 2017). ............................7

Figure 3.1: General illustration of automatic history matching workflow. .....................13

Figure 3.2: Graphical illustration of the permeability decay and fracture closure

concepts. (Glover et al. 2015) .......................................................................18

Figure 3.3: Schematic illustration of the Latin Hypercube sampling method. (Shams

2016) .............................................................................................................19

Figure 3.4: Summary of the workflow for inputting the hydraulic fractures in

arbitrary gridding system. .............................................................................21

Figure 3.5: Graphical illustration of the variables and definitions used for inputting

the hydraulic fractures in arbitrary gridding system. ....................................22

Figure 3.6: Graphical illustration of the fractal theory and the effects of different

fractal parameter on natural fracture network: (a) 2D fractal; (b) 3D

fractal. (Darcel et al. 2003b). The D is fractal dimension and a is length

dimension. .....................................................................................................24

Figure 3.7: Simplified schematic structure of XGBoost algorithm. ................................25


Figure 3.8: Simplified schematic structure of Markov Chain Monte Carlo algorithm. ..29

Figure 3.9: Illustration of deviation of original best match (left) model’s fracture

height to the P50 value (right) of the history matching solution

ensemble. ......................................................................................................31

Figure 4.1: Visualizations of basic reservoir and well information in the corner point

gridding system: (a) Top view; (b) 3D view; (c) Regional 3D well

landing location; (d) Global 3D well landing location. ................................38

xii
Figure 4.2: Statistical distributions of spatial configuration properties of natural

fractures characterized by geologist for the reservoir model. .......................39

Figure 4.3: Visualization of reservoir dip and depth variations for the bulk field

model (not upscaled). ....................................................................................39

Figure 4.4: Relative permeability curves for modeling gas and water flow for this

reservoir model. ............................................................................................40

Figure 4.5: Porosity field with heterogeneity in both horizontal directions and

vertical direction of the reservoir model used in this study. .........................40

Figure 4.6: Comparison between the conventional hydraulic fracture design and

simplified heel-biased hydraulic fracture design: (a) Uniform half-

lengths; (b) non-uniform half-lengths; (c) Non-uniform half-lengths with

natural fractures. ...........................................................................................42

Figure 4.7: Fracture modelling software outputs for the hydraulic fractures modelled

in this study: (a) Single stage; (b) Entire well...............................................43

Figure 4.8: Illustration of the problem of creating completely horizontal fractures

(light blue) from the perforation locations: (a) Overview; (b) Enlarged

view. ..............................................................................................................45

Figure 4.9: Visualization of tilting from the horizontally generated fractures (light

blue) to the fractures that aligns with the dip of the reservoir model

(green): (a) Overview; (b) Heel side view; (c) Toe side view. .....................46

Figure 4.10: Visualization of maximum and minimum permeability decline curve. ........49

Figure 4.11: Visualization of maximum and minimum half-lengths values for both

outer and inner fractures: (a) Minimum outer and inner fracture half-

lengths; (b) Maximum outer and inner fracture half-lengths. .......................50

Figure 4.12: Conversion from WHP to BHP used in this study........................................51


xiii
Figure 4.13: Locations of proxy prediction points for calculating values of response

variables’ objective functions during history matching iterations: (a)

Flowing BHP; (b) Water flow rate................................................................52

Figure 4.14: Comparison between simulation results from initial iteration (Latin

Hypercube sampling process) and actual production history: (a) Gas

flow rate; (b) Cumulative gas production; (c) Water flow rate; (d)

Cumulative water production; (e) Water gas ratio; (f) Flowing BHP. .........55

Figure 4.15: Number of history matching solutions versus iteration number. ..................57

Figure 4.16: Comparison between 55 filtered history matching solutions’ simulation

results and actual production history: (a) Gas flow rate; (b) Cumulative

gas production; (c) Water flow rate; (d) Cumulative water production;

(e) Water gas ratio; (f) Flowing BHP. ..........................................................60

Figure 4.17: Comparison between best match’s simulation results and actual

production history: (a) Gas flow rate; (b) Cumulative gas production; (c)

Water flow rate; (d) Cumulative water production; (e) Water gas ratio;

(f) Flowing BHP. ..........................................................................................63

Figure 4.18: Comparison between prior and posterior distributions of the studied

uncertain parameters based on 55 history matching solutions: (a) Matrix

permeability; (b) Fracture permeability decay factor; (c) Fracture water

saturation; (d) Fracture height; (e) Outer fracture half-length; (f) Inner

fracture half-length; (g) Hydraulic fracture conductivity; (h) Natural

fracture conductivity. ....................................................................................66

Figure 4.19: Characterized fracture geometry from AHM workflow: (a) Non-uniform

fracture half-lengths; (b) Fracture height. .....................................................67

xiv
Figure 4.20: Pressure-dependent fracture permeability decline curve of 55 history

matching solutions and the best match model. .............................................67

Figure 4.21: Parallel coordinate plot to visualize combinations of uncertain

parameters, including non-history matching solutions (light grey);

history matching solution (light orange); and best match solution (red). .....68

Figure 4.22: Comparison between response parameters’ values and relative errors of

proxy model and simulation results for iteration 2, 4, 6, 8 and 10: (a)

Water flow rate at 325 days; (b) Flowing BHP pressure at 334 days; (c)

Water flow rate relative error; (d) Flowing BHP relative error; (e) Global

relative error. .................................................................................................71

Figure 4.23: Scatter plots of values of different uncertain parameters and global error

versus simulation index: (a) Matrix permeability; (b) Fracture

permeability decay factor; (c) Fracture water saturation; (d) Fracture

height; (e) Outer fracture half-length; (f) Inner fracture half-length; (g)

Hydraulic fracture conductivity; (h) Natural fracture conductivity; (i)

global error. ...................................................................................................73

Figure 4.24: Production forecast of 55 history matching solutions: (a) flowing BHP;

(b) Gas flow rate; (c) Cumulative gas production; (d) Water flow rate;

(e) Cumulative water production. .................................................................76

Figure 4.25: Connectivity analysis of the natural fractures presented in the history

matched model: (a) Spatial distribution for different natural fracture

groups; (b) Total number of natural fractures in different groups. ...............79

Figure 4.26: Spatial connectivity analysis of the natural fractures magnified: (a) 1.5

times in both lengths and heights; (b) 1.8 times in both lengths and

heights. ..........................................................................................................80
xv
Figure 4.27: Visualizations of the effects of the natural fractures’ connectivity on

matrix pressure distributions after 20 years: (a) No natural fractures; (b)

Original natural fractures; (c) Natural fractures enlarged 1.5 times in

both length and height; (d) Natural fractures enlarged 1.8 times in both

length and height. ..........................................................................................81

Figure 4.28: Visualizations of the effects of the natural fractures’ connectivity on

drainage volumes after 20 years: (a) No natural fractures; (b) Original

natural fractures; (c) Natural fractures enlarged 1.5 times in both length

and height; (d) Natural fractures enlarged 1.8 times in both length and

height.............................................................................................................83

Figure 4.29: Visualizations of the effects of the natural fractures’ connectivity on

fracture pressure distributions after 20 years: (a) No natural fractures; (b)

Original natural fractures; (c) Natural fractures enlarged 1.5 times in

both length and height; (d) Natural fractures enlarged 1.8 times in both

length and height. ..........................................................................................85

Figure 4.30: Comparison of 20 year gas estimated ultimate recovery for four different

natural fracture connectivity scenarios. ........................................................85

Figure 5.1: A 3D example model visualizations for the reservoir, hydraulic fractures

(blue) and fractal natural fractures (green) in this study. ..............................89

Figure 5.2: Fixed relative permeability used for this study: (a) Matrix; (b) Hydraulic

fractures.........................................................................................................90

Figure 5.3: Connectivity check and non-connected natural fractures (pink) removal

for faster history matching. ...........................................................................91

xvi
Figure 5.4: Illustration of negligible effects from removing non-connected fractures

from the natural fracture network based on fractal theory: (a)

Bottomhole pressure; (b) Cumulative gas production; (c) Cumulative

water production. ..........................................................................................93

Figure 5.5: Sensitivity plots of varying fractal parameters vs natural fracture

distributions: (a) Df=2.6, Dl=2.45, α=4.5; (b) Df=2.8, Dl=2.45, α=4.5; (c)

Df=2.7, Dl=2.3, α=4.5; (d) Df=2.7, Dl=2.6, α=4.5; (e) Df=2.7, Dl=2.45,

α=4; (f) Df=2.7, Dl=2.45, α=5. .....................................................................98

Figure 5.6: Conversion from WHP to BHP used in this study........................................99

Figure 5.7: Locations of proxy prediction points for calculating values of response

variables’ objective functions during history matching iterations: (a)

Flowing BHP; (b) Water gas ratio. .............................................................100

Figure 5.8: Comparison between simulation results from initial iteration (Latin

Hypercube sampling process) and actual production history: (a) Gas

flow rate; (b) Cumulative gas production; (c) Water flow rate; (d)

Cumulative water production; (e) Water gas ratio; (f) Flowing BHP. .......103

Figure 5.9: Number of history matching solutions versus iteration number. ................105

Figure 5.10: Comparison between 52 filtered history matching solutions’ simulation

results and actual production history: (a) Gas flow rate; (b) Cumulative

gas production; (c) Water flow rate; (d) Cumulative water production;

(e) Water gas ratio; (f) Flowing BHP. ........................................................108

Figure 5.11: Comparison between best match’s simulation results and actual

production history: (a) Gas flow rate; (b) Cumulative gas production; (c)

Water flow rate; (d) Cumulative water production; (e) Water gas ratio;

(f) Flowing BHP. ........................................................................................111


xvii
Figure 5.12: Comparison between prior and posterior distributions of the studied

uncertain parameters based on 52 history matching solutions: (a) Matrix

permeability; (b) Fracture permeability decay factor; (c) Fracture water

saturation; (d) Matrix porosity; (e) Fracture height; (f) Outer fracture

half-length; (g) Inner fracture half-length; (h) Hydraulic fracture

conductivity; (i) 3D fractal dimension; (j) 3D length dimension; (k) 3D

alpha regulator. ...........................................................................................114

Figure 5.13: Characterized fracture geometry from AHM workflow: (a) Overview;

(b) Non-uniform fracture half-lengths; (c) Fracture height. .......................116

Figure 5.14: Pressure-dependent fracture permeability decline curve of 52 history

matching solutions and the best match model. ...........................................116

Figure 5.15: Parallel coordinate plot to visualize combinations of uncertain

parameters, including non-history matching solutions (light grey);

history matching solution (light orange); and best match solution (red). ...117

Figure 5.16: Comparison between response parameters’ values and relative errors of

proxy model and simulation results for iteration 2, 4, 6, 8 and 10: (a)

BHP at 30 days; (b) WGR at 20 days; (c) BHP at 745 days; (d) WGR at

743 days; (e) BHP at 1626 days; (f) WGR at 1623 days; (g) BHP

relative error; (h) WGR relative error; (i) Global relative error. ................119

xviii
Figure 5.17: Scatter plots of values of different uncertain parameters and global error

versus simulation index: (a) Matrix permeability; (b) Fracture

permeability decay factor; (c) Fracture water saturation; (d) Matrix

porosity; (e) Fracture height; (f) Outer fracture half-length; (g) Inner

fracture half-length; (h) Hydraulic fracture conductivity; (i) 3D fractal

dimension; (j) 3D length dimension; (k) 3D alpha regulator; (l) Global

error. ............................................................................................................122

Figure 5.18: Production forecasting of 52 history matching solutions: (a) Flowing

BHP; (b) Gas flow rate; (c) Cumulative gas production; (d) Water flow

rate; (e) Cumulative water production. .......................................................125

Figure 5.19: Pressure distribution of matrix at different production time: (a) 365 days;

(b) End of history (1626 days); (c) 10 years; (d) 20 years. .........................127

Figure 5.20: Pressure distribution of complex fracture network at different production

time: (a) 365 days; (b) End of history (1626 days); (c) 10 years; (d) 20

years. ...........................................................................................................128

xix
Chapter 1: Introduction

1.1 MOTIVATION

Shale rocks, which has nano-darcy level permeability, is conventionally thought

to be only the source rock that provides hydrocarbon or prevents hydrocarbons from

escaping. With the unlock of technology such as hydraulic fracturing and horizontal

drilling, hydrocarbons from the shale formations are engineered to be recoverable

nowadays. Nevertheless, significant challenges remain for shale reservoirs because of the

extensive number of uncertainties. In contrast with conventional reservoirs, shale

reservoirs are mainly feasible because of hydraulic fractures, and without proper

characterizations of these fractures, extracting economical values from these reservoirs

will be hard, and optimizations for future projects will be impossible.

With wide availabilities of production data and the continuous developments of

fracture modelling techniques, properties of the shale petrophysics and hydraulic fracture

geometry could be dynamically calibrated, by using numerical optimization procedure

called history matching. Due to current technology limits, it is not possible to fully

visualize the actual geometries of hydraulic fractures subsurface, making the history
matching problem ill-posed. Myriad combinations of the physical properties can match

the production data, and it is nonviable to validate which of these combinations is true. In

addition, some of the physical processes are still unknown to engineers, and others are

not fully understood yet. The advent of computer aided (or automatic) history matching

process provided some acceptable solutions to the problem, by quantitatively

characterizing the associated uncertainties.

1
From another aspects, the collaborations between geologist and engineers should

be even closer than before. The problem of how the work of field geologists can provide

insights to unconventional reservoir engineers is one of the keys to better solve the

problem of unconventional reservoir characterization. Two of the most important

knowledge would be the field-scale heterogeneity and natural fracture modelling, which

points to the objective of this study.

1.2 RESEARCH OBJECTIVES

The objective of this study is to utilize automatic history matching using a

Bayesian method and advanced machine learning algorithm to better study and

characterize unconventional reservoirs. Furthermore, more physics-based data, such as

geology model and fractal dominated natural fractures usually observed in outcrops, are

included in these numerical optimizations to explore their effects on the results of

automatic history matching.

1.3 THESIS OUTLINE

This thesis is structured mainly in 7 parts: In Chapter 2, an extensive literature

review on fracture modelling, history matching, corner point gridding of geology model,
and fractal theory behind natural fracture generation is provided.

In Chapter 3, we explain in detail the entire procedures/modules of the proposed

automatic history matching workflow, including input module; iterative optimization

module; solution module and post-processing module.

Chapter 4 introduces a shale gas field example to demonstrate how geology model

could be incorporated in the automatic history matching process. Probabilistic production

2
forecast, natural fracture sensitivity studies, and pressure visualizations are also

performed.

In Chapter 5, another shale gas well is history matched automatically to show how

fractal theory could assist to generate a more realistic natural fracture network. Similarly,

probabilistic production forecast and pressure visualizations are delivered.

In Chapter 6, all the studies and the corresponding findings are summarized, and

recommendations for future studies are described.

3
Chapter 2: Literature Review

In this chapter, we review and discuss the current literature of numerical

modelling and optimization of unconventional reservoirs, its challenges, present

developments, and potential solutions. Background in four main aspects, from fracture

modelling; history matching; complex gridding system and natural fractures, will be

thoroughly introduced.

2.1 FRACTURE MODELLING

With the advent of hydraulic fracturing and horizontal drilling, more hydrocarbon

resources can be extracted from previously challenging unconventional reservoirs. Even

though fracturing is vastly applied in nearly all unconventional reservoirs, the modelling

and numerical characterizations of the geometry of these man-made fractures is still

under developments, never minding the unfathomable myths about subsurface nature

factures. In the numerical optimization community, this modelling becomes more

significant because it greatly affects the quality of history matching results. Therefore, we

introduce some of the background behind fracture modelling below.

Numerical modelling of both hydraulic and natural fractures in the


unconventional reservoirs is difficult. McClure et al. (2020) summarized three major

topics in fracture modelling, namely categories of the model; workflow of the model; and

simplifications/assumptions of the model. They also outlined several challenging aspects

related to the fracture modelling, some of which are not well understood. These includes

spacing designs; factors dominating fracture growth; symmetry; proppant modelling;

RTA (rate transient analysis) analysis; net pressure modelling; initial shut-in pressure

(ISIP) and geomechanical heterogeneity (McClure et al. 2020). Let alone the complexity

4
of hydraulic fracture modeling, the occurrences of natural fractures in the shale reservoirs

presents more uncertainty (Cipolla et al. 2011).

In general, to solve the complicated issues presented by McClure et al. (2020),

there are several mainstream methods to numerically model fractures in the

unconventional reservoirs: dual permeability dual porosity (Al-Shaalan et al. 2003; Lu et

al. 2009; Yang et al. 2018), unstructured grid (Wang 2015; Bosma et al. 2017; Ding

2019), local grid refinement (Cipolla et al. 2009; Abdle Moneim et al. 2012; Xu et al.

2019), and embedded discrete fracture model (Xu et al. 2017; Yu et al. 2018b; Miao et al.

2018).

The first method, dual porosity dual permeability model (DPDK), has complex

mathematical constructs and is very confined to simplistic fractures configurations.

Despite its considerable computational efficiency advantage, DPDK model cannot study

the complex fracture network system such as natural fractures and 3D configurations. In

addition, according to Rubin (2010), DPDK model fails to accurately model the matrix-

fracture interaction in shale reservoirs, due to a substantial pressure drop in matrix.

Furthermore, Kuchuk and Biryukov (2014) also points out that DPDK model’s limitation

based on its assumption of uniformed spaced, regularly connected orthogonal fractures,

and Sun et al. (2016) points out that DPDK model and its’ variants couldn’t properly

handle faults and local anisotropies.

The second method, unstructured grid, is more accurate to model fractures

reservoirs. However, it comes with tremendous computational costs, as suggested by

Yang et al. (2017). Moreover, the application of unstructured grid in unconventional

reservoirs requires accurate description of permeability tensors, which can be challenging

5
to be computed (Hassanpour et al. 2008). Therefore, for complex geometry of fractures,

unstructured grid struggles with the gridding arrangements and poor efficiency.

The third method, local grid refinement (LGR), has enhanced accuracy for

modelling fractured reservoirs. Nevertheless, Cipolla et al. (2009) proposed that LGR can

suffer from poor computational efficiency. Extra refinements around well blocks are also

required to accurately model flow scheme (Ghosh 1998). Tijink and Cottier (2019)

pointed out that extensive refinements must be performed to avoid invalidating results

due to vertically partial LGR if complex geological feature is presented. This would

render impractical computational costs for unconventional reservoir simulation jobs in

corner point gridding model, which will be discussed later in this chapter.

The last method, embedded discrete fracture model (EDFM), has gained favorable

popularity in recent years. The idea of a generalized and 3D EDFM is developed by

several researchers (Moinfar et al. 2014; Cavalcante Filho et al. 2015; Shakiba and

Sepehrnoori 2015; Xu et al. 2017). Extensive benchmarking of EDFM method with LGR

method is completed by Xu et al. (2017), and it is found out that EDFM method could

achieve similar accuracy with LGR but could reduce computational time by 20 times for

some complex configurations of fractures. This balance of accuracy and superior

efficiency made it possible for properly model hydraulic or natural fractures in the history

matching process. Xu et al. (2017) constructed the EDFM method utilizing non-

neighboring connection (NNC) concepts, as illustrated in Figure 2.1. Additional grid

blocks are added to original reservoir model to account for modified transmissibility

between fracture media, matrix media, and well media. One main drawback of the EDFM

method is that it assumes linear pressure distributions in matrix blocks that contain

fracture media (Rao et al. 2019).

6
(a) Physical domain

(b) Numerical domain

Figure 2.1: Illustration of the concept for Embedded Discrete Fracture Model: (a)
Physical domain; (b) Numerical domain. (Xu et al. 2017).

In summary, various numerical methods have been proposed to solve the

challenging problems of characterization unconventional fractured reservoirs. However,

each method is not able to model hydraulic or natural fractures perfectly. Besides,

significant amount of uncertain physical phenomena still cannot be explained currently.

As the development of sensing technologies, such as distributed temperature/acoustic

sensing (DTS/DAS), advances, more clarity will be brought to the discipline of fracture

modelling. For now, the EDFM method stands out to be the best fracture modelling

method to be coupled in the numerical optimization.

7
2.2 HISTORY MATCHING

To properly characterize the approximate numerical values for the

petrophysical/fracture parameters in unconventional reservoirs and make estimation of

the estimated ultimate recovery (EUR), the numerical optimization method such as

history matching is required. In this section, we review some of the current history

matching techniques. In general, history matching is divided into two main categories:

manual and automatic (Rwechungura et al. 2011).

The manual history matching requires the engineer to manually adjust one

parameter at a time to potentially match the historical reservoir responses. Tavassoli et al.

(2004) pointed out that manual history matching is considered to be utterly time

inefficient due to its trial-and-error nature. In addition, Kabir et al. (2003) stressed that

manual history matching is highly subjective and depends exclusively on the engineer’s

field knowledge/experience. The manual history matching is also very difficult to keep

up with the pace of the updates of real-time data (Kabir et al. 2003). Moreover, due to the

ill-posed problem of history matching, it is often not enough to obtain single realization

of the reservoir properties. Results from manual history matching is deemed to be overly

deterministic and could not appraise the uncertainties quantitatively.

The computer aided history matching (or automatic history matching, AHM) is

becoming the industry trend currently. The idea of AHM is to use mathematical

optimization constructs to benchmark simulation trials and reduce fitting errors

automatically, through machine languages. Many history matching work with various

optimization algorithms have been developed and applied. Arroyo Negrete et al. (2018)

developed a simplified gradient-based method to perform a history matching on the well

test responses. However, this approach cannot handle larger number of uncertain

8
parameters, and the length of history data is too small to represent field production. Gao

et al. (2017) used Gauss Newton method to minimize objective function, a measure of the

history matching error, and obtained satisfactory results. Vazquez et al. (2015) applied

particle swarm algorithm in an oil reservoir, and Xie et al. (2015) implemented

evolutionary algorithm to history matching of a shale gas reservoir and generated an

ensemble of realizations. Nevertheless, the history matching methods discussed above all

fail to properly characterize the uncertainties associated with these reservoirs. In fact,

Goodwin (2015) proposed that these optimization-based algorithms for history matching

is statistically biased. Even though some of the algorithms could generate a solution

ensemble, they are not designed to analyze the probabilistic distributions of model

parameters.

With these challenges, the proxy-based Monte Carlo Markov Chain (MCMC)

algorithm is developed to properly handle the challenges of uncertainty characterizations.

The MCMC sampler is considered probabilistic because it is an enhanced version of

Monte Carlo simulation with memory. Large sample size can well approximate the shape

of posterior distributions. However, due to the requirements of tremendous sample size,

proxy models must be adopted to substitute the role of reservoir simulator and quantify

the objective functions for the generated samples. Wantawin et al. (2017) used a

polynomial proxy model for shale gas history matching. Yu et al. (2018a) and

Tripoppoom et al. (2018) experimented with k-nearest neighbor (KNN) model and

artificial neural networks (ANN) for history matching of shale gas wells. Among these

approaches, the one with ANN proxy model has the highest accuracy and computational

efficiency. However, one problem with the ANN method is that its accuracy is sub-

optimal if the training data size is small (Sejdinovic 2015), which is common for history

9
matching problems due to the computation costs associated with reservoir simulations.

An alternative method is by using an ensembled machine learning method called

XGBoost, which has great accuracy when the training dataset size is small. This method

includes several advanced features: hyperparameters auto-tuning, robust built-in

overfitting preventions, etc. (Malik et al. 2020).

2.3 CORNER POINT GRIDDING

In order to properly account for the geological structures and petrophysical

heterogeneities associated with unconventional reservoirs, a corner point gridding model

system must be adopted. In this section, we briefly review some of the work completed to

incorporate the corner point gridding model in the numerical simulation.

In corner point gridding system, combinations of coordinate lines and nodal depth

are given (Ponting 1989) to represent complex geological features such as faults, cross-

stratified beds, and sophisticated boundaries (Ding and Lemonnier 1995). The utilization

of geology model in corner point gridding system could thus greatly reduce the

uncertainties associated with some petrophysical parameters, such as porosity. Kumar et

al. (2013) employed the corner point geology model into the numerical simulation of oil

sand steam-assisted gravity drainage (SAGD) process, but the simulation time for the fine

gridded geology model is impractical. Therefore, upscaling of the geology model should

be performed to be suitable for history matching process, especially for the fractured

unconventional reservoirs (Fumagalli et al. 2016).

Given the limitation of traditional fracture modelling methods such as dual

continuum and unstructured grid, they are extremely challenging to be adapted into the

sophisticated corner point gridding system (Du et al. 2017). However, numerous studies

10
of implementing EDFM in the corner point gridding system are conducted (Panfili and

Cominelli 2014; Panfili et al. 2015; Shah et al. 2016). Xu and Sepehrnoori (2019) further

extended the EDFM to numerically handle complex geometries encountered in the corner

point model. Therefore, a robust performance of generalized EDFM method in the corner

point gridding system is achieved and the formulations laid a solid foundation to couple

geology with numerical simulations and optimizations.

2.4 NATURAL FRACTURES

Significant challenges exist to properly model the natural fractures in any

unconventional reservoir, due to different geology backgrounds and stress environments.

The effects of natural fractures on automatic history matching is not negligible, as natural

fractures both affect hydraulic fracture growth and also the probabilistic production

forecast from results of history matching. Wu and Olson (2016) extensively analyzed the

effects of natural fractures on propagation of hydraulic fractures. Yu et al. (2019)

demonstrated the effects of natural fractures on long term productions. However, the

characterization of natural fracture network is difficult. Random spatial generation of

natural fractures could be highly biased and scale dependent. Darcel et al. (2003a) stated

that the natural fractures observed in many outcrops exhibit fractal distributions of their

center locations and power law distributions in their lengths. This observation is

supported by several other authors (Okubo and Aki 1987; Ouillon and Sornette 1996;

Main et al. 1990; Odling 1997). Kim and Schechter (2007) also stressed that the

advantage of fractal theory is scale-independent, which would be more realistic and

unbiased.

11
Chapter 3: Methodology

3.1 INTRODUCTION

The objective of this chapter is to introduce the key steps in the automatic history

matching (AHM) workflow. The major components of the workflow contain Embedded

Discrete Fracture Model (EDFM), reservoir simulator, proxy model and Markov Chain

Monte Carlo sampling (MCMC). Flowcharts and schematic diagrams will be provided to

explain each step.

3.2 AUTOMATIC HISTORY MATCHING (AHM) WORKFLOW

The basic workflow of the automatic history matching is summarized in Figure

3.1. There are four major modules in this workflow: input module (light grey); iterative

optimization module (light yellow); solution module (light green); and post-processing

module (light blue). The input module checks data quality and assures that reservoir

model is correctly configured. It contains three components: field inputs; data treatments

and base model setup. The iterative optimization module generates hundreds of

realizations using artificial intelligence (AI) engine, advanced sampling methods, EDFM

and reservoir simulator. It contains components such as uncertainty identifications; smart

sampling; hydraulic fractures (HFs) and natural fractures (NFs) modelled by EDFM;

reservoir simulator; AI; and Markov Chain Monte Carlo (MCMC) algorithm. Solution

module finds satisfactory results of AHM and obtains characterized parameters. This

module contains: solution filtering; best match; and posterior distribution. Post-

processing module provides potential guidance to the engineers with pressure drainage

and estimated ultimate recovery information. It contains probabilistic production forecast

and visualizations. In the next few sections, each module is introduced in detail.

12
Figure 3.1: General illustration of automatic history matching workflow.
13
3.3 INPUT MODULE

In this section, the input module of the automatic history matching workflow is

briefly introduced. The function of this module is to receive data, inspect and treat the

data, and prepare a base case model that is error free and also computationally efficient so

that the no problems will be encountered when the next module (iterative optimization

module) is executed.

3.3.1 Field Inputs

For shale reservoirs, significant amount of engineering and monitoring data is

available. However, not all of the available data could be used in the AHM workflow.

The most important input data for the AHM workflow is the production data. Such data

mainly includes production monitoring dates, gas flow rate, water flow rate, wellhead

pressure (WHP), bottomhole pressure (BHP), etc. for shale gas reservoir and an

additional oil flow rate for shale oil reservoir. Some of the other data might also be

inputted in the AHM workflow, such as geological model with heterogeneities,

characterized natural fracture/faults network by geologists, microseismic events, etc.

3.3.2 Data Treatments

The bottomhole pressure available in the field inputs are usually based on the

computation from rate transient analysis (RTA), which could have significant errors

associated. Preferable methods to obtain accurate bottomhole data include bottomhole

pressure monitors, or transformation of surface pressure to bottomhole pressure using

wellbore model. However, these two methods are either unavailable or too costly in most

of the shale projects. Therefore, we proposed a simplistic approach to estimate

bottomhole pressure:

14
𝑊𝐺𝑅
𝐵𝐻𝑃 = 𝑊𝐻𝑃 + ∆𝑃 ∗ (3.1)
𝑊𝐺𝑅𝑚𝑎𝑥
where 𝐵𝐻𝑃 is bottomhole pressure, 𝑊𝐻𝑃 is wellhead pressure, ∆𝑃 is an empirical

constant pressure loss, 𝑊𝐺𝑅 is the water gas ratio, and 𝑊𝐺𝑅𝑚𝑎𝑥 is the maximum

water gas ratio observed over the history period.

In addition to the treatment of pressure data, it is also very important to inspect

the fluid flow rate data. It is acceptable to keep shut-in intervals if these intervals are

situated in the middle of the total history period. However, they should be removed if

they are situated near the end of the history period, because they increase the

computational costs for simulations and provides less useful information.

3.3.3 Base Model Setup

After we completed the input and preparation of relevant data, it is crucial to build

a simulation model base case. Arbitrary values could be assigned to fracture geometrical

properties, petrophysical properties, etc., some of which will later become uncertain

parameters. The purpose of this component is not to get a match with the production data,

but rather to make sure the reservoir simulation runs error free and the total

computational time is controlled within 10 to 30 minutes. By setting up the base model,

the iterative optimization module will run smoothly, and hundreds of or even thousands

of realizations could be achieved within reasonable amount of time.

Another significance of base model is to perform pre-checks for uncertainty

identifications. If the combination for the base model results in huge deviations of

simulation responses compared with actual production, we know that we either

underestimate or overestimated the parameters, and this prior knowledge would

somewhat guide the uncertainty identification process in the next module.

15
3.4 ITERATIVE OPTIMIZATION MODULE

In this section, each component of the iterative optimization module of the

automatic history matching workflow is explained in detail. The function of this module

is to selectively create massive number of realizations using artificial intelligence (AI)

engine, advanced sampling methods, EDFM and reservoir simulator. The principles and

algorithms of this module is described below.

3.4.1 Uncertainty Identifications

Based on the field experience and combinations of the base model, some of

reservoir/fracture/petrophysical parameters should be regarded as uncertain parameters

for automatic history matching process to characterize. The rest of the parameters could

be regarded as fixed/constant parameters, if they are not sensitive to dynamic production,

or information regarding these parameters are given/implied from other sources.

Parameters that the workflow current can assign to be either fixed/uncertain are listed in

Table 3.1. In this table, we also listed the sources of information that can provide some

suggestions regarding the sampling range if the corresponding parameter is set to be

uncertain. In general, numerical intervals of these ranges highly depends on the

unconventional reservoir type (gas or oil), depths intervals (deep, >3000 m or shallow,
<2000 m), geological structure (thin or thick net pay) and engineering designs (proppants,

pumping schedule, cluster/stage spacing). It is noteworthy to point out that for shale

reservoirs, the fracture closure phenomenon is very important to be modelled. This

physical phenomenon is simplified by incorporating a fracture permeability decay factor.

This variable reduces the equivalent fracture permeability if the matrix pressure is

lowered. The equation used to model this phenomenon is



𝑘 ′ = 𝑘𝑃𝑖 𝑒 −𝛽(𝑃𝑖 −𝑃 ) (3.2)
16
where 𝑘 ′ is the reduced permeability at a specific reservoir pressure, 𝑘𝑃𝑖 is the original

permeability at initial reservoir pressure, 𝛽 is the fracture permeability decay factor, 𝑃𝑖

is the initial reservoir pressure, and 𝑃′ is a specific reservoir pressure after production

begins. Figure 3.2 graphically illustrates the concepts of fracture closure and permeability

decay factor. At deeper depth, the fracture is expected to close more easily, and thus a

steeper fracture permeability decline curve would be appropriate.

General
Parameters Sources of information
classification
Matrix permeability DFIT/core data
Matrix properties Matrix porosity Well logging
Matrix water saturation Field experience
Fracture height DFIT/microseismic
Fracture half-length Microseismic
Fracture water saturation Empirical value
Fracture conductivity DFIT/experience
Fracture properties
Fracture aperture Usually fixed
Cluster efficiency Field experience
Fracture permeability decay
Depth interval
factor
Exponent of water relative
Empirical value
permeability curve
Endpoint of water relative
Relative Empirical value
permeability curve
permeability
Exponent of gas relative
properties Empirical value
permeability curve
Exponent of oil relative
Empirical value
permeability curve
Number of natural fractures Regional geology
Length of natural fractures Regional geology
Conductivity of natural
Natural fracture Regional geology
fractures
properties
3D fractal dimension Regional geology/Image Log
3D length dimension Regional geology/Image Log
3D alpha regulator Regional geology/Image Log

Table 3.1: Summary of significant parameters that is supported by AHM workflow.


17
Figure 3.2: Graphical illustration of the permeability decay and fracture closure
concepts. (Glover et al. 2015)

3.4.2 Smart Sampling

In order to obtain representative realizations from the iterative optimization

module and guarantee accurate results from solution module, we must perform an initial

sampling process. In addition, this sampling process must be smart enough to fully

consider the uncertainty space constructed by the possible combinations of the specified

uncertain parameters.

Shams (2016) has already discussed different types of sampling methods for the
initial sampling process. However, in our proposed workflow, the highly efficient Latin

Hypercube method is applied. The reason for this choice is because other methods, such

as factorial designs/space-based designs/sequence-based designs, require huge number of

samples to be representative, but Latin Hypercube (LH) method only requires 2k number

of samples to be indicative of the uncertain space, where k is the total number of

uncertain parameters. Actually, a greater number of uncertain parameters require a

greater number of samples, but 50 initial samples are deemed to be enough, based on the
18
balance between precision and time cost to run the reservoir simulation of these initial

samples.

The principle of the LH method is summarized next. First, the range for each of

the specified uncertain parameters is divided into N sub-intervals. Then, random samples

are drawn and only one sample is allowed to be drawn from each high dimension sub-

space. These samples are drawn in a way that no repeated samples are enclosed in the

same sub-interval for any uncertain parameters. Figure 3.3 schematically illustrates this

method. The black grids represent selected samples and notice two samples that share the

same column or row with each other is non-existent. The advantage of this method is that

a minimal number of generated samples can have a wide distribution across the entire

uncertainty space, so that the samples are not biased, or too stochastic.

Figure 3.3: Schematic illustration of the Latin Hypercube sampling method. (Shams
2016)

3.4.3 EDFM with Hydraulic Fractures

Based on the results from smart sampling, we need to input hydraulic fractures in

the model first. The main procedures for inputting hydraulic fractures are summarized in

Figure 3.4. First, we need to read in the well trajectory information and perforation

19
report. Then, according to geology-related information and also the sampling results of

the uncertain parameters, dip/strike/tilting angles are inputted as fixed variables while

half-length, height and heel bias information is given by one single realization of AHM

workflow. Next, the coordinates of all the perforation points are linearly interpolated

based on the X, Y, and Z, measured depth (MD) values of two adjacent well trajectory

points, and the equation is given as

ℎ 𝑡 ℎ)
𝑥𝑝 = 𝑥𝑤 + (𝑥𝑤 − 𝑥𝑤 ∗𝑟
ℎ 𝑡 ℎ
𝑀𝐷𝑝 − 𝑀𝐷ℎ (3.3)
{ 𝑦𝑝 = 𝑦𝑤 + (𝑦𝑤 − 𝑦𝑤 ) ∗ 𝑟 , 𝑟=
ℎ 𝑡 ℎ)
𝑀𝐷𝑡 − 𝑀𝐷ℎ
𝑧𝑝 = 𝑧𝑤 + (𝑧𝑤 − 𝑧𝑤 ∗𝑟

where 𝑥𝑝 , 𝑦𝑝 , and 𝑧𝑝 are the X, Y, and Z coordinates of any perforation point,



𝑥𝑤 , 𝑦𝑤ℎ , and 𝑧𝑤

are the X, Y, and Z coordinates of the adjacent well trajectory point
𝑡
toward heel side, 𝑥𝑤 , 𝑦𝑤𝑡 , and 𝑧𝑤
𝑡
are the X, Y, and Z coordinates of the adjacent well
trajectory point toward toe side, 𝑟 is the interpolation ratio, and 𝑀𝐷𝑝 , 𝑀𝐷ℎ , and 𝑀𝐷𝑡

are the measured depths of perforation point, adjacent well trajectory point toward heel

side, adjacent well trajectory point toward toe side, respectively.

Next, the coordinates of the two endpoints of the fracture’s middle axis are

calculated based on the interpolated perforation locations, using following equations:

𝑥𝑚𝑎 = 𝑥𝑝 ± 𝑥𝑓 ∗ cos(𝑡𝑖𝑙𝑡𝑖𝑛𝑔) ∗ cos(𝑠𝑡𝑟𝑖𝑘𝑒)


(3.4)
{ 𝑦𝑚𝑎 = 𝑦𝑝 ± 𝑥𝑓 ∗ cos(𝑡𝑖𝑙𝑡𝑖𝑛𝑔) ∗ sin(𝑠𝑡𝑟𝑖𝑘𝑒)
𝑧𝑚𝑎 = 𝑧𝑝 ± 𝑥𝑓 ∗ sin(𝑡𝑖𝑙𝑡𝑖𝑛𝑔)

where 𝑥𝑚𝑎 , 𝑦𝑚𝑎 , and 𝑧𝑚𝑎 are the coordinates of the two endpoints of the fracture’s
middle axis, 𝑥𝑓 is the half-length of the bi-wing planar hydraulic fracture, tilting is the

angle between x-y plane and the middle axis (refer to Figure 3.5b), and strike is the angle

between positive x-axis and middle axis (refer to Figure 3.5d).

20
Then, the coordinates of the four vertices of the planar hydraulic fracture is

computed based on the calculated 𝑥𝑚𝑎 , 𝑦𝑚𝑎 , and 𝑧𝑚𝑎 and geometrical projections of the

vertices pair on the same side of the wing on x-z and x-y planes, as shown in Figure 3.5c

and Figure 3.5d:

𝑥𝑣 = 𝑥𝑚𝑎 ± 0.5ℎ𝑓 ∗ cos(𝑑𝑖𝑝) ∗ cos(𝑠𝑡𝑟𝑖𝑘𝑒) + |0.5ℎ𝑓 ∗ cos(90° − 𝑡𝑖𝑙𝑡)|


(3.5)
∗ sin(𝑑𝑖𝑝) ∗ sin(𝑠𝑡𝑟𝑖𝑘𝑒)
𝑦𝑣 = 𝑦𝑚𝑎 ± 0.5ℎ𝑓 ∗ cos(𝑑𝑖𝑝) ∗ sin(𝑠𝑡𝑟𝑖𝑘𝑒) − |0.5ℎ𝑓 ∗ cos(90° − 𝑡𝑖𝑙𝑡)|
∗ sin(𝑑𝑖𝑝) ∗ cos(𝑠𝑡𝑟𝑖𝑘𝑒)
{ 𝑧𝑣 = 𝑧𝑚𝑎 ± |0.5ℎ𝑓 ∗ cos(𝑡𝑖𝑙𝑡)| ∗ sin(𝑑𝑖𝑝)

where 𝑥𝑣 , 𝑦𝑣 , and 𝑧𝑣 are the coordinates of the four vertices of the hydraulic fracture,
ℎ𝑓 is the hydraulic fracture height, dip is the angle between hydraulic fracture plane and

x-y plane. Then, these coordinates are outputted and also natural fractures’ coordinates if

they are available.

Figure 3.4: Summary of the workflow for inputting the hydraulic fractures in arbitrary
gridding system.
21
(a) 3D view (b) y-z plane view

(c) z-x plane view (d) x-y plane view

Figure 3.5: Graphical illustration of the variables and definitions used for inputting the
hydraulic fractures in arbitrary gridding system.

3.4.4 EDFM with Natural Fractures

Once the hydraulic fractures are successfully inputted into the reservoir model, we

can then add the background natural fractures in the model. There are generally three

ways to include natural fractures in our workflow. The first one involves a fixed or

characterized natural fracture network. This could come from a random distribution, or
22
from the field geologist working on the project. This method is deterministic and cannot

consider the associated uncertainty with natural fractures. The second method generate

random realizations of total number, length, and conductivity of the natural fractures

within the coordinate ranges of the model, with single/multiple sets of fixed dip and

strike angles. This method is probabilistic to properly consider the uncertainty associated

with natural fractures, but it is deemed to be too stochastic with high randomness. The

last method involves the use of 3D fractal theory. Darcel et al. (2003a) proposed that

natural fracture networks in nature, as observed in various outcrops, follow two unique

geometrical patterns: fractal distributions of fracture centers and power law distributions

of lengths. Hence, this method is beneficial because it is more realistic than the second

method, which lacks physical basis. The equation for fractal controlled natural fracture

density is (Kim & Schechter 2007):

3+𝐷
Γ ( 2 𝑙) (3.6)
𝑛(𝑙, 𝐿)𝑑𝑙 = 𝛼3𝐷 𝐿𝐷𝑐3𝐷 𝑙 −(𝐷𝑙3𝐷+1) , 𝛼3𝐷 = 𝛼√𝜋
2+𝐷
Γ ( 2 𝑙)

where 𝑛(𝑙, 𝐿) is the density of natural fractures given a minimum fracture length of 𝑙

and a model scale of 𝐿, 𝛼3𝐷 is the 3D alpha regulator are derived from 2D fractal

parameters, 𝐷𝑐3𝐷 is the 3D fractal dimension number, 𝐷𝑙3𝐷 is the 3D length dimension

number, 𝛼 is the 2D alpha regulator that controls natural fracture density, Γ is the

gamma function, and 𝐷𝑙 is the 2D length dimension number. The power law distribution

of the natural fracture can then be back-interpolated by this equation. Figure 3.6

schematically shows the effects of fractal and length dimensions on the spatial

distributions of natural fracture network in 2D and 3D.

23
(a) 2D fractal (b) 3D fractal

Figure 3.6: Graphical illustration of the fractal theory and the effects of different fractal
parameter on natural fracture network: (a) 2D fractal; (b) 3D fractal. (Darcel
et al. 2003b). The D is fractal dimension and a is length dimension.

3.4.5 Reservoir Simulator

After the preprocessing of both hydraulic and natural fractures in the reservoir

model using EDFM preprocessor, the reservoir simulations will be performed. The non-

intrusiveness of the EDFM preprocessor allows us to execute the simulations using any

of the third-party simulation software, including CMG, ECLIPSE, NEXUS, and etc. In

our proposed workflow, the black oil simulator – CMG’s IMEX is utilized. We did not

select the CMG’s compositional simulator – GEM because of its high computational

costs. One key aspect that needs to be considered is the Langmuir isotherm gas

desorption and absorption phenomenon that is unique for shale gas simulation. The

equation for this phenomenon is

𝐾 ∙ 𝑃𝑚
𝛼𝑎𝑑𝑠 = 𝜌(1 − 𝜙)𝑉̅𝑚𝑎𝑥 ( ) (3.7)
1 + 𝐾 ∙ 𝑃𝑚
24
where 𝛼𝑎𝑑𝑠 is the unit fractional adsorption, 𝜌 is the rock density, 𝜙 is the matrix

porosity, 𝑉̅𝑚𝑎𝑥 is the maximum volume of absorbed gas per unit mass of rock, 𝐾 is the

inverse Langmuir pressure, and 𝑃𝑚 is the matrix pressure.

3.4.6 Artificial Intelligence (AI)

Figure 3.7: Simplified schematic structure of XGBoost algorithm.

After we finished the EDFM preprocessing and reservoir simulations, we need to

establish the artificial intelligence engine using proxy models. The purpose of proxy

model is to establish a relationship between simulation input and output, so that it can

mimic a reservoir simulator and avoid actual simulation runs. There are many machine

learning algorithms that construct regression model on non-linear problems, including K-

nearest neighbor (KNN), deep neural network (DNN), polynomial, kriging, etc. However,

in our proposed workflow, the proxy model called XGBoost is adopted.


25
XGBoost is actually an advanced agglomeration of several machine learning

algorithms and methods. It has a basis in gradient boosting, which is an ensemble of

machine learning models that optimize the objective function collaboratively. The

ensemble learning models embedded in XGBoost are the regression decision trees, which

is predicts the outcome based on several inputs using the tree algorithms. A simplified

schematic structure for the XGBoost is provided in Figure 3.7. One unique feature about

XGBoost is the objective function is additive and customized, so that the loss function

could be very efficiently optimized. This is represented by (Bentéjac et al. 2019):


𝑁 𝑀

𝐿𝑋𝐺𝐵 = ∑ 𝐿(𝑦𝑖 , 𝐹(𝑥𝑖 )) + ∑ Ω(ℎ𝑚 ) (3.8)


𝑖=1 𝑖=1

where 𝐿𝑋𝐺𝐵 is the overall loss function, 𝑁 is the total training size, 𝐿(𝑦𝑖 , 𝐹(𝑥𝑖 )) is the

loss function value at data point i given input data of xi and output data of yi, 𝑀 is the

total number of sub-loss function, and Ω(ℎ𝑚 ) is the regularization function, which is

closely related to the tree structures.

As outlined by Malik et al. (2020), there are several advantages of XGBoost. The

model has a build-in cross validation methods, which prevents excessive overfitting

issues common among machine learning models. It also has some embedded search
algorithm to partially aid the selection of hyperparameters. The code structure is

constructed using parallelization and widely supported by different programming

languages. XGBoost also has a spare-awareness algorithm, which properly handles

missing data. In general, the implementation of neural network is optimal when the size

of the available data is massive. For the ill-posed history matching problem, the data size

is limited because of the limited computational time we assign to the task. Therefore, the

26
XGBoost algorithm would relatively converge faster and have better accuracy than neural

network over smaller amount of data (Sejdinovic 2015).

The exact method we utilized the proxy model is to use the model to predict the

values of selected representative time locations along the history data. The input of the

proxy model would be combinations of the uncertain parameters’ combinations, and the

output of the proxy model would be the value of different response variable (water flow

rate, bottomhole pressure, etc.) at a single time location. Usually 10 proxy locations are

selected, so for each of the response variable, 10 proxy models are built. For shale gas

reservoir, we generally select BHP and water flow rate (or water gas ratio) to be the

response variables to be predicted by proxy models, while for shale oil reservoir, an

additional variable of oil gas rate is incorporated. To measure the error of the multi-proxy

models or actual simulation run for each response variable, we calculate the objective

function as

𝑦𝑖,𝑚𝑜𝑑𝑒𝑙 − 𝑦𝑖,ℎ𝑖𝑠𝑡
∑𝑝𝑖=1 | × 100| × 𝑤𝑖 (3.9)
𝑦𝑖,ℎ𝑖𝑠𝑡
𝐹𝑗 =
∑𝑝𝑖=1 𝑤𝑖

where 𝐹𝑗 is the objective function value for response variable j, p is the total proxy

locations selected for the response variable j, i is the index of the proxy location, 𝑦𝑖,𝑚𝑜𝑑𝑒𝑙

is the value of the response variable j at time i from either simulation run or proxy model,
𝑦𝑖,ℎ𝑖𝑠𝑡 is the value of the response variable j at time i from actual history data, and wi is

the weight assigned to the proxy point i, given a response variable j. Similarly, the global

objective function is calculated in a similar way (just sum up all response variables). The

global objective function value will be used in the next component (MCMC), and the

individual objective function value will be used in the solution module.

27
3.4.7 Markov Chain Monte Carlo (MCMC) Algorithm

Figure 3.8 highly summarizes the entire procedure of the MCMC algorithm used

in the iterative optimization module. First of all, we start at the step of initial 𝜃, which is

the combination of uncertain parameters from the smart sampling component with lowest

global objective function value computed by Equation 3.9.

Next, we proceed into an iterative feature that repeats one hundred thousand steps

and a check is performed to see if the iterations reaches the maximum steps. If not, we

proceed to disturb the initial combination 𝜃 to a proposed combination 𝜃 ∗ with a small

change called “walk distance” as 𝛿. Next a step acceptance ratio, a*, is calculated as

(𝜀 ∗ )2 −𝜀 2

𝑎 = min {1, 𝑒 2𝜎2 } (3.10)

where 𝜀 ∗ is the global objective function of the proposed, disturbed combination of

uncertain parameters, 𝜀 is the global objective function of the previous (in this case

initial) combination of uncertain parameters, 𝜎 2 is the variance of the global objective

function values for the smart sampling results. The multi-proxy model is embedded in

this step to predict the values of response variables for all of the locations designated

previously and thus help us obtain the value for 𝜀 ∗ .


Then, the step acceptance ratio is checked with a random number between 0 and

1. If the ratio is greater than this number, the proposed combination is accepted as the

new sample. Otherwise, the old combination is regarded as the new sample. The

procedure then repeats, and a new walk distance is assigned, until maximum steps is

reached.

After the maximum step is reached, we check with the overall acceptance ratio

and see if it is between the empirical range of 15% to 50%. If this criterion is not

28
satisfied, we redo the previous steps by adjusting the variance value used in Equation

3.10 using an adjusting ratio, which usually falls between 2 to 10. Otherwise, we can

proceed to the next step and remove the initial 20% of the generated samples, which is

called “burn-in” period. This is done because the initial samples have a larger bias and is

not representative of the true posterior distribution. The remaining 80% of the samples

are thus the outputs from the MCMC algorithm. However, we only filter out 25 samples

with minimum global objective function values to be validated with the reservoir

simulator to optimize the accuracy.

Figure 3.8: Simplified schematic structure of Markov Chain Monte Carlo algorithm.

This brief explanation about MCMC concludes the principles for the iterative

optimization. In general, the quality of the multi-proxy model built from the smart

29
sampling process is poor. However, as MCMC procedures continue and actual validation

cases accumulate, the multi-proxy model will gradually improve qualitatively. Based on

our history matching experience, a maximum iteration number of 10 for this module is

sufficient to capture representative realizations of the reservoir, if the number of total

uncertain parameters is around 10 or below. However, as much as 20 iteration steps could

be run to guarantee superior accuracy, most robust characterizations of uncertainties, and

if the uncertain space is more complex (more than 10 uncertain parameters).

3.5 SOLUTION MODULE

In this section, we discuss each component of the solution module of the

automatic history matching workflow. The function of this module is to filter out optimal

ensemble of history matching solutions, obtain the most representative match with the

lowest global error, and obtain the final posterior distributions for various uncertain

parameters. Concise descriptions of each component are given below.

3.5.1 Solution Filtering

Once the iterative optimization module is finished, hundreds of actual simulation

results are available. However, some of results will still have high values of global
objective function, and it would be much less flexible if we use global function values to

treat all the simulations. Therefore, we need to selectively filter the history matching

results by using different threshold for each individual objective function. Simulation

cases whose values of individual objective functions satisfy all of the thresholds will be

grouped as the final history matching solution ensembles. Based on experience, the

threshold for bottomhole pressure’s objective function is much stricter than the one for

water flow rate/ water gas ratio, because significant data quality issues are associated
30
with water production due to the water flow back phenomenon. Compared to overly

deterministic traditional history matching methods, this ensemble of solutions greatly

considers the uncertainties associated with the ill-posed history matching problem.

3.5.2 Representative Best Match

There is a one optimal history matching solution with the lowest global objective

function value from the ensemble of the history matching solutions. However, one

problem exists with this best match model. For some of the uncertain parameters, it is

possible that this optimal model is not lying close to the statistical P50 values of the

solution ensembles, as shown in Figure 3.9. This presents a statistical bias, since the best

match model cannot catch the variability associated with the uncertainty space, as P50

value is defined to be the most likely value in any arbitrary distributions. Therefore, in

our workflow, the best match model is generated around the P50 values of the filtered

solution ensemble, to fully be as statistically representative as possible.

Figure 3.9: Illustration of deviation of original best match (left) model’s fracture height
to the P50 value (right) of the history matching solution ensemble.

31
3.5.3 Posterior Distribution

Once the solution ensembled is filtered and the representative best match model is

obtained, the visualization for posterior distributions can be achieved, as shown in the

right plot of Figure 3.9. Given the specific uncertain range, the histograms of the

distributions of the solution ensemble is plotted as the orange bars, and the representative

best match is shown as the red vertical line. The prior distribution, which is assumed to

be uniform in our workflow, is also plotted and shown as grey bars. Lastly, we included

the cumulative probability function based on the histograms as dashed curve to help us

judge the shape of the posterior distributions.

3.6 POST-PROCESSING MODULE

After history matching part is finished, we can proceed to conduct the

probabilistic analysis of ultimate estimated recovery (EUR) for the filtered solutions and

also visualize the pressure drawdown in the representative best match model, for both

matrix and fractures. This module’s function is intended for the work described above.

3.6.1 Probabilistic Production Forecast

After the history matching period and depending on the length of history, we need
to assign BHP drawdown schedule for the production forecast. If the length of history is

very short (around or less than 3 months), we assign a gradual drawdown from the last

BHP value of the history period to a minimum BHP constraint (usually between 1 to 2

Mpa) over a period of time, which depends on the decline rate of BHP in the history

period. We could also assign a direct BHP drop to the minimum constraint if the length

of history is long. Then, the production will continue until a total simulation time of 20

years is achieved. The EUR for both gas phase and water phase will be computed from

32
the reservoir simulator. The process is probabilistic because the work is repeated for the

entire solution ensemble and the representative best match, instead of a deterministic

approach for single model EUR usually observed in conventional methods. This

component provides important appraisal of the productivity of the studied shale reservoir.

3.6.2 Visualizations

The last step in our workflow would be the visualizations of pressure drawdown

in the both the matrix and fracture media. For both structural gridding system and corner

point gridding system, our workflow is able to extract the pressure drawdown

information. The ability of various visualizations enables us to dynamically capture the

connectivity of natural fractures, study the drainage extents of the hydraulic fractures.

These provides invaluable guidance towards well spacing optimization and future

refracturing job designs.

33
Chapter 4: Application of Automatic History Matching to Corner-Point
Gridding System in Shale Reservoirs

4.1 INTRODUCTION

The objective of this chapter is to demonstrate how the automatic history

matching workflow can be applied to a field-scale, single well corner-point gridding

model from an actual shale gas field. The chapter consists of sequential explanations of

steps necessary to complete the study, including basic reservoir model description,

fracture configuration preprocessing, uncertain/constant parameters identifications,

history matching and production forecast. We utilized the automatic history matching

module, which is composed of Markov chain Monte Carlo algorithm coupled with

XGBOOST proxy model, along with Embedded Discrete Fracture Model (EDFM) and

the corner-point gridding extension module to perform the study. The workflow is

applied to an actual shale gas well in Sichuan Basin in China with 334 days of production

history. The gas flow rate and water flow rate information are available, while the

bottomhole pressure information is derived from a special relationship between wellhead

pressure (WHP) and bottomhole pressure.

4.2 RESERVOIR MODEL

This field case contains a single well with a horizontal section length of around

2000 m. Since the model is in corner point gridding system, the well is deviated and is

observed to pass through several geological layers. We assumed water saturation and

permeability to be homogeneous, while porosity is heterogeneous as characterized by

field geologists. The well was completed with 27 stages of hydraulic fractures with a total

cluster number of 81. Each stage has 3 hydraulic fractures.

34
For this study, we used a static model in corner point gridding system, and there

are 1139 natural fractures (only one characteristic set) inscribed in this model as

described by the geologists. Both a view of the reservoir model and 3D visualization of

well landing location is given in Figure 4.1. Statistical distributions regarding the spatial

configurations of these natural fractures are given in Figure 4.2. We also tabulated Table

4.1 to summarize the spatial information obtained from the geologists. Average length

and height of natural fractures are 66 m and 32 m, respectively, and the mean dip angle is

about 22 degrees. These information imply that the natural fractures have poor

connectivity with both the wellbore and hydraulic fractures and is not well communicated

between vertical layers. This will be discussed in detail in section 4.7. The model

dimension is approximately 1000 m long × 2200 m wide and the thickness varies across

locations and across layers. The shallowest point of the model is around 3194 m and the

deepest point is about 3489 m, as shown in Figure 4.3. The model in z-direction is

upscaled into 10 layers, with average thickness of 30 m, five layers of 6 m, 6.2 m, 4.5 m,

1.53 m, and 9.92 m. The number of grid blocks is 64 and 144 in x and y direction. Basic

reservoir, fractures, and Langmuir isotherm gas desorption properties are summarized in

Table 4.2. The relative permeability curves (Figure 4.4) for modeling gas and water flow

are assumed to be fixed because they affect the dynamic production insensitively. The

heterogeneous porosity field can be visualized in Figure. 4.5.

Statistical Fracture Fracture Azimuth Dip Surface


parameters Length (m) Height (m) (°) (°) area (m2)
Mean 66.2 32.3 268.3 22.1 2248.9
Standard deviation 12.8 7.2 35.8 8.5 918.7
Minimum 47.0 1.5 3.2 1.4 231.6
Maximum 99.5 49.8 358.6 52.3 4951.0

Table 4.1: Summary of statistical information regarding natural fractures in this study.
35
Model parameters Value Unit
Model dimension (x × y) 1000 × 2200 m
Number of grid blocks (x × y × z) 64 × 144 × 10 -
Initial reservoir pressure 70 MPa
Reservoir temperature 150 ℃
Residual water saturation 30% -
Total compressibility 4.35×10-7 kPa-1
Reservoir depth 3194 ~ 3489 m
Well length ~ 2000 m
Number of stages 27 -
Clusters per stage 3 -
Average stage spacing 21 m
Average cluster spacing 17.3 m
Langmuir pressure 7 MPa
Langmuir adsorption volume 3.5 m3/ton
Rock density 2.54 g/cm3

Table 4.2: Summary of basic reservoir, fracture and gas desorption parameters used in
this study.

(a) Top view of deviated well (red) and natural fractures (light blue)

Figure 4.1: continued next page.


36
(b) 3D view of deviated well (red) and natural fractures (light blue)

(c) Regional 3D well landing location

Figure 4.1: continued next page.

37
(d) Global 3D well landing location

Figure 4.1: Visualizations of basic reservoir and well information in the corner point
gridding system: (a) Top view; (b) 3D view; (c) Regional 3D well landing
location; (d) Global 3D well landing location.

(a) Azimuth angle (b) Dip angle

Figure 4.2: continued next page.

38
(c) Length (d) Surface area

Figure 4.2: Statistical distributions of spatial configuration properties of natural


fractures characterized by geologist for the reservoir model.

Figure 4.3: Visualization of reservoir dip and depth variations for the bulk field model
(not upscaled).

39
Figure 4.4: Relative permeability curves for modeling gas and water flow for this
reservoir model.

Figure 4.5: Porosity field with heterogeneity in both horizontal directions and vertical
direction of the reservoir model used in this study.

40
4.3 HYDRAULIC FRACTURE GENERATOR IN THE CORNER-POINT GRID MODEL

A non-uniform fracture half-length per stage is assumed in our model. For

example, the two hydraulic fractures closest to subsequent, previous stage are called outer

fractures. The one middle hydraulic fracture is called inner fractures. Usually, outer

fractures have larger half-length than the inner fracture(s) (Figure 4.6b). This design is

significantly more beneficial compared to the traditional design in conventional history

matching work, in which the fractures are assumed to have uniform half-length (Figure

4.6a). This is because the characterized non-uniform half-lengths both honors the fracture

growth modelled by fracture modelling software and provides useful guidance regarding

a more realistic drainage area and thus better understanding of optimal spacing. To

support this statement, we modelled fractures of this well on both one stage and the entire

well using fracture modelling software, as shown in Figure 4.7. We inputted maximum

and minimum horizontal stress of 98 Mpa and 86 Mpa, with a Young’s modulus of 46

GPa and a Poisson’s ratio of 0.24, which are representative of the geomechanical

information for this reservoir. The single stage model provides us with non-uniform

hydraulic fracture half-lengths of 117 m, 73 m, and 99 m. In fact, the proppant

distributions in the stage designed hydraulic fractures tends to be heel-biased, which is

substantiated by the data provided by Wheaton et al. (2017). This implies that the non-

uniformity of the hydraulic fractures’ half-length could be more complex. However,

significant degrees of uncertainties could be introduced to our study if we adopted more

complex half-length geometries, so we assumed only two representative half-lengths (for

outer/inner fractures).

41
(a) Uniform half-lengths

(b) Non-uniform half-lengths

(c) Non-uniform half-lengths with natural fractures

Figure 4.6: Comparison between the conventional hydraulic fracture design and
simplified heel-biased hydraulic fracture design: (a) Uniform half-lengths;
(b) non-uniform half-lengths; (c) Non-uniform half-lengths with natural
fractures.

42
(a) Fracture modelling software output for a single stage

(b) Fracture modelling software output for the entire well

Figure 4.7: Fracture modelling software outputs for the hydraulic fractures modelled in
this study: (a) Single stage; (b) Entire well.

43
Since the actual perforation report for this well is available, the corner point

gridding fracture generation module can be applied directly. Nevertheless, special

attentions must be paid regarding the dip of the reservoir. The region of the reservoir

model near the toe side of the well has a lower dip angle compared with the region near

the heel side. If hydraulic fractures are generated perfectly horizontal to the well bore,

they will grow outside the model and thus affects the results of the simulation, as shown

in Figure 4.8. Therefore, we assigned tilting angles to be approximately 5 degrees near

for the cluster at toe side and linearly increased the tilting angle until the first cluster of

14th stage, which has a tilting angle of 8 degrees. The clusters in the rest of the stages

toward heel side all have tilting angles of 8 degrees. This tilting scheme can be visualized

in Figure 4.9.

(a) Overview

Figure 4.8: continued next page.

44
(b) Enlarged view

Figure 4.8: Illustration of the problem of creating completely horizontal fractures (light
blue) from the perforation locations: (a) Overview; (b) Enlarged view.

(a) Overview

Figure 4.9: continued next page.

45
(b) Heel side view

(c) Toe side view

Figure 4.9: Visualization of tilting from the horizontally generated fractures (light blue)
to the fractures that aligns with the dip of the reservoir model (green): (a)
Overview; (b) Heel side view; (c) Toe side view.

46
4.4 UNCERTAIN RESERVOIR AND FRACTURE PARAMETERS IDENTIFICATIONS

We identified eight important parameters to be uncertain parameters to perform

automatic history matching, including matrix permeability, fracture permeability decay

factor, fracture water saturation, fracture height, outer fracture half-length, inner fracture

half-length, fracture conductivity and natural fracture conductivity. These parameters are

deemed to be significant from aspect of engineering and reservoir descriptions. For

example, since the depth of this reservoir is more than 3000 meters, the stress tends to be

high and therefore fractures are more prone to be closed. Thus, we used pressure

dependent permeability to model this effect. The equation for pressure dependent

permeability is given in Equation 3.2. In theory, as many as infinite number of uncertain

parameters can be selected, but it would best to control the total number to be below 10 to

balance proxy efficiency and accuracy and to control non-linear degree of uncertainty

space. The porosity is already defined by the corner-point model, so it is a fixed

parameter. We also fixed the relative permeability curve (shown in Figure 4.4). The

matrix water saturation is assumed to be equal to residual water saturation, and the widths

of hydraulic and natural fractures are assumed to be 0.1 m to properly simulate water

flow back in early production period. Table 4.3 summarizes all the ranges used for the

eight uncertain parameters and deterministic values for the fixed parameters. The ranges

for the eight uncertain parameters are either based on history matching experience

(matrix permeability, conductivities, fracture permeability decay, etc.),

geological/engineering data (microseismic, DFIT, DTS, etc.) or suggestions from

engineers working on this field (fracture height, outer/inner fracture half-lengths). Figure

4.10 and Figure 4.11 contains the information regarding the maximum/minimum

47
pressure-dependent permeability decline and longest/shortest scenario for heel-biased

half-lengths, respectively.

Uncertain Parameters Distribution Minimum Maximum Unit

Matrix permeability Uniform 0.00001 0.001 md


Fracture permeability Uniform 0.0075 0.0425 1/kPa
decay factor
Fracture water saturation Uniform 0.5 0.9 -

Fracture height Uniform 5 20 m


Outer fracture Uniform 50 180 m
half-length
Inner fracture Uniform 20 80 m
half-length
Fracture conductivity Uniform 10 500 md-m
Natural fracture Uniform 1 20 md-m
conductivity
Exponent of water Deterministic 3.09 -
relative permeability
Exponent of gas relative Deterministic 2.86 -
permeability
Endpoint of water Deterministic 0.758 -
relative permeability

Table 4.3: Summary of uncertain parameters, their ranges, fixed parameters, and their
deterministic values used for automatic history matching.

48
Figure 4.10: Visualization of maximum and minimum permeability decline curve.

(a) Minimum outer and inner fracture half-lengths

Figure 4.11: continued next page.

49
(b) Maximum outer and inner fracture half-lengths

Figure 4.11: Visualization of maximum and minimum half-lengths values for both outer
and inner fractures: (a) Minimum outer and inner fracture half-lengths; (b)
Maximum outer and inner fracture half-lengths.

4.5 AUTOMATIC HISTORY MATCHING

We begin the process of automatic history matching by first designing the

location of proxy prediction points to calculate the objective function, a measure of

relative error, to be fed into the XGBOOST-MCMC algorithm. Usually for

unconventional reservoir with two phase flow, gas flow rate is used constraint as well

instead of flowing bottomhole pressure (BHP). This is because the bottomhole pressure is

interpolated from wellhead pressure (WHP). In this study, the conversion from WHP to

BHP based on a pressure loss correlation is shown in Figure 4.12. A total of 20 proxy

points is assigned for the proxy model to make predictions on bottomhole pressure and

water flow rate, and the location assignments are shown in Figure. 4.13. These included

10 proxies for bottomhole pressure at day 3, 20, 50, 85, 121, 153, 200, 238, 297, and 334

50
(Figure. 4.13a), 10 proxies for water flow rate at day 3, 20, 50, 85, 121, 153, 200, 238,

297, and 325 (Figure. 4.13b). In theory, we can include infinite proxy points, but 10

proxy points choice is considered optimal to balance accuracy and efficiency.

Figure 4.12: Conversion from WHP to BHP used in this study.

(a) Flowing bottomhole pressure

Figure 4.13: continued next page.


51
(b) Water flow rate

Figure 4.13: Locations of proxy prediction points for calculating values of response
variables’ objective functions during history matching iterations: (a)
Flowing BHP; (b) Water flow rate.

After determining the ranges of uncertain parameters and the proxy locations, we

can proceed to the first iteration of the history matching by using the Latin Hypercube

sampling process. The 50 initial iteration samples are generated, and their simulation

results are shown in Figure 4.14. Again, we used gas flow rate as constraint for all

simulation runs (shown in Figure 4.14a) because this response is measured relatively
accurately. For one of the samples, the gas flow rate cannot be constrained, probably due

to its’ sub-optimal combinations of uncertain parameters. For water flow rate (Figure

4.14c) and flowing BHP (Figure 4.14f), the current ranges covers the production history,

meaning there are samples both over-estimating and under-estimating the history data.

Therefore, we can proceed to subsequent iterations without modifying the ranges.

52
(a) Gas flow rate

(b) Cumulative gas production

Figure 4.14: continued next page.

53
(c) Water flow rate

(d) Cumulative water production

Figure 4.14: continued next page.

54
(e) Water gas ratio

(f) Flowing bottomhole pressure

Figure 4.14: Comparison between simulation results from initial iteration (Latin
Hypercube sampling process) and actual production history: (a) Gas flow
rate; (b) Cumulative gas production; (c) Water flow rate; (d) Cumulative
water production; (e) Water gas ratio; (f) Flowing BHP.
55
The automatic history matching process then continued iteratively. For each

iteration, we used single chain of MCMC and a proposal sample size of 20000. The

train:validation set ratio is 8:2, and the proxy model has an epoch size of 5 with a total

epoch number of 600. The maximum loop number for updating sampling variance is 8,

and the overall accept ratio should be between 0.15 and 0.51 for the posterior samples to

pass the filtering test. The variance update ratio is 2 for the chain if the overall accept

ratio is rejected. For the accepted posterior sample ensembles, the 25 samples with the

lowest global errors are selected to be inputted into the reservoir simulator. We also

defined different weight for each proxy location when calculating objective functions.

Proxies of BHP at days 3, 20, 297, and 334 and proxies of water flow rate at days 3, 20,

297, and 325 have weights of 1, while rest of the proxies have weights of 2. At early

period (days 3 and 20), the production is not stabilized, and the data does not necessarily

reflect the response from the reservoir. At late period (days 297+), there are some

fluctuations with the data as reflected in Figure 4.13. These two reasons caused the

weight of these proxies to be lower than the other ones.

For this study, a total of 10 extra iterations is run in addition to the initial iteration

(50 cases). For each of the extra iteration, 25 cases are included, thus giving a total of 300

simulation results. The entire history matching workflow took approximately 15 hours to

finish. We then filtered out the history matching solution out of the 300 simulations based

on one threshold for each of the response variables (flowing BHP and water flow rate).

Objective function threshold values of 10 and 33 are used for BHP and water flow rate,

respectively. The threshold value is stricter with BHP because the quality of BHP

response should be smoother, and issue of water flow back exists. A total of 55 solutions

satisfied the criteria. The number of solutions versus iteration run is visualized in Figure

56
4.15. As can be seen from this plot, the number of solutions first increases and then

decreases, implying that the maximum iteration number could be reduced to improve

efficiency. The simulation responses for these 55 history matching solutions can be

visualized in Figure 4.16. We can observe from this plot that all the history matching

solutions meet the well constraint of gas flow rate. In addition, the results for water flow

rate heavily concentrates near the actual production history, except for the fluctuations at

the late period. The results for pressure also cover the trend of the drawdown decline,

except for the time after the two shut-in periods. This can be explained by the operational

uncertainties associated with shut-in, and the quality of the match could be further

improved if shut-in time could be avoided. These general observations suggest that the

workflow successfully optimized the results compared to the initial iteration’s suboptimal

results. Figure 4.17 shows the simulation results for the best match model, and they

exhibit very satisfactory match with the actual production history.

Figure 4.15: Number of history matching solutions versus iteration number.

57
(a) Gas flow rate

(b) Cumulative gas production

Figure 4.16: continued next page.

58
(c) Water flow rate

(d) Cumulative water production

Figure 4.16: continued next page.

59
(e) Water gas ratio

(f) Flowing bottomhole pressure

Figure 4.16: Comparison between 55 filtered history matching solutions’ simulation


results and actual production history: (a) Gas flow rate; (b) Cumulative gas
production; (c) Water flow rate; (d) Cumulative water production; (e) Water
gas ratio; (f) Flowing BHP.
60
(a) Gas flow rate

(b) Cumulative gas production

Figure 4.17: continued next page.

61
(c) Water flow rate

(d) Cumulative water production

Figure 4.17: continued next page.

62
(e) Water gas ratio

(f) Flowing bottomhole pressure

Figure 4.17: Comparison between best match’s simulation results and actual production
history: (a) Gas flow rate; (b) Cumulative gas production; (c) Water flow
rate; (d) Cumulative water production; (e) Water gas ratio; (f) Flowing BHP.

63
The information regarding the best match model is given below. Matrix permeability is

38.1 nano-darcy (nd), fracture permeability decay factor is about 0.0249 and the decline

curve can be visualized in Figures 4.18 and 4.20. As can be implied from the

permeability decay curve, the degree of fracture closure is pronounced because the

fracture permeability multiplier reaches 0.18 when pressure drops to atmospheric

pressure (the range for this parameters produced a maximum multiplier of 0.6 to a

minimum multiplier of 0.05). Fracture water saturation is about 0.853 and fracture height

is 8.62 m. The outer fracture and inner fracture half-lengths are 130 m and 54.3 m,

respectively (visualized in Figure 4.19). Finally, conductivities of hydraulic fracture and

natural fractures are 90.8 md-m and 10.7 md-m. This best match model is generated

around the P50 values of all 55 history matching solutions to consider the effects of

uncertainties. From the plot, we can observe that the solution ensembles for matrix

permeability, fracture height, and fracture conductivity are more concentrated on the

lower end of their ranges, while solution ensembles for fracture water saturation and

outer fracture half-length are more concentrated on the higher end of their ranges. These

uncertain parameters more sensitively affect the dynamic production. However, other

parameters are less sensitive because any values within their ranges can compose the

solution ensembles. Usually, the half-length for hydraulic fractures are considered very

sensitive parameter. Nevertheless, in this study, the inner fractures with lower half-

lengths have less contributions to the production and are thus less sensitive. Parallel

coordinates plot is drawn in Figure 4.21 and it can also prove that the abovementioned

statements. More sensitive parameters should have solution ensemble lines passing

through focused part of the corresponding vertical coordinates, and less sensitive

parameters should have solution ensemble lines spanning the entire vertical coordinates.

64
(a) Matrix permeability (b) Fracture permeability decay factor

(c) Fracture water saturation (d) Fracture height

(e) Outer fracture half-length (f) Inner fracture half-length

Figure 4.18: continued next page.

65
(g) Hydraulic fracture conductivity (h) Natural fracture conductivity

Figure 4.18: Comparison between prior and posterior distributions of the studied
uncertain parameters based on 55 history matching solutions: (a) Matrix
permeability; (b) Fracture permeability decay factor; (c) Fracture water
saturation; (d) Fracture height; (e) Outer fracture half-length; (f) Inner
fracture half-length; (g) Hydraulic fracture conductivity; (h) Natural fracture
conductivity.

(a) Non-uniform half-lengths

Figure 4.19: continued next page.


66
(b) Fracture height

Figure 4.19: Characterized fracture geometry from AHM workflow: (a) Non-uniform
fracture half-lengths; (b) Fracture height.

Figure 4.20: Pressure-dependent fracture permeability decline curve of 55 history


matching solutions and the best match model.
67
Figure 4.21: Parallel coordinate plot to visualize combinations of uncertain parameters,
including non-history matching solutions (light grey); history matching
solution (light orange); and best match solution (red).

In order to gauge the accuracy and precision of the workflow, we plotted the

values of the two response variables (water flow rate and BHP) at the last proxy location

for both proxy model and the actual simulation runs (Figure 4.22a, b). In addition, the

objective function value for water flow rate and BHP, and global relative error are also

plotted, as shown in Figure 4.22c, d, e. Iteration 2, 4, 6, 8, and 10 are selected so that the
improvement of proxy model could be visualized along with progression of iterations.

From Figure 4.22, it is well perceived that the value of response variables tends to be

more consistent between proxy prediction and simulation runs, as training data for proxy

model accumulates. Furthermore, the number of maximum iteration could be reduced to

8, since the objective functions and global errors can be optimized at iteration 8.

68
(a) Water flow rate at 325 days

(b) Flowing bottomhole pressure at 334 days

Figure 4.22: continued next page.

69
(c) Water flow rate relative error

(d) Flowing bottomhole pressure relative error

Figure 4.22: continued next page.

70
(e) Global relative error

Figure 4.22: Comparison between response parameters’ values and relative errors of
proxy model and simulation results for iteration 2, 4, 6, 8 and 10: (a) Water
flow rate at 325 days; (b) Flowing BHP pressure at 334 days; (c) Water flow
rate relative error; (d) Flowing BHP relative error; (e) Global relative error.

We conclude this history matching section with the visualizations of values of

uncertain parameters and global error versus simulation cases (In Figure 4.23, red circles

are AHM solutions). From Figure 4.23a, c, we can conclude that the matrix permeability

and fracture water saturations are very sensitive to the dynamic production. Figure 4.23d,

e, g substantiates that fracture height, outer fracture half-length and hydraulic fracture

conductivity are also sensitive, but not at same level with matrix permeability and

fracture water saturation. Figure 4.23b, f, h shows that fracture permeability decay factor,

inner fracture half-length and natural fracture conductivity are least sensitive parameters

to dynamic production. Finally, Figure 4.23i shows that the workflow can optimize the

error quickly, and it is sufficient to end the workflow at simulation 8.


71
(a) Matrix permeability (b) Fracture permeability decay factor

(c) Fracture water saturation (d) Fracture height

(e) Outer fracture half-length (f) Inner fracture half-length

Figure 4.23: continued next page.

72
(g) Hydraulic fracture conductivity (h) Natural fracture conductivity

(i) Global error

Figure 4.23: Scatter plots of values of different uncertain parameters and global error
versus simulation index: (a) Matrix permeability; (b) Fracture permeability
decay factor; (c) Fracture water saturation; (d) Fracture height; (e) Outer
fracture half-length; (f) Inner fracture half-length; (g) Hydraulic fracture
conductivity; (h) Natural fracture conductivity; (i) global error.

4.6 PROBABILISTIC PRODUCTION FORECASTING

Next, we performed production forecasting using the 55 solutions obtained from

the previous section. The BHP drawdown schedule after the history period can be

visualized in Figure 4.24a. This schedule follows a gradual (spanning 9 months) decline

from the BHP value of the last history date of each solution’s simulation result and
73
remains at the minimum BHP constraint of 800 kPa until a total simulation time of 20

years is achieved. The reason we adopted this strategy is to avoid sudden increases in

water/gas flow rate right after history period ends. The gas flow rates, cumulative gas

productions, water flow rates, cumulative water productions are provided in Figure 4.24b,

c, d, e, respectively. Although the gas flow rates and water flow rates shows peaks after

production history, the magnitude of this peak is well controlled. For 20-year gas

estimated ultimate recovery (EUR), the P10, P50, P90 and best match values are 114.5,

124.4, 131.3, and 119.0 million cubic meters. For 20-year water EUR, the P10, P50, and

P90 and best match values are 26.7, 28.7, 30.8, and 28.9 thousand cubic meters. Gas

EUR for best match model tends to be little bit lower than the P50 values, and water EUR

is perfectly matching with the P50 values. Indeed, the integration of geology model and

the automatic history matching workflow greatly reduced the uncertain range for both 20-

year gas and water EUR. Without the heterogeneity of porosity, the ranges for these EUR

predictions would span much wider ranges.

(a) Flowing bottomhole pressure

Figure 4.24: continued next page.


74
(b) Gas flow rate

(c) Cumulative gas production

Figure 4.24: continued next page.

75
(d) Water flow rate

(e) Cumulative water production

Figure 4.24: Production forecast of 55 history matching solutions: (a) flowing BHP; (b)
Gas flow rate; (c) Cumulative gas production; (d) Water flow rate; (e)
Cumulative water production.

76
4.7 NATURAL FRACTURE SENSITIVITY STUDY AND VISUALIZATIONS

After probabilistic production forecast work is completed, we extended the study

to analyze the connectivity of natural fractures and its effects on the pressure drawdown

in both matrix and fractures, drainage volumes and 20-year gas EUR. Figure 4.25 shows

the grouped natural fractures and the total number of natural fractures in each group.

Although the number of connected fractures is high, the communication between these

fractures to near wellbore region is sub-optimal and the number of non-connected

fractures is not negligible. In fact, only 2 natural fractures are connected to the wellbore.

Therefore, the impact of the poor connectivity on gas EUR is profound. To prove this

point, we compared four scenario with different connectivity of natural fractures, which

are no natural fractures; the original dimension of natural fractures; natural fractures’

location preserved but their length and heights are magnified by 1.5 times (Figure 4.26a);

and natural fractures’ location preserved but their length and heights are magnified by 1.8

times (Figure 4.26b). The number of non-connected fractures (grey fractures in Figure

4.26) for 1.5 times magnification and 1.8 times magnification are 237 and 135, meaning

the connectivity significantly improved. In Figure 4.27, we also plotted the matrix

pressure distributions of the eighth layer from the top, where the most proportion of the

well lands in, for these four scenarios. These four models are created using the best match

properties, and the only difference is the presence of natural fractures. It is very

noticeable that for the scenario without natural fractures, the matrix pressure drawdown is

only confined to the region near the wellbore. However, as the presence and connectivity

of natural fractures increase, the pressure drawdown area extends to region far away from

the wellbore. In addition, for the enlarged natural fractures (Figure 4.27c, d), the pressure

drawdown are connected from the 7th layer to 8th layer, as opposed to Figure 4.27a, b

77
where the toe side region of 8th layer do not display any drawdown because the well

trajectory penetrated the top layer. Figure 4.28 depicts the drainage volumes for the four

scenarios studied. For no natural fractures and original natural fractures scenarios (Figure

4.28a, b), the drainage volumes are enclosed only in bottom five layers, or only small

portions of top five layers. However, the enlarged natural fracture scenarios exhibit

prominent drainage in top five layers as well. Figure 4.29 describes the pressure

drawdown in all the fractures. The original natural fracture scenario (Figure 4.29b) shows

a very limited network of natural fractures that is connected to hydraulic fractures.

However, with the enlarged natural fractures, a significant network is developed and with

this improved connectivity, more gas could be unlocked from this low permeability shale

reservoir. The comparison between 20-year gas EUR of the four scenarios are shown in

Figure 4.30. The gas EUR for these four scenarios are 90.9, 123.8, 219.5 and 297.1

million cubic meters (or -26.6%, 0%, 77.3%, 140% comparatively if the original scenario

is the base case).

(a) Spatial distribution for different natural fracture groups (color coded)

Figure 4.25: continued next page.


78
(b) Total number of natural fractures in different groups

Figure 4.25: Connectivity analysis of the natural fractures presented in the history
matched model: (a) Spatial distribution for different natural fracture groups;
(b) Total number of natural fractures in different groups.

(a) Magnified 1.5 times in both lengths and heights

Figure 4.26: continued next page.


79
(b) Magnified 1.5 times in both lengths and heights

Figure 4.26: Spatial connectivity analysis of the natural fractures magnified: (a) 1.5 times
in both lengths and heights; (b) 1.8 times in both lengths and heights.

(a) No natural fractures

Figure 4.27: continued next page.

80
(b) Original natural fractures

(c) Natural fractures enlarged 1.5 times in both length and height

(d) Natural fractures enlarged 1.8 times in both length and height

Figure 4.27: Visualizations of the effects of the natural fractures’ connectivity on matrix
pressure distributions after 20 years: (a) No natural fractures; (b) Original
natural fractures; (c) Natural fractures enlarged 1.5 times in both length and
height; (d) Natural fractures enlarged 1.8 times in both length and height.

81
(a) No natural fractures

(b) Original natural fractures

Figure 4.28: continued next page.

82
(c) Natural fractures enlarged 1.5 times in both length and height

(d) Natural fractures enlarged 1.8 times in both length and height

Figure 4.28: Visualizations of the effects of the natural fractures’ connectivity on


drainage volumes after 20 years: (a) No natural fractures; (b) Original
natural fractures; (c) Natural fractures enlarged 1.5 times in both length and
height; (d) Natural fractures enlarged 1.8 times in both length and height.
83
(a) No natural fractures

(b) Original natural fractures

(c) Natural fractures enlarged 1.5 times in both length and height

Figure 4.29: continued next page.

84
(d) Natural fractures enlarged 1.8 times in both length and height

Figure 4.29: Visualizations of the effects of the natural fractures’ connectivity on fracture
pressure distributions after 20 years: (a) No natural fractures; (b) Original
natural fractures; (c) Natural fractures enlarged 1.5 times in both length and
height; (d) Natural fractures enlarged 1.8 times in both length and height.

Figure 4.30: Comparison of 20 year gas estimated ultimate recovery for four different
natural fracture connectivity scenarios.

85
4.8 CONCLUSIONS

We developed the automatic history matching workflow by integrating the real

geology model with a modified proxy engine to achieve practicality, efficiency, and

proximity with reality. The workflow performed history match and statistical production

forecast for a real case of shale gas well. In this chapter, we concluded the followings:

1. Our developed workflow accurately executed the history matching of flowing

BHP and water rate for a shale gas well. The workflow found 55 HM solutions

with 8 uncertain parameters from 11 iterations or 300 total simulation runs, which

is around 18%. The total computational time is around 15 hours.

2. The eight posterior distributions were obtained from the workflow. The matrix

permeability, fracture water saturations are considered the most sensitive

uncertain parameter; fracture heigh, outer fracture half-length and hydraulic

fracture conductivity are moderately sensitive; and fracture permeability decay

factor, inner fracture half-length and natural fracture conductivity are least

sensitive.

3. The heterogeneous porosity field reduces the uncertainties associated with the

reservoir and thus improves quality of probabilistic production forecasting. The

non-uniform half-lengths of hydraulic fractures provides valuable benefits

towards optimal well spacing designs and future fracking job designs.

4. The connectivity study of the natural fractures also provides suggestions

regarding inter-well/inter-layer communications.

5. More subsurface data, such as DTS/DAS/etc., could be coupled with our

workflow to further reduce uncertainties and supply field engineers with practical

guidance.

86
Chapter 5: Development of Realistic Natural Fracture Module in AHM
by Using Fractal Theory

5.1 INTRODUCTION

Due to the facts that extensive number of natural fractures and faults exists among

the shale gas reservoirs; it is paramount to incorporate natural fractures within the

numerical model. These geological features have huge implications regarding long-term

EUR, well spacing optimization, and fracturing/refracturing job design. Therefore, the

objective of this chapter is to study the effects of natural fractures on the results of history

matching. Instead of random generation of natural fractures with arbitrary fracture

lengths, our workflow utilizes the fractal theory to generate natural fractures in 3D space.

Darcel et al. (2003a) proposed that natural fracture networks in nature, as observed in

various outcrops, follow two unique geometrical patterns: fractal distributions of fracture

centers and power law distributions of lengths. Hence, the proposed workflow is

beneficial because it is more realistic than random natural fracture generation, which

lacks physical basis.

All three main steps in the AHM workflow were performed including parameter

identification, history matching and probabilistic production forecasting. In this study, we

used another shale gas single well with 1626-day historical production data. To optimize

history matching efficiency, we included an extra step to analyze the fractal network’s

connectivity and remove the non-connected fractures from the network. Similar

constraints and BHP pressure conversion from last chapter is performed here.

5.2 RESERVOIR MODEL

The novel workflow is applied to a real shale gas well in Sichuan Basin in China.

We performed the case with both hydraulic fractures and natural fractures. The hydraulic
87
fractures are modelled considering the heel bias effect, which means the outer 2 fractures

in each stage have longer half-lengths than the fracture(s) in the middle. This design

actually honors the output characteristics of fracture modelling software, and is deemed

more realistic than simple, planar fractures with uniform half-lengths. For natural

fractures, we inputted the fractal dimension, fractal length dimension and alpha regulator

to be uncertain parameters, meaning that the total number of natural fractures will not be

constant. This field case contains a single well with a horizontal section length of around

1550 m. There is a total of 18 stages of hydraulic fractures with a total cluster number of

54, with a constant clusters per stage of 3. The cluster spacing is approximately 27.78

meters and it is equivalent to stage spacing.

For this study, since the geology model for this reservoir is lacking, we created a

homogeneous, rectangular reservoir model with a dimension of 1700 m long, 800 m wide

and a thickness of 20 m, with numbers of grid blocks in x, y, z dimensions 170, 80, 1

respectively. An example model visualization for the non-uniform hydraulic fractures and

fractal natural fracture scenario are provided in Figure 5.1. The fractal based natural

fractures are fully 3D, meaning that complex dip/azimuth/height distributions are given

for natural fractures. The reservoir has a uniform depth of 2500 m and a residual water

saturation of 0.30, which is equal to matrix water saturation. The initial reservoir pressure

is 44.79 MPa with a temperature of 100 ℃. A long history of 1626 days is available for

history matching. In normal situations, the available history is usually no longer than 1

year. Therefore, the availability of this long production history can truly test the

workflow’s robustness to capture long term production trends. Basic reservoir, fractures,

and Langmuir isotherm gas desorption properties are summarized in Table 5.1. The

relative permeability curves for modeling gas and water flow are assumed to be fixed

88
because they affect the dynamic production less sensitively. We adopted non-linear

relative permeability curves for matrix and straight-line relative permeability curves for

fractures. These fixed curves are shown in Figure 5.2.

Model parameters Value Unit


Model dimension (x × y × z) 1700 × 800 × 20 m
Number of grid blocks (x × y × z) 170 × 80 × 1 -
Initial reservoir pressure 44.79 MPa
Reservoir temperature 100 ℃
Residual water saturation 30% -
Total compressibility 4.35×10-7 kPa-1
Reservoir depth 2500 m
Well length 1550 m
Number of stages 18 -
Clusters per stage 3 -
Average stage/cluster spacing 27.78 m
Langmuir pressure 13.26 MPa
Langmuir adsorption volume 2.96 m3/ton
Rock density 2.6 g/cm3

Table 5.1: Summary of basic reservoir, fracture and gas desorption parameters used in
this study.

Figure 5.1: A 3D example model visualizations for the reservoir, hydraulic fractures
(blue) and fractal natural fractures (green) in this study.
89
(a) Matrix

(b) Hydraulic fractures

Figure 5.2: Fixed relative permeability used for this study: (a) Matrix; (b) Hydraulic
fractures.

90
5.3 CONNECTIVITY ANALYSIS AND EFFICIENCY OPTIMIZATION

To facilitate the efficiency of the workflow, an extra procedure is built into the

workflow. This procedure detects the connectivity between the fractal generated natural

fractures, hydraulic fractures, and wellbore. Only the natural fractures that are connected

to wellbore or hydraulic fractures are preserved for history matching purpose. The rest of

natural fractures that are isolated are removed from the reservoir model to save

computational time. For example, in Figure 5.3, pink fractures are isolated and will be

removed from the model. This procedure should not affect the results of history

matching/ production forecast since negligible contributions from these fractures are

observed. To validate this statement, we used the field production data of actual gas flow

rate as the well constraint for the history period, and then reduced bottomhole pressure

from the last history day stepwise to a minimum of 1000 kPa until 20 years (Figure 5.4a).

Figure 5.4b, c prove that if we remove the non-connected fractures, negligible

discrepancies can be found for long term cumulative gas and water production.

Figure 5.3: Connectivity check and non-connected natural fractures (pink) removal for
faster history matching.
91
(a) Bottomhole pressure

(b) Cumulative gas production

Figure 5.4: continued next page.

92
(c) Cumulative water production

Figure 5.4: Illustration of negligible effects from removing non-connected fractures


from the natural fracture network based on fractal theory: (a) Bottomhole
pressure; (b) Cumulative gas production; (c) Cumulative water production.

5.4 PARAMETERS IDENTIFICATIONS

A total of 11 important parameters are determined to be uncertain parameters to

be inputted into the history matching workflow. These parameters include matrix

permeability, fracture permeability decay factor, fracture water saturation, porosity,

fracture height, outer fracture half-length, inner fracture half-length, fracture

conductivity, 3D fractal dimension, 3D length dimension, and 3D alpha regulator. The

first eight parameters are important parameters related to reservoir and engineering, as

stated in Chapter 4. Fractal/length dimensions control the spatial distribution/max-min

length ratio of the natural fractures, and alpha regulator controls the overall natural

fracture density. Therefore, the fractal dimension really provides information regarding

93
what degree of randomness is associated with the fractal network, while the length

dimension reflects what ratio of faults and smaller micro-fractures/fissures exists. Darcel

et al. (2003a) states that the fractal dimension ranges between 1.5 to 2 for 2D (2.5 to 3 for

3D), but we focused the range more on lower end because the lower fractal dimension is,

the fractures tend to be less randomly distributed and rather exist in regional clusters. In

addition, the lower the length dimension, the larger is the difference between maximum

and minimum natural fracture lengths. The length dimension range here is designed also

to emphasis the combined effects of both faults and fractures on a smaller scale.

Because the permeability and aperture of natural fracture are much insensitive to

dynamic production, we assume constant values of 50 md and 0.01 m for them. Table 5.2

summarizes all the ranges used for the eleven uncertain parameters and the values for the

constant parameters. The ranges for the first eight uncertain parameters are based on the

experience from the last chapter, since the well in this study is situated in the region very

close to the shale gas well, we studied in the last chapter. The ranges for the fractal

parameters are determined so that the total number of natural fractures (both connected

and isolated) are controlled below 5000 to optimize computational speed. For better

conceptualization of the fractal parameters on natural fractures’ spatial characteristics, we

included sensitivity plots of fractal visualizations in Figure 5.5.

94
Uncertain Parameters Distribution Minimum Maximum Unit
Matrix permeability Uniform 0.00001 0.001 md
Fracture permeability decay
Uniform 0.01 0.052 -
factor
Fracture water saturation Uniform 0.5 0.9 -
Matrix porosity Uniform 5% 9% -
Fracture height Uniform 5 20 m
Outer fracture half-length Uniform 50 180 m
Inner fracture half-length Uniform 20 80 m
Fracture conductivity Uniform 10 500 md-m
3D fractal dimension (𝐷𝑓 ) Uniform 2.6 2.8 -
3D length dimension (𝐷𝑙 ) Uniform 2.3 2.6 -
3D alpha regulator (𝛼) Uniform 4 5 -
Exponent of water relative
Deterministic 1.95 -
permeability
Exponent of gas relative
Deterministic 1.8 -
permeability
Endpoint of water relative
Deterministic 0.842 -
permeability

Table 5.2: Summary of uncertain parameters, their ranges, fixed parameters, and their
deterministic values used for automatic history matching.

(a) Df=2.6, Dl=2.45, α=4.5

Figure 5.5: continued next page.


95
(b) Df=2.8, Dl=2.45, α=4.5

(c) Df=2.7, Dl=2.3, α=4.5

Figure 5.5: continued next page.

96
(d) Df=2.7, Dl=2.6, α=4.5

(e) Df=2.7, Dl=2.45, α=4

Figure 5.5: continued next page.

97
(f) Df=2.7, Dl=2.45, α=5

Figure 5.5: Sensitivity plots of varying fractal parameters vs natural fracture


distributions: (a) Df=2.6, Dl=2.45, α=4.5; (b) Df=2.8, Dl=2.45, α=4.5; (c)
Df=2.7, Dl=2.3, α=4.5; (d) Df=2.7, Dl=2.6, α=4.5; (e) Df=2.7, Dl=2.45,
α=4; (f) Df=2.7, Dl=2.45, α=5.

5.5 HISTORY MATCHING

Similar to last chapter, we begin the process of automatic history matching by

converting the WHP to BHP, as shown in Figure 5.6. Next, we selected a total of 20

proxy points for the proxy model to make predictions on bottomhole pressure and water

gas ratio (WGR), and the location assignments are shown in Figure. 5.7. In this study,

WGR was used as the response variable instead of water flow rate because it is less

noisy. The selected proxy locations included 10 proxies for bottomhole pressure at day

30, 90, 226, 374, 500, 745, 1000, 1200, 1400, and 1626 (Figure. 5.7a), 10 proxies for

water gas ratio at day 20, 140, 226, 374, 500, 743, 983, 1200, 1400, and 1623 (Figure.

5.7b). Although the historical data length is much longer than the one in the last chapter,

98
we did not increase the proxy points because this would jeopardize computational

efficiency extensively, without many improvements on the results.

Figure 5.6: Conversion from WHP to BHP used in this study.

(a) Flowing bottomhole pressure

Figure 5.7: continued next page.

99
(b) Water gas ratio

Figure 5.7: Locations of proxy prediction points for calculating values of response
variables’ objective functions during history matching iterations: (a)
Flowing BHP; (b) Water gas ratio.

Proceeding in a similar manner as the last chapter, we generated the 50 initial

Latin Hypercube samples, run the simulations for these samples and obtained the

corresponding results, as shown in Figure 5.8. Similarly, gas flow rate is used as
constraint for all simulation. Most initial samples can constrain the gas flow rate very

well, as can be seen from the Figure 5.8a. For water flow rate (Figure 5.8c) and flowing

BHP (Figure 5.8f), the current ranges covers the production history, meaning there are

samples both over-estimating and under-estimating the history data. Even though the

initial samples tend to overestimate flowing BHP, our algorithm should be able to find

optimal solution, as will be proved later in this section.

100
(a) Gas flow rate

(b) Cumulative gas production

Figure 5.8: continued next page.

101
(c) Water flow rate

(d) Cumulative water production

Figure 5.8: continued next page.

102
(e) Water gas ratio

(f) Flowing bottomhole pressure

Figure 5.8: Comparison between simulation results from initial iteration (Latin
Hypercube sampling process) and actual production history: (a) Gas flow
rate; (b) Cumulative gas production; (c) Water flow rate; (d) Cumulative
water production; (e) Water gas ratio; (f) Flowing BHP.
103
The automatic history matching process then continued iteratively. For each

iteration, we used single chain of MCMC and a proposal size of 30000. The other

parameters for MCMC are kept same as the last chapter. Similarly, differing weights for

each proxy location when calculating objective functions are defined. The first three

proxies of BHP and WGR at have weights of 1, while rest of the proxies have weights of

2. This is because there are huge uncertainties associated with water flow rate data at

early period, due to water flow back. There is also a rapid drop in BHP, meaning the well

is not very stabilized at early period.

Likewise, a total of 11 iterations, or 300 simulations are designed for this study.

The entire history matching workflow took approximately 13 hours to finish. Next,

history matching solutions are filtered out based on objective function threshold values of

35 and 70 for BHP and WGR. Each criterion is relaxed compared to the last chapter,

because more data uncertainties are introduced as production history increases. A total of

52 solutions are filtered out. The number of solutions versus iteration run is given in

Figure 5.9. This figure implies that the quality of proxy model improves progressively

with the iteration number, and the last 2 iterations can be removed to optimize efficiency.

The resulting production curves of response variables for these 52 history matching

solutions can be visualized in Figure 5.10. The figure shows that despite the early

period’s data fluctuations, our workflow is capable to match the late period production

data excellently even though the initial iteration overestimates actual BHP. Figure 5.11

shows the simulation results for the best match model, which proves that our proposed

workflow is robust in history matching production data that spans much longer time

period.

104
Figure 5.9: Number of history matching solutions versus iteration number.

(a) Gas flow rate

Figure 5.10: continued next page.

105
(b) Cumulative gas production

(c) Water flow rate

Figure 5.10: continued next page.

106
(d) Cumulative water production

(e) Water gas ratio

Figure 5.10: continued next page.

107
(f) Flowing bottomhole pressure

Figure 5.10: Comparison between 52 filtered history matching solutions’ simulation


results and actual production history: (a) Gas flow rate; (b) Cumulative gas
production; (c) Water flow rate; (d) Cumulative water production; (e) Water
gas ratio; (f) Flowing BHP.

(a) Gas flow rate

Figure 5.11: continued next page.


108
(b) Cumulative gas production

(c) Water flow rate

Figure 5.11: continued next page.

109
(d) Cumulative water production

(e) Water gas ratio

Figure 5.11: continued next page.

110
(f) Flowing bottomhole pressure

Figure 5.11: Comparison between best match’s simulation results and actual production
history: (a) Gas flow rate; (b) Cumulative gas production; (c) Water flow
rate; (d) Cumulative water production; (e) Water gas ratio; (f) Flowing BHP.

The following information contains the best match model parameters’ values.

Matrix permeability is 36.2 nano-darcy (nd), fracture permeability decay factor is about

0.0307 and the decline curve can be visualized in Figures 5.12 and 5.14. These results the

equivalent matrix permeability is low and thus the cluster spacing could be further
reduced to maximize recovery. The fracture permeability multiplier reaches 0.25 when

pressure drops to atmospheric pressure, meaning the fracture closure phenomenon is less

severe than the study in Chapter 4. Because the depth of the reservoir in this study is

shallower, and assuming similar fracture job design is applied for both study, the

compaction should be less severe for this study and thus less conductivity loss occurs,

which is corroborated by our results. Matrix porosity is 6.8 %. Compared to the previous

111
chapter, in which the model has a greater depth and a lower average porosity between 4-

5%, the obtained porosity is again backed up by physics. The fracture height, outer

fracture and inner fracture half-lengths are 14 m, 65.6 m and 57 m, respectively

(visualized in Figure 5.13). The fracture height is able to grow higher due to shallower

depth, but half-lengths tend to be limited possibly because of the effect of natural

fractures. Therefore, it is very important to rigorously model natural fractures for

geologists. Fracture water saturation is about 0.708 and conductivity of hydraulic

fractures 256 md-m, implying a less severe water flow back problem. Finally, the 3D

fractal dimension, 3d length dimension and 3d alpha regulator characterized are 2.65,

2.58, and 4.29. These results imply that the natural fractures tend to be less randomly

distributed, but the length discrepancy between different scales of natural fractures is low.

There is also a relatively lower density of natural fractures. Parallel coordinates plot is

shown in Figure 5.15, and we see that matrix permeability, fracture height, outer/inner

fracture half-lengths, 3D fractal/length dimensions are more sensitive parameters.

(a) Matrix permeability (b) Fracture permeability decay factor

Figure 5.12: continued next page.

112
(c) Fracture water saturation (d) Matrix porosity

(e) Fracture height (f) Outer fracture half-length

(g) Inner fracture half-length (h) Fracture conductivity

Figure 5.12: continued next page.

113
(i) 3D fractal dimension (j) 3D length dimension

(k) 3D alpha regulator

Figure 5.12: Comparison between prior and posterior distributions of the studied
uncertain parameters based on 52 history matching solutions: (a) Matrix
permeability; (b) Fracture permeability decay factor; (c) Fracture water
saturation; (d) Matrix porosity; (e) Fracture height; (f) Outer fracture half-
length; (g) Inner fracture half-length; (h) Hydraulic fracture conductivity; (i)
3D fractal dimension; (j) 3D length dimension; (k) 3D alpha regulator.

114
(a) Overview

(b) Non-uniform half-lengths

Figure 5.13: continued next page.

115
(c) Fracture height

Figure 5.13: Characterized fracture geometry from AHM workflow: (a) Overview; (b)
Non-uniform fracture half-lengths; (c) Fracture height.

Figure 5.14: Pressure-dependent fracture permeability decline curve of 52 history


matching solutions and the best match model.
116
Figure 5.15: Parallel coordinate plot to visualize combinations of uncertain parameters,
including non-history matching solutions (light grey); history matching
solution (light orange); and best match solution (red).

In order to gauge the robustness of the workflow to catch long term production

trend, we plotted the values of the BHP and WGR at the early period (Figure 5.16a, b),

mid-period (Figure 5.16c, d), and late period (Figure 5.16e, f). In addition, the objective

function value for water flow rate and BHP, and global relative error are also plotted, as

shown in Figure 5.16g, h, i, respectively. Iteration 2, 4, 6, 8, 10 are selected so that the


improvement of proxy model could be visualized along with progression of iterations.

For WGR at 1623 days, the values for water flow rate is almost 0, making the values for

proxy/simulation very low. From Figure 5.16, the proxy quality, especially for BHP

predictions, becomes progressively more accurate. However, the performance of the

workflow could be improved if we stop 2 iterations earlier, at iteration 8.

117
(a) BHP at 30 days (b) WGR at 20 days

(c) BHP at 745 days (d) WGR at 743 days

(e) BHP at 1626 days (f) WGR at 1623 days

Figure 5.16: continued next page.


118
(g) BHP relative error (h) WGR relative error

(i) Global relative error

Figure 5.16: Comparison between response parameters’ values and relative errors of
proxy model and simulation results for iteration 2, 4, 6, 8 and 10: (a) BHP at
30 days; (b) WGR at 20 days; (c) BHP at 745 days; (d) WGR at 743 days;
(e) BHP at 1626 days; (f) WGR at 1623 days; (g) BHP relative error; (h)
WGR relative error; (i) Global relative error.

Lastly, we plotted the visualizations of values of uncertain parameters and global

error versus simulation cases (In Figure 5.17, red circles are AHM solutions). This figure

proves our previous statements about sensitivities of the various uncertain parameters.

119
For global error (Figure 5.17l), the average values for history matching solutions are

higher than the one obtained in last chapter, because of the longer data history.

(a) Matrix permeability (b) Fracture permeability decay factor

(c) Fracture water saturation (d) Matrix porosity

Figure 5.17: continued next page.

120
(e) Fracture height (f) Outer fracture half-length

(g) Inner fracture half-length (h) Fracture conductivity

(i) 3D fractal dimension (j) 3D length dimension

Figure 5.17: continued next page.

121
(k) 3D alpha regulator (l) Global error

Figure 5.17: Scatter plots of values of different uncertain parameters and global error
versus simulation index: (a) Matrix permeability; (b) Fracture permeability
decay factor; (c) Fracture water saturation; (d) Matrix porosity; (e) Fracture
height; (f) Outer fracture half-length; (g) Inner fracture half-length; (h)
Hydraulic fracture conductivity; (i) 3D fractal dimension; (j) 3D length
dimension; (k) 3D alpha regulator; (l) Global error.

5.6 PROBABILISTIC PRODUCTION FORECASTING

Next, we performed probabilistic production forecasting using the obtained 52

history matching solutions. The BHP drawdown schedule after the history matching

period is visualized in Figure 5.18a. To maintain the BHP drawdown rate from the late

history period, we designed a gradual, linear drawdown in 60 months that reduces to a


minimum BHP of 1000 kPa. The gas flow rates, cumulative gas productions, water flow

rates, cumulative water productions are provided in Figure 5.18b, c, d, e, respectively.

Because the history data spans 4 years, the reservoir has lost a large amount of energy

and there is no more flow rate peaks after history period. For 20-year gas estimated

ultimate recovery (EUR), the P10, P50, P90 and best match values are 112.0, 121.2,

127.7, and 115.1 million cubic meters. For 20-year water EUR, the P10, P50, P90 and

best match values are 5.24, 6.50, 7.91, and 7.06 thousand cubic meters. Compared to last
122
chapter, the ranges for the probabilistic production forecasting in this study, without

porosity heterogeneity and inclusion of natural fractures, span more widely, further

validating that the characterization of natural fracture distributions and geology model is

crucial in determining the productivity of the shale gas reservoir.

(a) Flowing bottomhole pressure

(b) Gas flow rate

Figure 5.18: continued next page.

123
(c) Cumulative gas production

(d) Water flow rate

Figure 5.18: continued next page.

124
(e) Cumulative water production

Figure 5.18: Production forecasting of 52 history matching solutions: (a) Flowing BHP;
(b) Gas flow rate; (c) Cumulative gas production; (d) Water flow rate; (e)
Cumulative water production.

5.7 MATRIX AND FRACTURE PRESSURE VISUALIZATIONS

At last, the pressure in matrix and complex fracture network are visualized

separately for both the short term and long term. In Figure 5.19, we plotted the matrix

pressure at 365 days, end of history period (1626 days), 10 years and 20 years. From this

figure, we can see that the drainage area gradually increases as time progresses. However,

this area is only confined to the region near wellbore, implying that the characterized

fractal network has very deficient connectivity. Figure 20 contains the pressure in

fractures at the same time. This figure again shows that the natural fracture network in

this reservoir is under-developed.

125
(a) 365 days

(b) End of history (1626 days)

(c) 10 years

Figure 5.19: continued next page.


126
(d) 20 years

Figure 5.19: Pressure distribution of matrix at different production time: (a) 365 days; (b)
End of history (1626 days); (c) 10 years; (d) 20 years.

(a) 365 days

(b) End of history (1626 days)

Figure 5.20: continued next page.

127
(c) 10 years

(d) 20 years

Figure 5.20: Pressure distribution of complex fracture network at different production


time: (a) 365 days; (b) End of history (1626 days); (c) 10 years; (d) 20 years.

5.8 CONCLUSIONS

We applied the AHM workflow to another shale gas well in southwest china and

investigated the effects of fractal based natural fracture networks on history matching
results. It is found out that the presence of natural fractures reduced possibility of

existence of longer and taller fractures. In other words, our workflow is robust enough to

substantiate the fact that the existence of natural fractures has profound impacts on the

stress shadowing effect of hydraulic fractures propagation. From another perspective, the

history matching results suggest that the natural fractures in this shale reservoir,

according to fractal theory, are less randomly distributed, have a smaller number of faults

and have a rather low density. Compared to the study performed in the last chapter and
128
field experience from geologists, a much more developed natural fracture network is

usually seen in the deeper shale gas reservoirs, and our results indeed agree with the

statement. To close the loop, more subsurface data (such as good-quality microseismic

data, well image log, etc.) should be introduced to further validate the findings if

available.

129
Chapter 6: Summary, Conclusions, and Recommendations for Future
Work

This chapter aims to summarize all the conclusions and findings from this study,

and to deliver some of the potential suggestions for future improvements and extensions

for this study.

6.1 SUMMARY AND CONCLUSIONS

In this study, we developed an automatic history matching (AHM) workflow for

shale reservoirs with natural fractures characterized from both a geology model and based

on fractal theory. This workflow is made possible by implementation of EDFM in both

simple structural grids and complex corner point grids. In addition, the coupling of

XGBoost with MCMC algorithm enables us to optimize the workflow efficiency and

obtain accurate matching results. These results are then applied to probabilistically

perform production forecast, or EUR estimations.

In Chapter 4, we successfully integrated the geology model in corner point

gridding system and geologist-characterized natural fracture network into the AHM

process for a shale gas well. We introduced the complex characteristics of the reservoir,

discussed the difficulties of inputting fractures in this model, and explained how this

problem is solved. Then, we employed the AHM workflow to this reservoir model and

obtained satisfactory results. Probabilistic EUR estimations and natural fracture

connectivity analysis are lastly studied. We found out that the inclusion of porosity

heterogeneity reduced the uncertainties associated with the probabilistic EUR forecasts,

and that the geologist-characterized natural fracture network has very limited

connectivity with both the wellbore and hydraulic fractures, creating a bottleneck for the

20-year EUR values.


130
In Chapter 5, we explored the effects of realistic natural fracture networks on the

results of automatic history matching. Again, each of the module in our proposed AHM

workflow is implemented and acceptable results are achieved. In this chapter, we found

out that the presence of natural fracture does affect the growth of hydraulic fractures, and

based on the characterizations of fractal parameters, we know that the natural fractures in

this reservoir are less randomly distributed and more fractal-distributed. Furthermore,

there is a small number of large-scale fractures and a low natural fracture density.

This study is very beneficial, since the integration of the geology with engineering

is always neglected in most of the previous reservoir characterization studies. The

workflow proposed here is efficient, accurate, and generalized enough to be applied to

any shale reservoirs. By using a physical geology model, a realistic natural fracture

description, and robust AHM workflow to characterize reservoir properties and

uncertainties, invaluable guidance regarding well spacing optimization and engineering

design optimizations is made with ease.

6.2 RECOMMENDATIONS FOR FUTURE WORK

To further improve/extend some aspects of this study and make the numerical

optimization process more realistic, the following bullet points are enumerated:

• For MCMC sampling, we utilized the Metropolis-Hasting (MH) algorithm. There

are other algorithms such as Gibbs sampling, and Hamiltonian sampling.

Furthermore, as the technology of machine learning rapidly improves nowadays,

more advanced proxy models could be available. The exploration of other

sampling methods, potential suitable proxy models and their results comparison

would bring a more wholesome picture for the iterative optimization module.

131
• Microseismic data could be brought into the AHM process. For example, plane

fittings to seismic events and perform automatic cut off of the originally fitted

planes to obtain equivalent half-length and height of hydraulic fractures would

enhance the physical basis between AHM.

• Generation of fractal natural fracture network could be improved. The current

algorithm could not control the total number of natural fractures. Moreover,

physical measurements such as wellbore image logs and microseismic events

should be coupled with fractal theory to delineate natural fracture distributions in

unconventional reservoirs.

• Distributed acoustic sensing (DAS) and distributed temperature sensing (DTS)

data could be utilized in the AHM workflow to further reduce the uncertainties

associated with hydraulic fracture geometry.

• Outputs from specialized fracture modelling software that consider geomechanics

could be integrated with our AHM workflow, and automatic cutoff features can

be added to deliver another source of reference.

132
Glossary

ACRONYMS

AHM = Automatic history matching

AI = Artificial intelligence

ANN = Artificial neural network

BHP = Bottomhole pressure

DAS = Distributed acoustic sensing

DFIT = Diagnostic fracture injection test

DNN = Deep neural network

DPDK = Dual porosity dual permeability model

DTS = Distributed temperature sensing

EDFM = Embedded discrete fracture model

EUR = Estimated ultimate recovery

HFs = Hydraulic fractures

ISIP = Initial shut-in pressure

KNN = K-nearest neighbors

LGR = Local grid refinement


LH = Latin Hypercube

MCMC = Markov-chain Monte Carlo

MD = Measured depth

MH = Metropolis-Hasting

NFs = Natural fractures

NN = Neural network

NNC = Non-neighboring connection


133
P = Pressure

POR = Porosity

RTA = Rate transient analysis

SAGD = Steam-assisted gravity drainage

WGR = Water gas ratio

WHP = Wellhead pressure

NOMENCLATURE

𝐷𝑐3𝐷 = 3D fractal dimension number

dip = Angle between hydraulic fracture plane and x-y plane

𝐷𝑙 = 2D length dimension number

𝐷𝑙3𝐷 = 3D length dimension number


𝐹𝑗 = Objective function value for response variable j
ℎ𝑓 = Height of the hydraulic fracture

i = Index of the proxy location

𝑘′ = Reduced permeability at a specific reservoir pressure


𝑘𝑃𝑖 = Original permeability at initial reservoir pressure

𝐾 = Inverse Langmuir pressure

𝐿𝑋𝐺𝐵 = Overall loss function of XGBoost model

Loss function value of XGBoost model at data point i given input


𝐿(𝑦𝑖 , 𝐹(𝑥𝑖 )) =
data of xi and output data of yi

𝑀 = Total number of sub-loss function of XGBoost model

𝑀𝐷ℎ = Measured depth of the adjacent well trajectory point toward heel side
𝑀𝐷𝑝 = Measured depth of an arbitrary perforation point

134
𝑀𝐷𝑡 = Measured depth of the adjacent well trajectory point toward toe side

Density of natural fractures given a minimum fracture length of 𝑙


𝑛(𝑙, 𝐿) =
and a model scale of 𝐿

𝑁 = Total training size for XGBoost model

p = Total proxy locations selected for the response variable j

𝑃𝑖 = Initial reservoir pressure

𝑃𝑚 = Matrix pressure

𝑃′ = A specific reservoir pressure after production begins

𝑟 = Interpolation ratio

𝑠𝑡𝑟𝑖𝑘𝑒 = Angle between positive x-axis and middle axis

𝑡𝑖𝑙𝑡𝑖𝑛𝑔 = Angle between x-y plane and the middle axis

𝑉̅𝑚𝑎𝑥 = Maximum volume of absorbed gas per unit mass of rock

wi = Weight assigned to the proxy point i, given a response variable j

𝑊𝐺𝑅𝑚𝑎𝑥 = Maximum water gas ratio observed


𝑥𝑓 = Half-length of the bi-wing planar hydraulic fracture

𝑥𝑚𝑎 = X coordinates of two endpoints of the fracture’s middle axis


𝑥𝑝 = X coordinate of an arbitrary perforation point

𝑥𝑣 = X coordinates of the four vertices of the hydraulic fracture



𝑥𝑤 = X coordinate of the adjacent well trajectory point toward heel side
𝑡
𝑥𝑤 = X coordinate of the adjacent well trajectory point toward toe side
𝑦𝑖,ℎ𝑖𝑠𝑡 = Value of the response variable j at time i from actual history data

Value of the response variable j at time i from simulation run or


𝑦𝑖,𝑚𝑜𝑑𝑒𝑙 =
proxy model

𝑦𝑚𝑎 = Y coordinates of two endpoints of the fracture’s middle axis

135
𝑦𝑝 = Y coordinate of an arbitrary perforation point

𝑦𝑣 = Y coordinates of the four vertices of the hydraulic fracture

𝑦𝑤ℎ = Y coordinate of the adjacent well trajectory point toward heel side

𝑦𝑤𝑡 = Y coordinate of the adjacent well trajectory point toward toe side

𝑧𝑚𝑎 = Z coordinates of two endpoints of the fracture’s middle axis


𝑧𝑝 = Z coordinate of an arbitrary perforation point

𝑧𝑣 = Z coordinates of the four vertices of the hydraulic fracture



𝑧𝑤 = Z coordinate of the adjacent well trajectory point toward heel side
𝑡
𝑧𝑤 = Z coordinate of the adjacent well trajectory point toward toe side

𝛼 = 2D alpha regulator

𝛼𝑎𝑑𝑠 = Unit fractional adsorption

𝛼3𝐷 = 3D alpha regulator

𝛽 = Fracture permeability decay factor

∆𝑃 = Empirical constant pressure loss

Global objective function of the previous or initial combination of


𝜀 =
uncertain parameters

Global objective function of the proposed, disturbed combination of


𝜀∗ =
uncertain parameters

Γ = Gamma function

Ω(ℎ𝑚 ) = Regularization function of XGBoost model

𝜙 = Matrix porosity

𝜌 = Rock density

Variance of the global objective function values for the smart


𝜎2 =
sampling results

136
References
Abdle Moneim, S. S., Rabee, R., Shehata, A. M., and Aly, A. M., 2012. Modeling
Hydraulic Fractures in Finite Difference Simulators Using Amalgam LGR (Local
Grid Refinement). Paper SPE 148864, presented at the North Africa Technical
Conference and Exhibition, Cairo, Egypt, 20-22 February.
Arroyo Negrete, E., Rodiguez, J., Goryachev, S., Belova, N., Al Blooshi, A., and Basioni,
M., 2018. Automatic History Matching Theory, Implementation, and Field
Applications. Paper SPe 193018, presented at the Abu Dhabi International
Petroleum Exhibition & Conference, Abu Dhabi, UAE, 12-15 November.
Al-Shaalan, T. M., Fung, L. S. K., and Dogru, A. H., 2003. A Scalable Massively Parallel
Dual-Porosity Dual-Permeability Simulator for Fractured Reservoirs with Super-
K Permeability. Paper SPE 84371, presented at the SPE Annual Technical
Conference and Exhibition, Denver, Colorado, 5-8 October.
Bentéjac, C., Csörgő, A., and Martínez-Muñoz, G., 2019. A Comparative Analysis of
XGBoost. doi: 10.1007/s10462-020-09896-5.
Bosma, S. B. M., Hajibeygi, H., Tene, M., and Tchelepi, H. A., 2017. Multiscale Finite
Volume Method for Discrete Fracture Modeling with Unstructured Grids. Paper
SPE 182654, presented at the SPE Reservoir Simulation Conference,
Montgomery, Texas, 20-22 February.
Cavalcante Filho, J.S.A., Shakiba, M., Moinfar, A., and Sepehrnoori, K., 2015.
Implementation of a Preprocessor for Embedded Discrete Fracture Modeling in an
IMPEC Compositional Reservoir Simulator. Paper SPE 173289, presented at the
SPE Reservoir Simulation Symposium, Houston, Texas, 23-25 February.
Cipolla, C. L., Lolon, E., Erdle, J. C., and Rubin, B., 2009. Reservoir Modeling in
ShaleGas Reservoirs. Paper SPE 125530, presented at the SPE Eastern Regional
Meeting, Charleston, West Virginia, 23-25 September.
Cipolla, C. L., Weng, X., Mack, M. G., Ganguly, U., Gu, H., Kresse, O., and Cohen, C.
E., 2011. Integrating Microseismic Mapping and Complex Fracture Modeling to
Characterize Hydraulic Fracture Complexity. Paper SPE 140185, presented at the
SPE Hydraulic Fracturing Technology and Conference, the Woodlands, Texas,
24-26 January.
CMG-IMEX., 2017. IMEX User’s Guide, Computer Modeling Group Ltd.
Darcel, C., Bour, O., Davy, P., and de Dreuzy, J.R., 2003a. Connectivity Properties of
Two-Dimensional Fracture Networks with Stochastic Fractal Correlation. Water
Resources Research 39 (10): SBH1-1 – SBH1-13.

137
Darcel, C., Bour, O., Davy, P., 2003b. Stereological Analysis of Fractal Fracture
Networks. Journal of Geophysical Research 108 (9): ETG13-1 – ETG13-14.
Ding, X., 2019. Using Unstructured Grids for Modeling Complex Discrete Fracture
Network in Unconventional Reservoir Simulation. Paper SPE 195051, presented
at the SPE Middle East Oil and Gas Show and Conference, Manama, Bahrain, 18-
21 March.
Ding, Y., & Lemonnier, P., 1995. Use of Corner Point Geometry in Reservoir
Simulation. Paper SPE 29933, presented at the International Meeting on
Petroleum Engineering, Beijing, China, 14-17 November.
Du, S., Liang, B., and Yuanbo, L., 2017. Field Study: Embedded Discrete Fracture
Modeling with Artificial Intelligence in Permian Basin for Shale Formation.
Paper SPE 187202, presented at the SPE Annual Technical Conference and
Exhibition, San Antonio, Texas, 9-11 October.
Fumagalli, A., Pasquale, L., Zonca, S., and Micheletti S., 2016. An Upscaling Procedure
for Fractured Reservoirs with Embedded Grids. Water Resources Research (52)
8: 6506-6525.
Gao G., Jiang H., Hagen P., Vink J., and Wells T., 2017. A Gauss-Newton Trust-Region
Solver for Large-Scale History-Matching Problems. SPE Journal 22 (6): 1999-
2011.
Ghosh, S., 1998. Curvilinear Local Grid Refinement. Paper SPE 50633, presented at the
European Petroleum Conference, The Hague, Netherlands, 20-22 October.
Glover, K., Naser, G., and Mohammadi, H., 2015. Creep Deformation of Fracture
Surfaces Analysis in a Hydraulically Fracture Reservoir Using the Finite Element
Method. Journal of Petroleum and Gas Engineering 6 (6): 62-73.
Goodwin, N., 2015. Bridging the Gap Between Deterministic and Probabilistic
Uncertainty Quantification Using Advanced Proxy Based Methods. Paper SPE
173301, presented at the SPE Reservoir Simulation Symposium, Houston, Texas,
23-25 February.
Hassanpour, R. M., Leuangthong, O., and Deutsch, C. V., 2008. Calculation of
Permeability Tensors for Unstructured Grid Blocks. Paper PETSOC-2008-187,
presented at the Canadian International Petroleum Conference, Calgary, Alberta,
Canada, 17-19 June.
Kabir, C. S., Chien, M. C. H., and Landa, J. L., 2003. Experiences With Automated
History Matching. Paper SPE 79670, presented at the SPE Reservoir Simulation
Symposium, Houston, Texas, 3-5 February.
Kim, T. H., and Schechter, D. S., 2007. Estimation of Fracture Porosity of Naturally
Fractured Reservoirs with No Matrix Porosity Using Fractal Discrete Fracture

138
Networks. Paper SPE 110720, presented at the SPE Annual Technical Conference
and Exhibition, Anaheim, California, 11-14 November.
Kuchuk, F. J., and Biryukov, D., 2012. Transient Pressure Test Interpretation from
Continuously and Discretely Fractured Reservoirs. Paper SPE 158096, presented
at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, 8-
10 October.
Kumar, A., Close, J. C., Picone, M. M., Agustsson, H., Card, C. C., and Kjosavik, A.,
2013. A New and Practical Workflow for Large Multi-pad SAGD Simulation- A
Corner Oil Sands Case Study. Paper SPE 165511, presented at the SPE Heavy Oil
Conference-Canada, Calgary, Alberta, Canada, 11-13 June.
Lu, J., Zhu, T., and Tiab, D., 2009. Pressure Behavior of Horizontal Wells in Dual-
Porosity, Dual-Permeability Naturally Fractured Reservoirs. Paper SPE 120103,
presented at the SPE Middle East Oil and Gas Show and Conference, Manama,
Bahrain, 15-18 March.
Main, I. G., Meredith, P. G., Sammonds, P. R., and Jones, C., 1990. Influence of Fractal
Flaw Distributions on Rock Deformation in the Brittle Field. Geological Society
Special Publications 54: 71-79.
Malik, S., Harode, R., and Kunwar, A., 2020. XGBoost: A Deep Dive into Boosting
(technical report). doi: 10.13140/RG.2.2.15243.64803.
McClure, M., Picone, M., Fowler, G., Ratcliff, D., Kang, C., Medam, S., and Frantz, J.,
2020. Nuances and Frequently Asked Questions in Field-Scale Hydraulic Fracture
Modeling. Paper SPE 199726, presented at the SPE Hydraulic Fracturing
Technology and Conference, the Woodlands, Texas, 4-6 February.
Miao, J., Yu, W., Xia, Z., Zhao, W., Xu, Y., and Sepehrnoori, K., 2018. An Easy and Fast
EDFM Method for Production Simulation in Shale Reservoirs with Complex
Fracture Geometry. Paper ARMA-DFNE-18-0926, presented at the 2nd
International Discrete Fracture Network Engineering Conference, Seattle,
Washington, 20-22 June.
Moinfar, A., Varavei, A., Sepehrnoori, K., and Johns R.T., 2014. Development of an
Efficient Embedded Discrete Fracture Model for 3D Compositional Reservoir
Simulation in Fractured Reservoirs. SPE Journal 19 (2): 289-303.
Odling, N. E., 1997. Scaling and Connectivity of Joint Systems in Sandstones from
Western Norway. Journal of Structural Geology 19 (10): 1257-1271.
Okubo, P. G., and Aki, K., 1987. Fractal Geometry in the San Andreas Fault System.
Journal of Geophysical Research 92 (B1): 345-355.
Ouillon, G., and Sornette, D., 1996. Unbiased Multifractal Analysis: Application to Fault
Patterns. Geophysical Research Letters 23 (23): 3409-3412.

139
Panfili, P., and Cominelli, A., 2014. Simulation of Miscible Gas Injection in a Fractured
Carbonate Reservoir Using an Embedded Discrete Fracture Model. Paper SPE
171830, presented at the Abu Dhabi International Petroleum Exhibition and
Conference, Abu Dhabi, UAE, 10-13 November.
Panfili, P., Colin, R., Cominelli, A., Giamminonni, D., and Guerra, L., 2015. Efficient
and Effective Field Scale Simulation of Hydraulic Fractured Wells: Methodology
and Application. Paper SPE 175542, presented at the SPE Reservoir
Characterization and Simulation Conference and Exhibition, Abu Dhabi, UAE,
14-16 September.
Peaceman, D. W., 1983. Interpretation of Well-Block Pressures in Numerical Reservoir
Simulation with Nonsquare Grid Blocks and Anisotropic Permeability. SPE
Journal 23 (03): 531-543.
Ponting, D. K., 1989. Corner Point Geometry in Reservoir Simulation. Presented at the
1st European Conference on the Mathematics of Oil Recovery, Cambridge, UK,
14-16 July.
Rao, X., Cheng, L., Cao, R., Jia, P., Dong, P., and Du, X., 2019. A Modified Embedded
Discrete Fracture Model to Improve the Simulation Accuracy During Early-Time
Production of Multi-Stage Fractured Horizontal Well. Paper SPE 196263,
presented at the SPE/IATMI Asia Pacific Oil & Gas Conference and Exhibition,
Bali, Indonesia, 29-31 October.
Rubin, B., 2010. Accurate Simulation of Non Darcy Flow in Stimulated Fractured Shale
Reservoirs. Society of Petroleum Engineers. Paper SPE 132093, presented at the
SPE Western Regional Meeting, Anaheim, California, 27-29 May.
Rwechungura, R. W., Dadashpour, M., and Kleppe, J., 2011. Advanced History Matching
Techniques Reviewed. Paper 142497, presented at the SPE Middle East Oil and
Gas Show and Conference, Manama, Bahrain, 25-28 September.
Sejdinovic, D., 2015. Model complexity and generalization. Lecture, Oxford University,
Oxford, UK.
Sepehrnoori, K., Xu, Y., and Yu, W., 2020. Embedded Discrete Fracture Modeling and
Application in Reservoir Simulation. 1st Ed.; Publisher: Elsevier, Cambridge,
USA. ISBN: 978-0-12-819688-5.
Shah, S., Møyner, O., Tene, M., Lie, K.A., and Hajibeygi, H., 2016. The Multiscale
Restriction Smoothed Basis Method for Fractured Porous Media (F-MsRSB).
Journal of Computational Physics 318 (C): 36-57.
Shakiba, M., and Sepehrnoori, K., 2015. Using Embedded Discrete Fracture Model
(EDFM) and Microseismic Monitoring Data to Characterize the Complex
Hydraulic Fracture Networks. Paper SPE 175142, presented at the SPE Annual
Technical Conference and Exhibition, Houston, Texas, 28-30 September.
140
Shams, M., 2016. Reservoir Simulation Assisted History Matching: From Theory to
Design. Paper SPE 182808, presented at the SPE Kingdom of Saudi Arabia
Annual Technical Symposium and Exhibition, Dammam, Saudi Arabia, 25-28
April.
Sun, J., Schechter, D., and Huang, C.-K., 2016. Grid-Sensitivity Analysis and
Comparison Between Unstructured Perpendicular Bisector and Structured
Tartan/Local-Grid-Refinement Grids for Hydraulically Fractured Horizontal
Wells in Eagle Ford Formation with Complicated Natural Fractures. SPE Journal
21 (6): 2260-2275.
Tavassoli, Z., Carter, J. N., and King, P. R., 2004. Errors in History Matching. SPE
Journal 9 (3): 352-361.
Tijink, P., & Cottier, J., 2019. The Description and Quantification of the Truncation
Errors Produced by Local-Grid Refinement in Reservoir Simulation. SPE
Reservoir Evaluation & Engineering 22 (2): 660-672
Tripoppoom, S., Yu, W., Sepehrnoori, K., and Miao, J., 2019. Application of Assisted
History Matching Workflow to Shale Gas Well Using EDFM and Neural
Network-Markov Chain Monte Carlo Algorithm. Paper URTEC-2019-659,
presented at the SPE/AAPG/SEG Unconventional Resources Technology
Conference, Denver, Colorado, 22-24 July.
Tripoppoom, S., 2019. Assisted History Matching Workflow for Unconventional
Reservoirs. Master thesis, The University of Texas at Austin, Austin, Texas (May
2019).
Vazquez, O., Young, C., Demyanov, V., Arnold, D., Fisher, A., MacMillan, A., and
Christie, M., 2015. Produced-Water-Chemistry History Matching in the Janice
Field. SPE Reservoir Evaluation & Engineering 18 (4): 564-576.
Wang, Y., 2015. A Hybrid Dual-Continuum Discrete Fracture Modeling Approach for
Numerical Simulation of Production from Unconventional Plays. Paper SPE
178749, presented at the SPE Annual Technical Conference and Exhibition,
Houston, Texas, 28-30 September.
Wantawin, M., Yu, W., and Sepehrnoori, K., 2017. An Iterative Response-Surface
Methodology by Use of High-Degree-Polynomial Proxy Models for Integrated
History Matching and Probabilistic Forecasting Applied to Shale-Gas Reservoirs.
SPE Journal 22 (6): 2012-2031.
Wheaton, B., Haustveit, K., Deeg, W., Miskimins, J., and Barree, R., 2016. A Case Study
of Completion Effectiveness in the Eagle Ford Shale Using DAS/DTS
Observations and Hydraulic Fracture Modeling. Paper SPE 179149, presented at
the SPE Hydraulic Fracturing Technology Conference and Exhibition, The
Woodlands, Texas, 9-11 February.
141
Wu, K., and Olson, J.E., 2016. Numerical Investigation of Complex Fracture Networks in
Naturally Fractured Reservoirs. SPE Production & Operations 31 (4): 300-309.
Xie, J., Yang, C., Gupta, N., King, M. J., and Datta-Gupta, A., 2015. Integration of Shale
Gas-Production Data and Microseismic for Fracture and Reservoir Properties
With the Fast Marching Method. SPE Journal 20 (2): 347-359.
Xue, X., Yang, C., Onishi, T., King, M. J., and Datta-Gupta, A., 2019. Modeling
Hydraulically Fractured Shale Wells Using the Fast Marching Method with Local
Grid Refinements LGRs and Embedded Discrete Fracture Model EDFM. Paper
SPE 193822, presented at the SPE Reservoir Simulation Conference, Galveston,
Texas, 10-11 April.
Xu, Y., Cavalcante Filho, J. S. A., Yu, W., and Sepehrnoori, K., 2017. Discrete-Fracture
Modeling of Complex Hydraulic-Fracture Geometries in Reservoir Simulators.
SPE Reservoir Evaluation & Engineering 20 (2): 403-422.
Xu, Y., and Sepehrnoori, K., 2019. Development of an Embedded Discrete Fracture
Model for Field-Scale Reservoir Simulation with Complex Corner-Point Grids.
SPE Journal 24 (4): 1552-1575.
Yang, C., King, M. J., and Datta-Gupta, A., 2017. Rapid Simulation of Naturally
Fractured Unconventional Reservoirs with Unstructured Grids Using the Fast
Marching Method. Paper SPE 182612, presented at the SPE Reservoir Simulation
Conference, Montgomery, Texas, 20-22 February.
Yang, D., Xue, X., and Chen, J., 2018. High Resolution Hydraulic Fracture Network
Modeling Using Flexible Dual Porosity Dual Permeability Framework. Paper SPE
190096, presented at the SPE Western Regional Meeting, Garden Grove,
California, 22-26 April.
Yu, W., Tripoppoom, S., Sepehrnoori, K., and Miao, J., 2018a. An Automatic History-
Matching Workflow for Unconventional Reservoirs Coupling MCMC and Non-
Intrusive EDFM Methods. Paper SPE-191473-MS, presented at the SPE Annual
Technical Conference and Exhibition, 24-26 September, Dallas, Texas.
Yu, W., Miao, J., and Sepehrnoori, K., 2018b. A Revolutionary EDFM Method for
Modeling Dynamic Behaviors of Complex Fractures in Naturally Fractured
Reservoirs. Paper ARMA-DFNE-18-0928, presented at the 2nd International
Discrete Fracture Network Engineering Conference, Seattle, Washington, 20-22
June.
Yu, W., and Sepehrnoori, K., 2018. Shale Gas and Tight Oil Reservoir Simulation, 1st
Ed.; Publisher: Elsevier, Cambridge, USA. ISBN: 978-0-12-813868-7.
Yu, W., Hu X., Liu, M. and Wang, W., 2019. Investigation of the Effect of Natural
Fractures on Multiple Shale-Gas Well Performance Using Non-Intrusive EDFM
Technology. Energies 12 (5): 932.
142

You might also like