INTEGRATION OF CORE AND WIRELINE DATA FOR THE
Transcription
INTEGRATION OF CORE AND WIRELINE DATA FOR THE
Akademia Górniczo-Hutnicza im. Stanisława Staszica Wydział Geologii, Geofizyki i Ochrony Środowiska Katedra Geofizyki AGH University of Science and Technology Faculty of Geology, Geophysics and Environment Protection Department of Geophysics ROZPRAWA DOKTORSKA DISSERTATION KOMPLEKSOWA CHARAKTERYSTYKA WŁASNOŚCI ZBIORNIKOWYCH DLA MODELOWANIA PRZEPŁYWU MEDIÓW W ZŁOŻU GAZU ZIEMNEGO Z W ZAPADLISKU PRZEDKARPACKIM INTEGRATED RESERVOIR CHARACTERIZATION FOR FLUID FLOW MODELING OF THE Z GAS DEPOSIT AT THE CARPATHIAN FOREDEEP M.Sc. MAN HA QUANG Hanoi University of Mining and Geology, Vietnam Promotor/Supervisor: PROF. DR HAB. INŻ. JADWIGA JARZYNA KRAKÓW, 2011 ABSTRACT The purpose of this study was to use many statistical and geostatistical methods to integrate data of various scales: starting from microscopic scale (core plugs) through mesoscopic scale (logs) to megascopic (seismic) for improved reservoir characterization and reservoir modeling in order to generate a reliable geological model which was then used for dynamic simulation. This case study was performed for the Z gas reservoir depositional environment, of very complicated deltaic facies distribution which belongs to a group of the Miocene gas reservoirs in the northern-eastern part of the Carpathian Foredeep of Poland. The Hydraulic Flow Unit (HU) method was used to subdivide the reservoir into hydraulic flow units on the basis of conventional core data and conventional statistical methods. Two statistical approaches, namely Alternating Conditional Expectation (ACE) and Linear Multiple Regression (LMR) were applied to predict HU_log from core and log data. Since there is no rock type (RT) descriptive data available in the study area, unsupervised K means clustering was used for RT classification based on log data. Building static geological model both deterministic and stochastic geostatistical methods including Unsupervised Neural Network (UNN), Sequential Indicator Simulation (SIS), and Sequence Gaussian Simulation (SGS) were used to integrate 2D seismic and log data to perform static model. The innovative approach of the methodology developed and demonstrated in this study was building the pseudo 3D seismic cube from all available 2D seismic lines which were used as conditional data for hydraulic flow unit modeling. In order to verify the static geological model, history matching was performed constrained by the hydraulic flow unit distribution in the reservoir. The study has shown that even without 3D seismic and with limited well log control, the new hydraulic flow unit method can be used successfully to integrate reservoir data from different domains and from a wide range of scales. The resulting robust 3D property model can be validated by history matching. This methodology can be usefully extended to the upper part of the study area or other hydrocarbon reservoirs in Poland where dense networks of 2D seismic data are commonly available. i SUMMARY IN POLISH Geologiczna, petrofizyczna i geofizyczna charakterystyka złoża węglowodorów jest cennym źródłem informacji dla zrozumienia procesów zachodzących w formacji skalnej podczas eksploatacji. Przedstawienie wzajemnych relacji między porowatością efektywną i przepuszczalnością oraz zaileniem z uwzględnienie facjalnych zmian w trójwymiarowym ośrodku skalnym jest znakomitym ułatwieniem analizy procesów przepływu mediów w złożu. Szybki rozwój metod matematycznych i technik informatycznych oraz specjalistycznego oprogramowania stosowanego w modelowniach geofizycznych i geologicznych oraz inżynierii złożowej znacznie ułatwił parametryzację i opis procesów geologicznych niezbędnych przy poszukiwaniu i eksploatacji złóż węglowodorów. Wymusił także konieczność przedstawienia w postaci wzorów związków pomiędzy opisywanymi wielkościami. Do tego celu bardzo przydatne okazały się procedury geostatystyczne oraz sieci neuronowe pozwalające powiązać koncepcje geologiczne ze ścisłą, liczbową charakterystyką procesów. W pracy przedstawiono charakterystykę petrofizyczną i geofizyczną wybranego złoża gazu w zapadlisku przedkarpackim oraz zaproponowano podejście geostatystyczne do modelowania facji sejsmicznych oraz modelowania procesu przepływu mediów w złożu gazu w przypadku ograniczonej liczby danych. Celem pracy było połączenie danych geologiczno-geofizycznych pochodzących z różnych źródeł: punktowych wyników badań laboratoryjnych (skala mikro), jednowymiarowych danych geofizyki otworowej (skala mili) oraz sejsmiki 2D (skala makro) dla udokładnienia opisu złoża gazu oraz dla utworzenia poprawnego statycznego, trójwymiarowego modelu złoża w celu wykonania modelowania przepływu. Końcowym efektem pracy było sprawdzenie wyników symulacji przepływów z danymi otworowymi uzyskanymi podczas produkcji. Wykorzystane procedury wymagały spełnienia następujących warunków: dostępność danych i oprogramowania oraz efektywność finansowa, możliwość obliczenia wskaźników Flow Zone Index, FZI, oraz podziału obszaru złoża na jednostki hydrauliczne, HU, dysponowanie danymi sejsmicznymi 2D (jeśli byłyby dostępne dane sejsmiczne 3D wyniki będą dokładniejsze), ii dostęp do danych umożliwiających poprawne skalowanie modeli statycznych i dynamicznych od poziomu mikro do skali makro, sprawdzenie poprawności wyników na podstawie historycznych danych eksploatacyjnych. Kolejne kroki wykonane w pracy są przedstawione na diagramie (Fig. 1.1). Diagram ten może być podstawą organizacji podobnych prac przy spełnieniu warunków wymienionych powyżej. Praca doktorska składa się z 7. rozdziałów. Pierwszy rozdział stanowi wprowadzenie do tematu konstrukcji statycznego i dynamicznego modelu ośrodka skalnego w warunkach ograniczonej dostępności danych. Zawiera również krótki opis treści pozostałych rozdziałów. W rozdziale 2. przedstawiono zarys budowy geologicznej złoża gazu ziemnego Z w zapadlisku przedkarpackim. Złoże należy do typowych wielowarstwowych nagromadzeń gazu ziemnego w utworach sarmatu w północno-wschodniej części zapadliska. Do prac modelowych wybrano fragment cienkowarstwowej formacji piaskowcowo-mułowcowoiłowcowej, powstały w warunkach sedymentacji deltowej. Zwrócono uwagę na skomplikowany charakter utworów zbiornikowych i pokazano, że cechy obserwowane na anomaliach geofizyki otworowej oraz widoczne na przekrojach sejsmicznych (amplitud oraz innych atrybutów) wskazują na określone warunki sedymentacji. W rozdziale 3. przedstawiono koncepcję konstrukcji jednostek o jednakowych zdolnościach do przepływu (Hydraulic Units, HU) na podstawie parametru Flow Zone Index, FZI, wyznaczonego tylko na podstawie porowatości i przepuszczalności. FZI oraz HU zostały wyznaczone na podstawie wyników badań laboratoryjnych dostępnych w 10. otworach na złożu gazu Z (560 analiz). W rozdziale 4. omówiono wyniki połączenia danych laboratoryjnych i profilowań geofizyki otworowej w celu utworzenia jednowymiarowych modeli zmian HU w profilach otworów. W tym celu zastosowano regresję wielowymiarową, Linear Multiple Regression, LMR, oraz metodę Alternating Conditional Expectation, ACE. Nie dysponowano geologicznym opisem facji w interwałach, z których pobrano rdzenie, zatem zastosowano statystyczną metodę K mean do wydzielenia charakterystycznych warstw w badanej formacji, czyli wyznaczenia Rock Types. Uzyskane wyniki podziału mioceńskiej formacji cienkowarstwowej na 6 typów iii skalnych (litofacji), RT, okazały się spójne z wynikami uzyskanymi wcześniej na drodze wykorzystania ACE do przeniesienia rozkładu HU na pełne profile geologiczne w badanych otworach. W rozdziale 5 opisano sposób wyznaczenia statycznego, trójwymiarowego modelu ośrodka skalnego na bazie wcześniejszych wyników oraz 25 profili sejsmicznych 2D. Do połączenia wyników profilowań geofizyki otworowej i danych sejsmicznych wykorzystano metody deterministyczne, Unsupervised Neural Network, UNN, oraz geostatystyczne, Sequential Indicator Simulation, SIS, i Sequence Gaussian Simulation, SGS. Zastosowano innowacyjne przejście od danych sejsmicznych 2D do modelu trójwymiarowego (Pseudo 3D Seismic Cube). Następnie wykorzystano trójwymiarowy rozkład facji sejsmicznych w połączeniu z litofacjami, RT, wyznaczonymi wcześniej na podstawie geofizyki otworowej do utworzenia trójwymiarowego modelu jednostek przepływu, HU. W celu udokładnienia informacji pozyskiwanej na podstawie danych o zróżnicowanej pionowej rozdzielczości podczas tworzenia trójwymiarowego modelu statycznego wprowadzono dodatkowo więzy oparte na wielkości stosunku Net to Gross. Ostatecznym sprawdzianem poprawności przygotowanego modelu było porównanie wartości współczynnika przepuszczalności wyznaczonego w jednostkach jednakowego przepływu w modelu, K_HU i wartości przepuszczalności wyznaczonych na podstawie danych laboratoryjnych i geofizyki otworowej, K_log. W rozdziale 6 przeprowadzono modelowanie przepływu gazu na podstawie trójwymiarowego modelu opracowanego w rozdziale 5. Jako weryfikację poprawności modelowania zastosowano historyczne dane produkcyjne w 3. otworach na badanym obszarze złoża Z. Zastosowanie jednostek jednakowego przepływu okazało się skuteczne pod wieloma względami - i: uzyskano bardzo dobre korelacje między porowatością i przepuszczalnością w jednostkach HU, ii: dzięki użyciu HU możliwe było zredukowanie obliczeń w programie Eclipse do jedynie tych jednostek, które miały wysoki współczynnik FZI, czyli dużą porowatość i przepuszczalność oraz wysoki stosunek NTG, iii: funkcja kompakcji była dobrze dopasowana do modelu porowatości w każdej jednostce HU. W rozdziale 7 podsumowano wykonane prace oraz podkreślono innowacyjne aspekty modelowania statycznego – tworzenia trójwymiarowego ośrodka geologicznego z podziałem na jednostki HU oraz dynamicznego dla obliczenia przepływu mediów. Zwrócono uwagę na celowość proponowanego połączenia danych laboratoryjnych i geofizyki otworowej poprzez iv wyliczanie parametrów FZI oraz HU. Włączenie do obliczeń jednostek litologicznych (Rock Types) wyznaczonych na podstawie geofizyki otworowej stanowi dodatkowy element skalowania danych i przejścia od wartości punktowych do ciągłej informacji wzdłuż osi otworu. Włączenie danych sejsmicznych 2D, dzięki wykorzystaniu metod geostatystycznych i deterministycznych do konstrukcji facji sejsmicznych, pozwoliło na utworzenie statycznego modelu trójwymiarowego (Pseudo 3D Seismic Cube). Obliczenie porowatości i przepuszczalności w trójwymiarowym modelu statycznym zostało dodatkowo uzupełnione informacją na temat Net to Gross. Wyniki modelowania przepływów zostały zweryfikowane na podstawie danych historycznych. Ocena niepewności uzyskanych wyników jest najtrudniejszym do wykonania elementem przedstawionego schematu postępowania, zalecanego do wykorzystania w obszarach, gdzie do dyspozycji pozostaje ograniczona liczba danych geofizycznych i geologicznych. Stosowanie metod statystycznych i geostatystycznych powoduje konieczność wykorzystania dużej ilości danych laboratoryjnych dla uzyskania wiarogodnych wartości średnich i odchyleń standardowych porowatości i przepuszczalności. Stosowanie procedur statystycznych w modelowniach skutkuje brakiem powtarzalności przy powtarzaniu obliczeń, a nie uwzględnienie szczegółowych informacji o elementach tektonicznych w badanym obszarze z powodu niepełnych danych geologicznych powoduje ograniczenie wiarogodności ostatecznych wyników modelowani. Ograniczenie wiarogodności predykcji zachowania złoża w przyszłości wynika z braku historycznych danych z długiego okresu produkcji złoża. Zatem, ograniczona wiarogodność uzyskanych wyników wynika z ograniczeń zastosowanych metod i dostępu do danych. v ACKNOWLEDGMENTS I would like to take this opportunity to thank all the family and friends who have helped and inspired me during my doctoral study! I would like to express my sincere thanks to Prof. dr Jadwiga Jarzyna, who patiently supervised the progress of my work and it is for her patient supervision, useful advice and discussions that led to the completion of this project. Many thanks also go to Prof. Le Hai An for his helpful counsels and arguments. I owe my deepest gratitude to the Rector of AGH University of Science and Technology, Cracow, Poland for granting me the scholarship and also Hanoi University of Mining and Geology for their support to my accomplishing this study. I am grateful to Prof. Jacek Matyszkiewicz, the Dean of the Faculty of Geology Geophysics and Environmental Protection AGH UST, Cracow, Poland. My special thanks to the Polish Oil and Gas Company Ltd., Warsaw, Poland for the permission to use the data. Petrel®, Eclipse®, Interactive Petrophysics® were used thanks to the university’s grant donation by Schlumberger to AGH UST. STATISTICA software was used thanks to AGH UST, Cracow, Poland license. I am also grateful to dr Jerzy Ziętek and his family, Ms. Maria Cicha and all of staff in Geophysics Department for helping me to adopt myself into Polish life style and other official matter. Without their helps this project would not have been completed. Thanks also go to Mr. Graham Dryden from Subsurface Consultants & Associates, Houston, USA for help with the English corrections and also helpful discussions. A special thanks to many of the students at AGH University. In particular, thanks to Wojciech Machowski, Michał Michna, Paulina Krakowska for the friendship we have developed. Also thank to all my Vietnamese friends in Cracow for their invaluable support and consideration during my stay here. I would like to thank my dear parents, parents-in-law and my brothers, sisters for their encouragement and support during the last 4 years when I was absent from home. Finally, I am deeply indebted to my dear wife Thu and daughter Minh Hằng for their love and their encouragement to my course-choosing and study. vi DEDICATION This thesis is dedicated to my Grandparents, my Parents, my Parents-in-law, my wife Thu and my daughter Minh Hằng vii TABLE OF CONTENTS ABSTRACT…………………………………………………………………………………….i SUMMARY IN POLISH……………………………………………………………………....ii ACKNOWLEDGMENTS……………………………………………………………….........vi DEDICATION…………………………………………………………………………...….. vii Chapter 1……………………………………………………………………………………...1 INTRODUCTION Chapter 2 GEOLOGICAL SETTING 2.1 Introduction to the geology and tectonic framework of the study area…………………….5 2.2 Lithology and sedimentological environment…………………………….………………10 2.3 Core Analysis……………………………………………………………………………..13 2.4 Relationship between reservoir parameters and sedimentary environment………………17 2.5 Organization of succession, its division and correlation diagram……………...………...20 2.6 Conclusions……………………………………………………………………….………28 Chapter 3 HYDRAULIC FLOW UNIT 3.1 Concept of Hydraulic Flow Unit……………………………….…………………………29 3.2 Hydraulic Flow Unit classification technique…………………………………………….35 3.2.1 Histogram……………………………………………………………….….…..36 3.2.2 Probability plot………………………………………………………….……...38 3.2.3 Ward's algorithm approach…………………………………………….…….....39 3.3 Critical review on determining HU in the Lower Miocene reservoir of Z gas field……..40 3.3.1 Preliminary outcomes…………………………………..………………...…….40 3.3.2 Global Hydraulic Elements………………….…………………………..……..41 3.4 Conclusions……………………………………………………………………………….46 viii Chapter 4 CORE – LOG INTEGRATION 4.1 Core - log depth matching…………………………………………...……………………47 4.2 Hydraulic flow unit prediction…………………………………………………….……...50 4.2.1 Methodology……………………………………………………………..…......50 4.2.2 Flow Zone Indicator prediction………………………………………….……..52 4.2.2.1 Linear Multiple Regression (LMR)…………………………….…….52 4.2.2.2 Alternating Conditional Expectations algorithm (ACE)…….….…….56 4.2.2.3 Comparison of LMR and ACE algorithm…………………….………59 4.2.3 Hydraulic Flow Unit Prediction (HU_log) from FZI_pre_ACE……..….……..60 4.2.4. Validation of results………………………………………..………………..…63 4.3 Rock types classification………………………………………………………..…….…..66 4.3.1 K means clustering background ………………………………………..……….66 4.3.2 Applying K means for the data group 3 (G3: Z-76, Z-81, Z-82)……….………69 4.4 Relationship between Hydraulic Flow Unit (HU) and Rock Types (RT)……...…………69 4.5 Conclusions…………………………………………………………………………….…77 Chapter 5 STATIC MODELING 5.1 Reservoir modeling overview………………………………………………….…………79 5.1.1 Reservoir modeling workflow……………………………………….………...…79 5.1.2 Deltaic facies and spatial relationship………………………………..…….…….80 5.1.3 Geostatistical methods overview…………….…………………..….………...….83 5.2 Structure modeling……………………………………………………………….……….86 5.2.1 Miss-tie correction for 2D seismic survey…………………………..……….…...87 5.2.2 Horizons picking and 3D grid…………………………………….….………......87 ix 5.3 Rock Type and Hydraulic Flow Unit Modeling…………….……………..……………..91 5.3.1 Convert 2D seismic to pseudo 3D seismic……………………….…….……...…93 5.3.2 Rock Type modeling constrained by seismic facies model…………………..….96 5.3.2.1 Seismic facies extraction volume…………………………………….96 5.3.2.2 Seismic facies classification………………………….……………...100 5.3.2.3 Rock Type modeling………………………………..………..……...104 5.3.3 Hydraulic Flow Unit modeling constrained by Rock Type model..….……...….107 5.4 Properties modeling constrained by HU model………………………………..………..112 5.4.1 Porosity and permeability modeling……………………………….…………....112 5.4.2 Water saturation and Net to Gross modeling………………………………...….116 5.5 Conclusions……………………………………………………………………….……..120 Chapter 6 HISTORY MATCHING UNDER HYDRAULIC FLOW UNIT CONTROL 6.1. History matching under Hydraulic Flow Unit control………………………………….123 6.1.1 History Matching Overview……………………………………..……….……..123 6.1.2 History matching under Hydraulic Flow Unit control………………….…..…..125 6.2. Upscaling...…………………….……………………………………………...…….….129 6.3. Reservoir initial condition………………………………………………………………134 6.4 History matching and discussions……………………………………………………….137 6.4.1 Results…………………………………………………………………..………137 6.4.2 Discussion……………………………………………………………..………..138 6.5 Conclusions……………………………………………………………………………...143 Chapter 7 CONCLUSIONS AND RECOMMENDATIONS 7.1 Conclusions…………………………………………………………………………...…144 7.2 Recommendations……………………………………………………………………….147 x REFERENCES.......................................................................................................................151 LIST OF FIGURES AND TABLES………………………………………………….....…..158 LIST OF ABBREVIATIONS…………………………………………….…………...…….168 APPENDIX A…………………………………………………………………….…………169 xi Chapter 1 INTRODUCTION In the upstream petroleum industry, reservoir characterization and description plays an important role giving geologists and geophysicist a more accurate and detailed understanding of the subsurface reservoirs. Understanding key reservoir properties such as porosity, permeability, their relationship as well as their spatial distribution by having precise and realistic reservoir model, assists petroleum engineers in improving the characterization of the reservoir, and to enhance the development and performance of the reservoir throughout its life. In recent years, with the robust development of the mathematics and computer science such as modeling, geostatistics, and neural networks, geological ideas have become easier to realize and be verified, especially with the advanced tools available through specialized software in the petroleum industry like Schlumberger Petrel and Eclipse. The need to bring the latest developments to the petroleum industry has led to one of the major motivations of this study. This thesis looks at various aspects of reservoir characterization and modeling where the geological and geophysical data available is rather limited. The aim of this thesis is to integrate G&G data from diverse disciplines; from core plugs through wireline logs to seismic. To this aim we develop suitable methodologies for improved reservoir characterization and reservoir modeling with rather limited data (mainly 2D seismic available) in order to generate a reliable geological model which was then used for dynamic simulation. On top of that, this study has successfully verified a recently emerged Hydraulic Flow Unit approach in various scales from micro scale of the core plug to mega scale of the seismic within a constraint of dynamic data through history matching processes. The methodologies developed in this study were designed to be: reliabile and cost-effective, applicable to rocktyping by mean of hydraulic flow unit in reservoir characterization, applicable to dealing with limited data availability, especially where no 3D seismic data available, 1 applicable and reliable in upscaling from micro to mega scale in the framework of geological reservoir modeling constrained by dynamic data. The workflow, which while by no means universally applicable, illustrates a practical approach to handling sparse data in a deltaic environment gas reservoir. The workflow was used throughout the course of this study, and is summarized in figure 1.1. Core, Logs 2D Seismic Production Qg, BHP, PVT,… FZI_core FZI_log K-mean Clustering Pseudo 3D Seismic HFU_core HFU_log RT_log 3D Seismic Structure Facies 3D Rock Type Modeling 3D HFU Modeling Dynamic Modeling History Matching 3D Properties Modeling (PHI, K) Fig. 1.1 The main workflow in this study for integration of static and dynamic models The rest of the thesis consists of 6 further chapters. Chapter 2 reviews the geological setting, including geology, tectonic framework, lithology and sedimentological environment of the Z gas deposit belonging to a group of Miocene gas reservoirs in the northern part of the Carpathian Foredeep of. This chapter also addresses the relationship between reservoir parameters and deltaic sedimentary environment due to the complexity of the structure of the Miocene (Sarmatian) succession in study area. 2 Chapter 3 introduces the concept of the hydraulic flow unit (HU) and applies this approach to determine HU throughout the field based on conventional core data and conventional statistical methods. The results show that using all available 570 core plugs from the study area resulted in a reliable model of 6HUs. These were consequently used for the reservoir simulation. Chapter 4 discusses the application of two statistical approaches, namely Alternating Conditional Expectation (ACE) and Linear Multiple Regression (LMR), which were applied to predict hydraulic flow units (HU_log) from cores and logs. Since there is no rock type (RT) descriptive data available in the study area, unsupervised K means clustering was used for facies classification based on log data. The results demonstrate that statistical methods are useful, flexible, and effective, and that they delivered acceptable results. Unsupervised K means clustering is effective for classifying six rock types. The ACE method is good for predicting FZI from core and log. Chapter 5 proposes an appropriate approach in building static geological modeling for the reservoir in study area. Both deterministic and stochastic geostatistical methods including Unsupervised Neural Network (UNN), Sequential Indicator Simulation (SIS), and Sequence Gaussian Simulation (SGS) were used to integrate 2D seismic and log data to perform static modeling. The innovative approach of the methodology developed and demonstrated in this chapter was building the pseudo 3D seismic cube from all available 2D seismic survey for use as conditional data for hydraulic flow unit modeling. The derived results show clearly that there is a good correlation between permeability log (K_log) and 3D permeability model (K_HU) was and also confirm the advantage of applying the HU approach in this study. Due to lack of SCAL data for reservoir rock parameters, the 3D NTG model was calculated directly from 3D HU and it again showed the advantages of HU methods. In Chapter 6, in order to verify the static geological model built in chapter 5, history matching was performed constrained by the hydraulic flow unit distribution in the reservoir. Some recent accomplishments in jointly integrating static and dynamic information into a reservoir flow simulation model and history matching were reviewed. History matching was tested by manually adjusting a few reservoir model parameters through a trial-and-error procedure. Manual history matching was performed by running the simulator over the static 3D reservoir model for the historical production period, and then the results were compared 3 with known field performance. The history matching results of three cases from 3 wells (Z74, Z75 and Z76) showed good results for gas flow rate for a few years production. As the study evolved, it became abundantly clear that when comparing the results of the traditional method (two stages) to the HU method (four stages), there are three distinct advantages of the HU approach: (i) at each HU we have very good correlation between porosity and permeability that is good for classification of cells in fluid flow for Eclipse simulation, (ii) from 6HUs distribution we can easy control NTG model for Eclipse simulation cutting off HU1 with lower PHI and K while reducing number of cells in the 3D grid to increase CPU processing speed, and (iii) the rock compaction function can respond to the porosity model (PHI) and each hydraulic flow unit. Chapter 7 summaries all of the innovative aspects of the thesis, the conclusions from various studies in the thesis and gives some recommendations and suggestions for future work that might benefit the reservoir in the study area. 4 Chapter 2 GEOLOGICAL SETTING 2.1 Introduction to the geology and tectonic framework of the study area The multi-pay gas field, Z-L, is located in the north-eastern part of the Polish Carpathian Foredeep (Fig. 2.1). The Carpathian Foredeep is a peripheral molasse sedimentary basin which has been being overthrust in a northerly direction. The most hydrocarbon prospective sedimentary section is Miocene age (Karnkowski, 1999). The Miocene sediments can be up to 3500 m thick in the south part of the Carpathian Foredeep and are thinner in the northern part (Fig. 2.2). Fig. 2.1 General overview of the Carpathians and Carpathian Foredeep (Oszczypko, et al., 2006) Two main geological components are important in the Carpathian Foredeep: first – the structural Precambrian - Palaezoic unit – constituting the base for the Neogene sediments, whilst in the Z-L region the unit is comprised of Proterozoic rock. The second structural unit is formed by the autochtonous Miocene formation including Lower Badenian, Middle Badenian and Upper Badenian and Lower Sarmatian and Quaternary sediments. 5 Fig. 2.2 Schematic map showing the submarine deposits diversity in the eastern part of the Carpathian Foredeep (Myśliwiec, 2004) The Precambrian basement is made of fyllitic Table 2.1 Top of the Precambrian in shales. The rocks are ductile, with common study area cracks and fissures, most of which are filled with Well Depth [m] clay minerals or calcite. Also present are thin Z2 1283 laminae of hard sandstone. In the study area the Z24 1223 top of the Precambrian was observed in four Z72 1237 wells as follows in Table 2.1. Z75 1097 In the Z72 well, the Precambrian formation consists of mudstones of low or moderate cohesion, fissured and cracked, and moderately to poorly lithified. These beds dip at 45o. In the Z74 well, the Precambrian rocks comprise mudstones and sandy-mudstones having a 6 wackestone texture. In the D2 well the Precambrian rocks comprise massive mudstones and claystones. The mudstones sporadically include thin layers of fine-grained sandstone lenses, and more rarely continuous laminae. Sandstones textures include subarkose or arkose arenites, and rarely wackestones. The dense network of fissures, fractures and cracks range from closed to partially-open and are filled with quartz, chlorite and kaolinite. In the Z77 well the Precambrian rocks comprise claystones and mudstones. The Precambrian formation was also found in the Ch.D. 1, 2 and 3 wells and in the G.D. 1 and 3 wells. Generally, the Precambrian succession is covered by the autochtonous Miocene sediments, but in the southern part of the central element of the Z gas field, in the cores in the Z76 well, weathered fragile mudstones, claystones and fine grained quartzite sandstones with carbonateclay cement, and calcite cement are found. Organic matter is also present. There is no paleontological documentation, although by way of analogy with other formations in the Carpathian Foredeep it was assumed that the sediments are of Paleogene age. In the most wells in the Carpathian Foredeep the basement is covered by the Badenian and Sarmatian sediments. To the south of the Z-L gas field there is a zone without the Lower and Middle Badenian sediments, the so called Rzeszów Iceland. Its northern border coincides with the southern edge of the central element of the Z-L gas field. The Miocene formations of the Carpathian Foredeep have many years been divided into the Lower Badenian sub-evaporitic series, Middle Badenian evaporites, and the Upper Badenian and Sarmatian supra-evaporitic formations. Newer and more precise stratigraphic studies slightly modify the age determined for the above-mentioned principal series of sediments occurring within the Z-L bed. The former divisions are still used, however, because evaporites are the most important reflecting horizon identified in seismic sections in the foredeep (Myśliwiec, 2004, 2006a). The gas field Z-L consists of three components: the eastern component, central component and western component (Fig. 2.3). To the south there is another gas field, G.D., and to the east there is the Ch.D. gas field. The studied gas field Z-L and the surrounding fields are formed on the uplifts of the older base of the Miocene succession (Myśliwiec, 2004, 2006a). 7 Fig. 2.3 Location of Z gas-field with cored and logged wells; tectonic elements marked in red (after Myśliwiec, 2006b and Myśliwiec et al., 2004) Lower Badenian This formation comprises grey-green shales deposited in the outer shelf environment and green glauconite sandstones from the shallow shelf environment. It has not been thoroughly explored, because most of the wells were drilled only to the bottom of the Sarmatian. Sandstones are average in terms of sorting and hardness, but they are fractured. However they are not important as a reservoir because they are generally of insufficient thickness (Dziadzio 2000). In the Z74 well, the Lower Badenian (Baranów beds) were penetrated from 1199 to 1204 m, in the Z72 well: 1228 to 1237 m, and in the Z77 well: 1215 to 1220 m. According to core descriptions from the Z72 well, the Baranów beds comprise clay shales and mediumgrained sandstones, medium-sorted, with only slightly rounded grains. Middle Badenian In the eastern and southern part of the central element of the Z gas field, the Middle Badenian comprises anhydrite with clays, salts and calcites, including also epigenetic calcite (Mysliwiec 2004). The anhydrite is about 10 – 25 m thick. In the west central element, the Middle Badenian comprises hard and massive limestones, with mudstone intercalations. The rocks are 10-15 m thick and are deformed. This evaporite series was formed during the shallowing of marine environment in which the reservoir rocks were being deposited, and this resulted in its partial isolation (Oszczypko, 1999). The evaporites in the eastern part of the foredeep are called the Krzyżanowice formations (Table 2.2) Their age, determined on the basis of nanoplankton, was established as the top of the Upper Badenian (Olszewska, 1999), thus 8 these formations are not as old as previously thought (Middle Badenian). The thickness of the evaporites fluctuates around ten meters or so. Upper Badenian - Lower Sarmatian The sedimentation associated with the Upper Badenian and Sarmatian subsidence formed thick shaly-sandy layers, overlaying evaporites or, whenever the latter are absent, lying directly on the substrate rocks. There are many divisions concerning these formations, but they are most often called the Machów formations (Table 2.2). In the area of the Z gas field, these sediments consist of hard shales rarely laminated with fine-grained sandstones and mudstones. The sediments dip at from 5 to 30o and their thickness increases to the south. Separating the formations of the Upper Badenian from the Lower Sarmatian is difficult and usually made on the basis of increased sand content in the Sarmatian formations. In the places where the Middle Badenian is absent, the Upper Badenian is deposited directly on the Precambrian basement. The thickness of the Upper Badenian sediments in the central element of the Z gas field varies from 10-40 m (Z82). The greatest thickness is where the Middle Badenian is absent. Sarmatian The most important and thickest deposits of the autochtonous Miocene are the Sarmatian sandstones and shales, which are about 1100 m thick in the area of the Z gas field. The Sarmatian formations are most prospective in the study area because of significant accumulations of bitumens in the Z-L, G.D., and Ch.D. gas fields. The origin of the geobodies, was based on sedimentological investigation of cores and dipmeter interpretation (Documentation of wells: Z75, 79 and 84). Sedimentary environment studies of the Sarmatian, as well as the studies of the geometry and the lithology and facies of the geobodies enable us to recognize relationships between the lithofacies and gas reservoir rocks of Miocene age. These studies also show that the Miocene sediments form stacks of layers, which vary from very thin (several centimeters thick) up to 10 meters thick. These studies also indicate that these particular sandy bodies are comparable to those of other Miocene gas deposits. These studies have also contributed to our knowledge about horizontal and vertical variation in lithology and reservoir properties, and allow us to predict which sand bodies are likely to be the best gas prone reservoir rocks. 9 Table 2.2 Stratigraphy of the Miocene formation (Rögl, 1996; Martini, 1971) 2.2 Lithology and sedimentological environment A characteristic feature of the Miocene deposits in the Carpathian Foredeep is the considerable diversification of facies. Facies changes related to the differentiation of a sedimentation environment are distinctly visible in the Sarmatian rocks. The depositional environment strongly influences the reservoir parameters of rocks (Dziadzio et al., 1997). In terms of their lithology, petrography and particularly facies characteristics, the Miocene reservoir rocks are extremely diversified. It occurs so often that each region, bed or even a well or single gas-bearing horizon, have their own specific characteristics. For this reason it is very difficult to systematize the data relevant to them. However, it is possible to classify them into several such types which, although of different age, lithology and origin, are - at the same time – of major economic importance (Myśliwiec, 2004; Myśliwiec, 2006a; Matyasik et al., 2007). In the Carpathian Foredeep, four main litho-facies have been identified. These are, starting from the deepest sediments: 1. A complex of turbidite sediments of the lowest Sarmatian, 450-500 m thick. It comprises sandstones of submarine fans and heteroliths of sandstone-mudstone-claystone from the basin plain. This sequence is organized as about 50 m cyclothems. A characteristic feature is the presence of a thin or very thin laminae. Some of these laminae are over one meter thick. Sandstone sequences do not cover very large areas and on their edges, they grade into mudstones. Heteroliths are good reservoir rocks; their porosity exceeds 20% and 10 permeability reaches 500 mD (Myśliwiec, 2006a; Dziadzio, 2000). These rocks appear to have been developed on the slopes of vast prograding accumulation levees typical of the distal parts of submarine complex. This type of prograding or accreting submarine fan complex, with thin laminae suggests that the environment of deposition was distal from the source, where hydrodynamic energy was almost exhausted. 2. A transitional sediment complex about 220 m thick. It comprises sandstone laminae fining upwards into sandstone-mudstones. In the upper part of the complex, sandstones prevail over fine-grained sediments. The complex is a result of progradation of a deltaic depositional system and is transitional between turbidites and deltaic sediments. 3. A complex of deltaic sediments is the most diverse in terms of a sedimentation environment and facies. Its most important feature is a vertical cycling. The 300 m thick complex comprises several parasequences (cyclothems) each 20-30 m thick. The parasequences have a complicated structure, but in each, one can observe sets of laminae of coarse grained sandstones with the thickness of each set of laminae increasing towards the top of the parasequence. Parasequences start with sediments of prograding accumulation levees (piles or bars). Inside of these sediments one can select parts of various sedimentation environments: delta slope, estuary levees (piles/bars) and deltaic plain. Sandstones in parasequences occur in the form of several meters thick packets bordered by underlying heteroliths of similar thickness. In the region of the Z gas field, not all cyclothems underwent typical or complete development, particularly in the middle and upper part of the deltaic sediment complex. Deltaic sediments are built in a zone where the river and the sea meet.They include a considerable amount of organic matter. This fact together with prograding deltas, is important from a hydrocarbon prospecting point of view because: 1) organic matter can be converted into hydrocarbons, which means the Sarmatian deltaic deposits are mature source rock for hydrocarbons and 2) the maximum flooding surfaces (MFS) indicated by the muddy-shaly upper limits of the parasequences form hydrocarbon seals for the reservoirs below. In fact deltaic sediments, especially those built in estuary ostiary levees (piles/bars), are good reservoir rocks. The sandstones have porosities ranging from 15-32 % and permeabilities of around 900 mD (Myśliwiec, 2006a, Śmist, 2003). 11 4. A complex of shallow shelf, littoral sediments represents the final stage of sedimentation in the Carpathian Foredeep. Deposits are poorly sorted and include fine grained lithofacies and muddy-shaly litho-facies. There is great vertical and horizontal variation in facies, which makes it difficult to recognize, classify and correlate them. Despite these litho-facial complexities, these sediments include gas accumulations in the shallow horizon (nr I) of Z gas field, representing sediments of a near shore coastal environment and open shelf and in the horizon nr II, as sediments from sandy barriers delimiting lagoons (Myśliwiec, 2004). The sandstones of the Sarmatian deltaic formations, and particularly the sandstones occurring in heterolithic or clayey formations, are characterized by low textural and mineral maturity. This first feature means that there is a relatively high content of a clayey-shaly matrix in the cement, and low rounded detritus. The second feature shows as a varied petrographic composition with minerals of platy crystal habit, plagioclases and various rock fragments. The deltaic sandstones of the Sarmatian usually contain large quantities of dispersed calcium carbonate, thus they are sometimes called marly sandstones. The total calcium carbonate content is increased also by the mud derived from from erosion of limestone rock, which, together with shale, forms cement. The grain matrix of sandstones is typically fine- to medium-grained with an admixture of silty-shaly material. Using the typical classification of lithoclasts, the deltaic sandstones include lithic and sublithic wackes, quartz-mica wackes, and – when better-sorted – sublithic arenites. The dominant mineral components include quartz and fragments of other rocks (limestones, dolomites, radiolarites, quartzites, quartz-muscovite shales, inter alia), as well as plagioclases, micas, calcium feldspars, glauconite, and carbonized plant detritus. Also present are bioclasts – foraminifera, fragments of bryozoans, and bivalves. The cement is made of shaly material, limestone mud, and terrigenous silt. Incidentally, crystalline calcite and dolomitic cements occur, filling single pores. On the basis of geological and geophysical studies it may be stated that the trap for natural gas in the Z gas deposit, which consists of a number of gas-bearing horizons, is of a structuralstratigraphic type. It is an anticline formed above an uplift of substrate rocks. In some horizons, the factors creating the trap also include lithological changes and more precisely, changes in facies together with local variations in porosity and permeability. 12 In the Z gas deposit several types of gas-water contact were identified depending on the facies and lithology of the reservoir rocks (Geological Documentation of the Z-L gas field, 2007). In terms of different positions of separation surfaces between parts saturated with natural gas and those saturated with formation water, in the horizons where edge water occurs, the deposit is partly a thinly bedded sheet-deposit, whereas it is partly a more massive deposit in those horizons where bottom water occurs. In the sandstone horizons, and particularly these occurring in the upper parts of deltaic prograding accumulation levees, sealed by shaly sediments of the basin bottom, or those in sand-filled canals of the upper undersea fans, seismic profiling shows a visible gas-water boundary. The reservoir drive mechanism in these types of sediments is either edge drive in the case of thin laminated reservoirs, or bottom water in the case of thick and massive sandstone reservoirs When the gas is accumulated in heterolithic sediments of the delta plain or in deposits in the bays of a shallow shelf/near-shore coastal environment zone each thin sandstone or shale layer may have has its own individual gas-water contact. 2.3 Core Analysis Results of porosity and permeability obtained by laboratory measurements on cores are presented on the examples from 3 wells: Z75, Z84 and Z78. Core samples in other wells from the Z gas field provided similar results. Porosity distribution was measured using an Auto Pore 9320 mercury porosimeter. A scanning electronic microscope, SEM, and Roentgen microanalyser were used to examine the pore structure. Information was also obtained on cement type and distribution. Porosity in Z75 ranged commonly from 23 to 28 percent with rare samples having a range of 16-20 percent. The grain diameters fall within the 0.032 – 2 mm range, with the dominant fraction in the range of 0.033 – 0.125 mm (Fig 2.4). The average diameter of a pore throats is high and ranges from 0,2 to 2 μm. Over 50% of throat diameters exceed 1 μm. Polished sections impregnated with blue resin and images from the scanning microscope illustrate the nature of porous space perfectly (Fig 2.5). The porosity of this sample is 20.85%. The grain material is loosely packed and the proportion of diagenetic cement is low. Envelopes of quartzite cement and aggregations of clay materials are seen on scanning microscope images along with single quartzite crystals (Fig 2.6). Porosity is distinctly macropore in character. Pores with a diameter greater than 1 μm total some 80 to 95 percent, which means they 13 possess good filtration abilities. The SEM showed that there is not much cement among the mineral components. It was also shown that porosity is inter granular or inter crystalline and pores are in communication, not isolated, suggesting high permeability. There is also a minor secondary porosity related to pores and micropores in grains of feldspar and plagioclases and in the lithoclasts. In some cases porosity exceeded 30 percent with permeability higher than 1 Darcy. These excellent reservoir characteristics were observed in well sorted sediment deposited in a high energy sedimentary environment, typically in the channel axes at the head of the delta or in distributary channels. Number of data (%) 45 40 35 30 25 20 Fig.2.4 The result of grain size 15 analysis of powdered sandstone 10 taken at 677m, the depth of the 5 0 0.031 Z75 well 0.063 0.125 0.25 0.5 1 2 Grain size (mm) Fig. 2.5 Photomicrographs of a polished section impregnated with blue resin, from a sample obtained at a depth of 855.20m in well Z75; crossed nicoles (left – magnification 14X, right – 120X) (Documentation of wells: Z75, 79 and 84) 14 Fig. 2.6 Scanning microscope photomicrographs of a polished section from a sample obtained at a depth of 855.20m in well Z75; magnification 600X (Documentation of wells: Z75, 79 and 84) Samples in the Z84 well were highly diversified. Many of them, taken from the shallower part of the borehole, were too soft and shaly to make measurements. Only samples from the depth section of 720-729 m enabled regular measurements. Porosity in that section ranged from 24 to 25 percent and permeability in some samples was higher than 1 Darcy. Most of the samples revealed permeability of 100-500 mD. However, a few samples with a high porosity of 20 percent - show permeability lower than 2 mD. SEM investigations showed that rocks are porous and permeable, and that there is a small amount of carbonate-shaly matrix cement. SEM investigation showed conclusively that even when there was intergranular and intercrystal porosity with clear interpore communication, the presence of significant amounts of matrix cementation, degraded porosity and permeability so that these rocks were probably not capable of producing hydrocarbons. In the Z78 well, the reservoir parameters were also very diversified. In the upper part, from 528-537 m, the rock displayed high porosity (almost 30 percent) and high permeability (more than 100 Darcy). In other sections there are also good reservoir parameters except for 717-726 m, where extensive cementation lowers the porosity to 5-6 percent. Average porosity is equal to 23-26 percent and permeability about 100 mD. Similar to previous wells, porosity values of about 20 percent with permeability close to zero, were observed. Such samples are common in the lowest part of the Sarmatian. Geological core descriptions were also used for lithofacies calibration (Mastalerz et al., 2004). The short descriptions of the cores in two selected depth sections in the Z76 well are presented as example: 15 691-700 m Lithology: the upper part of the section comprises thinly bedded heteroliths of mudstones and sandstones, whilst the lower section exhibits a higher proportion of sandstone heteroliths. Also observed were the frequent synsedimentological deformations, mostly in one directional slope of diagonal laminae, rarely bimodal, there are frequently observed small nodules. In the lower part of the section, the normal sorting of grains is observed, and in some parts frequent bioturbacies have been detected. In addition lithoclasts are occasionally deposited in waves. In other areas nodular concretions are present. Facies: The environment of deposition is characterized by traction currents with a considerable amount of suspension; suspension currents, high movement of deposits in an area with a soft nonconsolidated bottom, a lower share of hemipelagic suspension deposition; probably sporadic influence of wave action and tidal currents; this sedimentation being typical of submarine deltaic slope. Other: Dip is generally horizontal or slightly inclined (0-5o); transportation is generally from one direction, in parts bimodal; the cores from that interval were correlated with logs in the depth section 689 – 698.5 m. 700-709 m Lithology: the upper part of the section is mostly heterolithic with mudstones and sandstones, of various lamina thickness; the dip direction in the sandstones is primarily in one direction, , rarely bimodal; frequent small nodules are visible, current ripple marks are deformed synsedimentologically; in thicker sets a normal grain size distribution is present, and not the frequent modifications due to wave activity; in the middle part of the interval sandstone clasts become more common; laminae are thicker and grains are coarser; in the lower part poorly sorted material occurs, and coarse and medium-coarse sandstones with few mudstone laminae are observed. Facies: the upper part of this section is similar to the previous one (691-700 m). In the lower part there is observed influence of weak suspension currents and traction currents. A considerable result of traction currents is also observed in the upper part. 16 Other: horizontal lamination or minor dips (0-5o), transportation from one direction - of structural dip, is rarely bimodal. The cores from the interval were correlated with logs in the depth section 698.5 – 707.5 m. 2.4 Relationship between reservoir parameters and sedimentary environment The discussion presented below is based on data from the Z75, Z78, Z79 and Z84 wells selected as the typicall wells of the Z gas field. The best reservoir rocks were deposited in turbidite fans. They are thinly laminated, the laminae are equal in thickness and are parallel. Grain size is generally fine. Fining upwards sequences are common. The thickest laminae occur in coarse grained poorly sorted sandy lenses. They are formed in zones of migrating channels in the upper submarine fan. In the middle fan and between channels, sediments are more fine grained and their characteristics are more akin to typical turbidites. Turbidites in the Z gas field exhibit high porosity (up to 20 percent) and permeability (up to 500 mD). In the discussed wells sediments from the submarine fan turbidites do not contain any gas accumulations. Proximally the turbidites are replaced by a series of deltaic sediments. This transition occurs gradually, over a long distance. The transition zone is typified by laminae of fining upwards cycles typical of turbidites diminishing as they are replaced by sandy-muddy packets with a characteristic increase in grain size. Turbidites have low values of GR and rather low NPHI, whilst the occurrence of sandstones predominates over shales. An example is presented from the depth section of 925 – 1025 m in the Z84 well (Fig 2.7a). Deltaic sediments are where the richest gas accumulations are found. Thirteen of the seventeen gas horizons identified in the Z gas field are located in deltaic facies. They are cyclic and it is the reason why the Sarmatian gas deposits are found in stacked reservoirs. Changes in grain size and their sorting are related to cycles in parasequences. At the beginning of the parasequence, grain size increases except for in the top laminae where the grain size decreases. A similar pattern in terms of thickness of the laminae can also be seen. The frequent erosion of bars is typical of deltaic sediments caused by changes in the direction of the channel axes supplying the sedimentary material. The erosion cuts are filled with sands. From a hydrocarbon prospecting point of view there is an important stable geometrical shape of the deltaic sediments. It is related to a decrease in the energy of the sedimentation environment from stream outlets to the open sea. The shape of the deltaic area also depends 17 on the mutual relationship between energy of sources, i.e. current flow in river, sea wave energy and tides. The fusion of the river environment with a considerable inflow of clastic material and the sea environment enables accumulation in the same place, of large amounts of organic matter, sandy material and shaly components. A typical parasequence consists of claystone forming the basis for levee (bar), overlying this are heteroliths Heterolithic bedding is defined as a closely interbedded deposit of sand and mud, generated in environments where current flow varies considerably. The three main types of heterolithic bedding are called flaser, wavy, and lenticular. Flaser bedding is characterized by cross-laminated sands with thin mud drapes over foresets. Wavy bedding consists of rippled sands with continuous mud drapes over the ripples. Lenticular bedding consists of isolated lenses and ripples of sand set in a mud matrix. Heterolithic sediments can be deposited in storm-wave influenced shallow marine environments, river floodplains, tidal flats, or delta front settings where fluctuating currents or sediment supply permit the deposition of both sand and mud. At the beginning of the sedimentation process the heteroliths form as a result of prograding of the bar slope. During the parasequence development phase, the genesis of heteroliths is connected with the channel fill. The top of the parasequence is built up of sandstones from the prograding head of the accumulation levee. The sequence terminates with the deposition of pelagic shales. These represent the maximum flooding surface which coincides with the beginning of the next parasequence. This pattern of deltaic sediments is well illustrated by pairs of GR and NPHI and electric logs (depth sections of 550-690 in the Z75 well) (Fig. 2.7b). In the study area the deltaic sediments are not homogeneous. A typical parasequence is observed in the Z75 well in the eastern part of the central element of the Z gas field. Boundaries between parasequences are distinct and relatively easy to discriminate. A similar picture is observed in the Z84 well. Moving to the west, the distinct picture of the parasequences is not observed. The same problem appears in the Z79 well, where distinguishing the boundaries is difficult. Sediments of the shallow shelf form the top of the Sarmatian succession. Their genesis is associated with the last stage of sedimentation in the Carpathian Foredeep, in which they filled the remaining free sedimentation space. They are not vertically organized, and are not well sorted, but they are composed of fine-grained sediments, mainly claystone and mudstone. 18 They are not prospective for hydrocarbons. However, despite this, four of the highest horizons in the Z gas field are composed of those layers, and are located in heteroliths and mudstones intercalated with very fine-grain sandstones. GR and NPHI track each other closely which is interpreted to mean that these sediments are shaly, as evidenced by the high neutron porosity (high NPHI is observed in shaly formations due to water bound in clay minerals or in small micropores). The depth section of 410-500 m in the Z79 well presents a typical littoral deposit (Fig. 2.7c). RMS amplitude map and GR log for each well from top (~400m) to bottom (~900m) in the study area are presented in the figure 2.8. Shallow marine shelf Fluvial system/Open sea ? Baries Coarsening Upward Fining Upward (b) Shallow marine shelf Deltaic (a) (c) Fig. 2.7 Well logs (GR and NPHI) showing vertical facies distribution; a) an example of turbidities from the depth section of 925 – 1025 m in the Z84 well; low value of GR and rather low NPHI, sandstones dominate over shales, b) a typical parasequence coarsening upward in deltaic sediments in the Z75 well, c) sediments of shallow shelf form the top of the Sarmatian succession (GR and NPHI are very close to one another), sediments are shaly and have high neutron porosity 19 Fig.2.8. RMS amplitude map and GR log for each well shows deltaic facies distribution in the study area 2.5 Organization of succession, its division and correlation diagram Several analyses of sedimentological and structural conditions for the Z gas field area were provided on the basis of well log data (including dipmeter measurements and interpretation). The type example for sedimentological and structural analysis of the Miocene sediments in the Z75 and the Z76 wells is presented on the basis of dipmeter SED (Halliburton) measurements and interpretation (Mastalerz et al., 2004). Organization of a series and its structural-textural characteristics shows that in the profile we find series of sequences (cyclothems), mostly of a prograding character at the beginning and retrograding character in the next stage. In the Sarmatian sediments one can distinguish a few dozen sedimentary sequences with a hierarchic organization with varying character (part of them there are parasequences or sequences of a higher order in a sequential stratigraphical sense). Single sequences run from several to more than a hundred meters in thickness (mainly depending on their rank). A considerable part of the sequences show an almost symmetric 20 structure: in the lower parts there is a visible general increase in the sandy character at higher elevations (it is connected with the gradual increase in the average thickness of the sandy laminae), however shallower sandstones are gradually fine grained, and laminae, especially the sandy ones, become thinner and thinner. Considerably rarer are sequences that are distinctly asymmetric (i.e. almost solely increase in grain size or almost solely decrease of grain size up to the top of sequences). Most of the sequences of a higher order are complex, i.e. they have internal subsequences of higher frequency. Generally speaking, for an accuratecorrelation of sediments in two geological profiles from the, it is important to distinguish marker correlation horizons with a characteristic structure. However in the case of the Z76 and the Z75 wells it was impossible due to their belonging to different structural units and the considerable distance between them. Additionally, the Z76 well and the Z75 well penetrate two sedimentary accumulations related to different depositional systems (as seen on well logs, especially on the arrow plot of the dipmeter) (Mastalerz et al., 2004). In the basin influenced by differential synsedimentary tectonics (as seen in the Miocene succession in the Carpathian Foredeep), a correlation of horizons between separate depositional systems is very difficult. A basic step of differentiation of a sedimentary series, due to genetic sequence stratigraphy criterions, is to find maximum flood surfaces (MFS). These are associated with periods of relative increase in water depth of the basin in which the tempo of growth of accommodation space was not compensated by an accumulation of sediments. These surfaces are connected to phases of maximum range of successive transgressions and are boundaries of genetic sequences. Such surfaces are of fundamental significance due to the chronostratigraphy of horizons in sedimentary series of transitional and shallow shelf types of environment. Parts of them can be connected with episodes of intensive subsidence in the parts of the basin or in the whole basin or with eustatic episodes. Then come the surfaces of the maximum range of progradation, named by some of the scientists as transgression surfaces, which may be treated as helpful horizons of chronostratigraphic significance. Their identification on well logs is more difficult and needs experience and a Dipmeter may be helpful here. Identification of these surfaces is indispensable for more detailed differentiation of the sedimentary series into tract systems in genetic sequences. 21 The basis of a more detailed differentiation may be identification of boundary surfaces for depositional sequences, i.e. distinct erosion surfaces (erosion in sub-areal conditions) or their correspondents from more distant parts of the basin. In this stage it is important to show regressive surfaces of submarine erosion and transgressive erosion surfaces (connected with retrogradation of the shoreline). In Table 2.3 there are important horizons which can be treated as candidates of correlation horizons of various rank. Table 3 and 4 present sketches of divisions in the Sarmatian series of the Z region into stratigraphic sequences (also semi-genetic). This division may form the basis of a general prognosis for the place and geometry of sandy lithosomes. An overview of sequence stratigraphy results in the study area based on 2D seismic data are shown in figure 2.9. In figure 2.10 a comparison of the results of well log correlation (the cross section correlation: Z81 - Z76 - Z82 based on GR and NPHI) and the cross section based on seismic facies classification (Chapter 5) is presented and showed facies distribution various from well to well. In Table 2.3 there are important horizons which can be treated as candidates of correlation horizons. Table 2.4 and 2.5 present sketches of divisions in the Sarmatian series of the Z region into stratigraphic sequences (also semi-genetic). This division may form the basis of a general prognosis for the areal extent and geometry of sandy lithosomes. On the basis of presented data we can say that the main factor controlling sedimentation and the means of filling a basin was local tectonic subsidence. The analyzed series consists of various rank sequences. Each sequence of the main scale is several dozen meters thick and comprises a few sub-ranked sequences of a lower scale, which frequently reveal a complete inner structure and almost symmetrical character. The degree of development of the sequences of lower rank is diverse in different areas, and as such, the correlation of inner parts of sequences can be difficult. In each of the basal sequences, there are several candidates for maximum flooding surfaces, MFS, maximum progradation surfaces, MPS, or erosion surfaces, which may be treated as surface boundaries. This means also that the development of sequences of higher rank (lower scale) was not only controlled by tectonics, or eustatics, but also authocycling (parasequencies). It is worth remembering that sequences of higher rank can range on average from a few meters to several tens of meters in thickness. Differentiation of such sequences and describing their character and variability is crucial for a credible 22 prognosis of the extent and geometry of single reservoir lithosomes. In addition, the inner geometry of such sequences cannot be followed by routine seismic methods. (a) (b) Upper delta Delta lobe SB MFS SB MFS SB Fault ? MFS SB (c) (d) MFS: Maximum Flooding Surface SB : Sequence Boundary Fig. 2.9 Overview sequence stratigraphy in the study area based on 2D seismic data a) seismic facies classification based on seismic attributes (Chapter 5), b) balancing section a) c) some basic sequence stratigraphy interpretation, d) successive progradation of delta lobes deposits an offlap succession with a clinoform geometry (after Frazier 1974) 23 24 Fig. 2.10 Comparison of the result of well log correlation; a) cross section correlation Z-81 – Z76 - Z-82 based on GR and NPHI, (after Mastalerz, et al., 2004) b) cross section basic on seismic facies classification (Chapter 5) Table 2.3 Important correlation horizons (event markers) distinguished in the Z75 and Z76 wells in the selected depth section according to its geological core description (after Mastalerz, et al., 2004) Depth Depth Horizon Well logging features Comments Z75 Z76 the lowest rank MFS 481 490 ? distinct maximum on GR top of fundamental sandy sequence SG9, close to depicted seismic horizon boundary not distinct in the MFS_SG8/9 538 (525 distinct maximum on Z76 well (at the top a ?)/534 GR relatively fine-grained sequence is visible) distinct maximum on MFS_SG7/8 612 623-638 GR and dramatic change of plots in Z76 boundary not distinct in Z76, probable tectonic discontinuity distinct maximum on MFS_SG6/7 696 699 GR and minimum on close to described seismic resistivity, insignificant horizon extremes in other curves MFS_SG5/6 737 733 distinct maximum on probably boundary of lower GR rank not very distinct boundary in MFS_SG4/5 777 776 distinct maximum on GR the Z75 well. Close to described seismic horizon (probably boundary of lower rank) distinct maximum in MFS_SG3/4 863 850 GR and extremes in other logs MFS of lower rank distinct maximum in 908 897 GR and extremes in other logs close to depicted seismic horizon 25 Table 2.4 Genetic sequences, SG, and other stratigraphic intervals distinguished in the Miocene succession in the Z75 well (after Mastalerz, et al., 2004) Depth Depth Seque Paleo Lithology, lamination from to nce slope 463 538 SG9_ Z75 Complex sequence (two distinguished sandy intervals) with almost symmetrical structure. Various lamination, mostly of medium thickness and up to thick laminae WSW (ENE) Complex sequence of a complicated structure and with a significant part comprised of relatively fine grained 538 612 SG8_ material – mostly heterolithic and sandy-shaly with Z75 thinly laminated sandstones. Sandstones are mainly in NNW lower and middle part, mostly with middle and low thickness laminae Sequence of complicated structure and great litho – 612 696 SG7_ Z75 facial variability; from thinly laminated mudstones including nodules and ferrous cements (in top and in bottom about 605 m) to various in facies sandstones NNWSSE (thickly laminated are rare) 696 737 SG6_ Z75 Sequence with distinct tendency of increasing average SSE size of grains to the top and distinct boundaries in parts (signifi of fine-grained sediments. Sandstones of variable cant characteristics with a relatively large amount of thickly dispers laminated ones ion) Sequence of a complex structure (probably two 737 777 SG5_ Z75 equivalent sequences) and various lithofacial characteristics: from thinly-bedded mudstones to thickly bedded sandstones. At the depth of 768 m –a distinct zone MFS of ferrous cement variabl e SSESW? Complex sequence with almost symmetric structure. 777 863 SG4_ Domination of sandstones (including a distinct part of ESE- Z75 thick and very thickly laminated) about fine-grained SE sediments 26 Table 2.5 Genetic sequences and other stratigraphic intervals distinguished in the Miocene succession in the Z76 well (after Mastalerz, et al., 2004) Depth Depth Seque Paleoslope Lithology, lamination nce from to /transport Complex sequence with domination of fine grained (465) (525) SG9_ sediments, various in facies; only in the middle part 491 534 Z76 is there a packet 10 m thick dominated by thin bedded sandstones Sequence of a very complicated structure: several thick intervals of largely sandstone (various (525) 623- SG8_ lamination, mostly middle and thin-bedded). Upper 534 638 Z76 sandstones have much better marked progradation segments, and a relatively large proportion of fine- SW-NE (significant dispersion) grained sediments Complex sequence with an almost symmetric 623638 699 SG7_ Z76 vertical structure; several thicker intervals with a NE ? larger share of sandstones (mostly middle-size (significant lamination, only in the lower part are they thicker); dispersion) generally dominated by fine-grained sediments 699 733 SG6_ Z76 Relatively homogeneous, slightly asymmetric sequence with hard progradation segment; NNE sandstones, only thin and middle laminated Complex sequence with two thicker intervals with 733 776 SG5_ Z76 a domination of sandstones of variable lamination (in the lower part there is a larger share of thick laminae), in the upper part fine-grained sediments NE (great dispersion) dominate. Complex interval (at least two thick sequences) 776 850 SG4_ Z76 with a large share of sandstones of variable lamination, increasing up to very thick in the central part (815-830 m), distinct domination of NNE to NNW fine-grained sediments (MFS zone) SG – genetic sequences 27 2.6 Conclusions The sources of the presented data were geological information, core descriptions, results of laboratory measurements of porosity, permeability, volume of carbonates, mineralogical information, well logging data, and especially dipmeter interpretation and parts of seismic structural interpretation. We do not have a complete data set from all wells in the discussed Z gas field, but on the basis of those which we gathered, we have attempted to build the geological (sedimentological and depositional) framework of the study area. After studying all of the geological information we are aware of the complexity of the structure of the Miocene (Sarmatian) succession. To realize the goals presented in the Ph.D. dissertation, we decided to use all lab data available in all wells and select a consistent group of data from the depth section between 500 and 900 m, comprising the deltaic sedimentation environment. After this we selected other geological and depositional data according to the depth section limit and according to the area surrounding the wells. We intentionally limited the amount of data to make them more coherent and uniform. Petrel® software (Schlumberger) provides us on the one hand, with very good mathematical and computer science tools which make it easier to develop a static geological model on the basis of presented data. On the other hand the model has many limitations and constraints, with some simplification forced by automation of the method. 28 Chapter 3 HYDRAULIC FLOW UNIT Petrophysical parameters including porosity, permeability and water saturation are compulsory input to modeling and simulation of any hydrocarbon reservoir. Among them, porosity and permeability are the most important features controlling productivity of reservoir that can be assessed by its flow properties through a factor named Flow Zone Indicator (FZI). This chapter introduces the concept of hydraulic unit, HU on the basic of FZI and its application to the Z gas field. The HU throughout the field is determined based on conventional core data and conventional statistic methods such as histogram, probability plot and cluster analysis. This chapter introduces the concept of hydraulic unit (HU) and its relationship to the Flow Zone Indicator (FZI) as it applies to the Z gas field. The Fluid Zone Indicator incorporates numerous petrophysical parameters critical for modeling and simulation of a hydrocarbon reservoir, including porosity and permeability. The Hydraulic Flow Units at Z field are determined from conventional core data analysis and conventional statistical approach. 3.1 Concept of Hydraulic Flow Unit The hydraulic flow unit concept provides a method for classifying rock types and predicting their flow properties based on geological parameters and the physics of flow at the pore scale. This concept is significant because it helps unify several theories relating to reservoir rocks and the fluids they contain. Amaefule et al. (1993) suggested that the hydraulic quality of a rock is controlled by its pore geometry. In contrast, other authors have proposed various definitions of hydraulic flow units based on depositional and diagenetic processes. More recently, the concept of the HU has become an important tool in describing a reservoir in term of its flow zones. Bear (1972) defined the hydraulic (pore geometrical) unit as the representative elementary volume of the total reservoir rock within which the geological and petrophysical properties are the same. Hear et al. (1984) defined a flow unit as a reservoir zone that is laterally and vertically continuous and has similar permeability, porosity, and bedding characteristics. Gunter et al. (1997) defined a flow unit as a stratigraphically continuous interval of similar 29 reservoir process that honors the geological framework and maintains the characteristics of the rock types. The focus on detail in one or more aspects of the reservoir flow modeling process can obscure the fundamental reservoir concept in a model study. One way to integrate available data within the context of a “big picture” is to apply the flow unit concept. The concept of hydraulic flow unit was introduced by Ebanks (1987) who defined a HU as a mappable portion of a reservoir within which the geological and petrophysical properties that affect the fluid flow are internally consistent and predictably different from the properties of other reservoir volumes. He described the flow units as the following: 1. a specific volume of a reservoir; it is composed of one or more reservoir-quality lithology and any none-reservoir-quality rock types within that same volume, as well as the fluids they contain, 2. a correlative and mappable unit at the interwell scale, 3. a recognizable section on wireline logs, 4. a unit being in communication with other flow units. However, flow units based on lithostratigraphic characteristics are not always in pressure communication (Fig. 3.1). Fig. 3.1 Various parameters used in defining geologic flow units (Ebanks et al., 1992); four flow units are defined on the basis of lithofacies, pore types, porosity, and permeability crossplots, capillary pressure measurements, and gamma-ray log response (after Ebanks et al. (1992) 30 Understanding flow unit concept in reservoir is very important for reservoir modeling and simulation. Permeability is one of the key parameters influencing reservoir flow properties. Ideally, permeability is measured directly from core analysis or repeated formation test equipment. However, in the absence of direct measurements of permeability, more indirect measurement methods must be used. Several models have been developed to reveal the relationship between porosity and permeability. These are primarily based on empirical and theoretical techniques, most of which employ simple regression analysis. Table 3.1 shows a chronological summary of the models. In reality, permeability not only depends on porosity but also on other factors such as pore space geometry, grain size distribution, specific surface area, tortuosity, fluid saturation and others. The most common equation used to calculate permeability is the Kozeny-Carman equation (3.1) (Kozeny, 1927; Carman, 1937). This approach suffers from two parameters which are either unknown or difficult to calculate: special surface (Sgr) and tortuosity (). e 1 K 2 2 2 S gr (1 e ) 2 3 (3.1) Where: K: permeability (m2), e: effective porosity (fraction), : tortuosity, Sgr : specific surface area per unit grain. In equation 3.1, factor 2 accounts for the assumption that pores are cylindrical with circular cross-sections. In 1993 Amaefule et. al. generalized this equation to include the pore shape (Fs) parameter as follows: K 1 2 Fs 2 S gr e (1 e ) 2 3 (3.2) The term (Fs or Kozeny constant, usually has a value between 5 and 100 in most reservoir rock. The term (FsSgr) is a function of geological characteristics of porous media and varies with changes in pore geometry. The determination and discrimination of the (FsSgr) group is the main point of the HU classification technique. Equations 3.1 and 3.2 are only partially successful in predicting permeability from porosity. In reality, most pores are not circular cylinders and so equation 3.1 has limited applicability. Some parameters such as , Sgr, and Fs 31 are not readily available and so equation 3.2 is also difficult to apply. Amaefule et al. (1993) addressed the variability of Kozeny’s constant by dividing Eq. 3.2 by effective porosity (e) and taking the square root of both sides resulted in: K 1 e Fs SVgr e (1 e ) (3.3) Where K is in m2 and if permeability is presented in millidarcy then the following parameter can be defined: RQI (m) = Reservoir Quality Index RQI ( m) 10 2 K K 0.0314 e e (3.4) z is defined as the pore volume to grain volume ratio: z e 1 e (3.5) FZI (m), designated as the Flow Zone Indicator, is given by: FZI 1 Fs S gr RQI z (3.6) Substituting these variables into equation 3.3 and taking the logarithm of both sides resulted in: log RQI log z log FZI (3.7) RQI vs. z can be plotted on a log-log plot as a straight line with the slope equal to one if FZI is constant for all core samples. Representation of data on a log-log graph is more useful because unit slope lines can be distinguished easily. Data samples with similar but not identical FZI values are located around a single unit-slope straight line with a mean FZI value. Samples with significantly different FZI lie on other parallel unit-slope lines. Each line represents a HU with an associated mean FZI value. The mean FZI value is the intercept of a unit slope line with the coordinate z = 1. The scatter of data about the straight lines is owing to measurement errors in core data analysis and minor fluctuations around main geological controls on pore throat characteristics of rock samples. The basis of HU classification is to identify groups of data that form the unit-slope straight lines on a log-log plot of RQI versus z. The permeability of a sample point is then calculated 32 from a pertinent HU using the mean FZI value and the corresponding sample porosity using the following equation: e K 1014.24( FZI ) (1 e )2 3 2 (3.8) In this equation, the FZI must be correlated to wireline log responses for known core permeability and porosity data. As a first step, the FZI is classified using some sort of cluster analysis, such as a histogram analysis, probability plot or other. Then the FZI must be predicted in wells where only logs are available. This will be discussed in more detail on the next section. Table 3.1 Permeability correlations developed on the basis of pore and grain properties (modified after Babdagli and Al-Slmin, 2004) Model Equation KozenyCarman (1937) Archie (1942) Krumbein and k f Variables 3 2 g 1 2 r 2 eff k 8F k 0.760D2g exp( 1.31D) Berg (1970) Effective pore radius, eff, and formation factor, F. Geometric mean grain diameter, Dg diameter, D. (1943) (1968) , shape factor, and porosity, . and standard deviation of grain Monk Timur Specific surface area, f, tortuosity, 0.136 4.4 k 2 Sw k 80.8 5.1D2 1.385 j Irreducible water saturation, Sw, and porosity, . Median grain diameter, D, and sorting term, , porosity, . 33 Van Baaren k 10Dd2 3.64 mC 3.64 (1979) For clean formation and Domain grain size, D, and sorting index, C, porosity,. Irreducible water saturation, total and 100 (1 Swt) k Swt Denoo For clayey material (1981) 100 2e ( Vwt) 2 k Vwt Swanson Maximum capillary pressure and (1981) S k a b Pc max Herron f 3 exp( Mf ) Mineral composition feldspar Coates (1987) 2 e e k (1990) Coates (1991) 2.08 k 10 0.1 T 2.15 m 2 FFI BVI 3 k 1,014FZI 1 2 2 Amaefule et al. (1993) Ohen et al. (1995) FZI 1 F g k 1,014( FZI )2 water saturation, volume of bound water, bulk irreducible water, and total immovable water content, and textural maturity of sediments k 106.59 Vp / S 2.56 m k 10 / Qv 2.11 k C porosity, effective porosity, bound saturation percent (1 )2 m Sen et al. 2 RQI z 3 1 2 Volume (Vp) to the surface (S) ratio, exchange cation molarity (Qv), and proton NMR decay constant (T1) FFI = free flow index and BVI = Bulk volume irreducible water Flow-zone indicator (FZI), Reservoir quality index (RQI) Kozeny constants (shape factor, tortuosity, and specific surface area) Flow-zone indicator (FZI), Reservoir quality index (RQI), and 34 b1 NMRSWR FZI 1 aNMRSWR 1 1/ C k 1,014FZI 2 3 1 2 Altunbay NMR measured water saturation Flow zone indicator (FZI), FZI a bR1 cRxo fR1Rxo gRGR (Resistivity, gamma ray, GR, grain (1997) hRGD jRxoGD iGRGD mR2 density, GD) Quintero k Cpf 4.6 T et al. et al. nR 2xo pGR2 rGD2 4 2 2m (1999) pf 10i 2 widlk _ 1 / 2 widlh _ m Haro b 1 log R RW a m (2004) Width of T2 distribution at a specific depth interval (T2 width) and of the T2 distribution in a mudsupported facies a m log k n log Swf Water saturation, Swf formation resistivity, Rt, formation water resistivity, Rw, cementation factor, m, and saturation component, n. 3.2 Hydraulic Flow Unit classification technique A reservoir can be modeled by subdividing its volume into several HUs. Recall that a HU is a continuous part of the reservoir that has similar flow properties as characterized by geological and petrophysical properties. Several flow unit identification techniques are proposed in the literature. Some of the more prominent techniques are based on identification of rock type and illustration of the results on the Winland porosity – permeability crossplot (R35) (Kolodziej, 1980) and on the modified Lorenz plot (Gunter, et al., 1997). After calculating pore-throat related parameters of RQI and FZI from core data, HU’s can be discriminated based on FZI values. Although each HU should be associated with one FZI value, in reality a HU typically exhibits a distribution for each FZI around its true mean. This is most commonly caused by random measurement errors in core analysis. When multiple HU groups exist, the overall FZI distribution function is an upper position of the individual 35 distribution functions around their mean FZI. Identification of each mean FZI, or the corresponding HU, can be achieved by decomposition of the overall FZI distribution into its constituting elements. This desuperposition problem can be solved using cluster analysis techniques (Abbaszadeh, et al., 1995). In this section, we apply the following statistical analysis: histogram, probability plot, Ward’s algorithm, and Global Hydraulic Elements (GHE) for HU classification based on core data. 3.2.1 Histogram Because FZI distribution is a superposition of multiple log-normal distributions, a histogram of FZI (with the log scale in the x-axis) should show “n” number of normal distributions for “n” number of HU’s. When clusters are distinctly separated, the histogram clearly delineates each HU and provides their corresponding FZI values. This is the easiest and simplest approach. However, it is often difficult to separate the overlapped individual distributions from a histogram plot. Therefore, this method is not suitable for most field applications because the transition zones between HU’s often clouds the judgment on their identity (Abbaszadeh, et al., 1995). To explore this method, the experimental FZI was calculated from the core data by Eq. 3.6. The distribution of the log(FZI) is shown in Fig. 3.2 which is similar with log(K). Each of the peaks in the log(FZI) distribution should represent a group or cluster with Gaussian distribution, since log(K) have normal distributions. However, the group classification is not apparent since there is a strong overlapping. 36 Phi (dec) K Multi-w ell Interval histogram Multi-w ell Interval histogram 60 60 50 Number of points 40 Number of Points Number of Points 50 40 30 30 20 20 10 10 (a) (b) 0 0. 570 points plotted out of 82307 Curv e Well Depths Mean (%) Std Dev 0 0.01 0.35 0.1 1. 570 points plotted out of 82307 Curv e Well 10. 100. D_2 680.85M - 938.9M 0.23732 0.03449 K D_2 680.85M - 938.9M 1.7518 histogram 0.9244 Multi-w ell Interval Phi Z_72 253.1M - 967.35M 0.24891 0.02485 K Z_72 253.1M - 967.35M 2.0065 0.6968 Phi Z_74 675.4M - 1071.6M 0.22928 0.03607 K Z_74 675.4M - 1071.6M 1.4271 0.9059 Phi Z_75 509.05M - 951.3M 0.26502 0.03111 K Z_75 509.05M - 951.3M 2.5008 0.6602 Phi Z_76 501.1M - 1155.85M 0.25385 0.04303 K Z_76 501.1M - 1155.85M 2.0756 0.9935 Phi Z_77 708.15M - 1091.95M 0.2298 0.02228 50 K Z_77 708.15M - 1091.95M 2.1262 0.6398 Phi Z_78 528.4M - 1092.55M 0.24706 0.03442 K Z_78 528.4M - 1092.55M 1.7368 0.9079 Phi Z_79 642.3M - 798.65M 0.24743 0.05572 K Z_79 642.3M - 798.65M 2.0597 1.126 Phi Z_81 741.1M - 1018.1M 0.259 0.0288 K Z_81 741.1M - 1018.1M 2.5588 0.5286 Phi Z_82 646.05M - 837.35M 0.27038 0.02245 K Z_82 646.05M - 837.35M 2.5332 0.647 Phi Z_84 650.1M - 726.8M 0.22261 0.05854 K Z_84 650.1M - 726.8M 1.4053 1.305 0.24845 0.03691 2.0781 0.8945 40 All Zones 5 4 Gaussian fit for each group Number of Points All Zones 10000. Std Dev Phi Gaussian fit for all data 1000. logK (mD) FZI Mean Depths 30 3 20 2 6 10 1 (c) 0 0.05 0.1 570 points plotted out of 82307 Curv e Well 0.2 0.5 Depths 1. 2. logFZIMean 5. 10. Std Dev FZI D_2 680.85M - 938.9M 0.19939 0.3551 FZI Z_72 253.1M - 967.35M 0.28482 0.2933 FZI Z_74 675.4M - 1071.6M 0.06734 0.3252 FZI Z_75 509.05M - 951.3M 0.48351 0.2635 Fig. 3.2 Histogram of porosity (a), permeability (b), FZI (c) for 570 core data measurements FZI Z_76 501.1M - 1155.85M 0.31387 0.3568 FZI Z_77 708.15M - 1091.95M 0.40795 0.2564 FZI Z_78 528.4M - 1092.55M 0.15977 0.3477 FZI Z_79 642.3M - 798.65M 0.33969 0.3322 FZI Z_81 741.1M - 1018.1M 0.53128 0.1866 FZI Z_82 646.05M - 837.35M 0.4808 0.2734 FZI Z_84 650.1M - 726.8M 0.09748 0.4714 0.32805 0.3394 All Zones 37 3.2.2 Probability plot The probability plot, or cumulative distribution function, is the integral of the histogram (probability density function) and it is smoother than the histogram. The scatter in data is reduced on this plot and the identification of clusters becomes easier. A normal probability plot has a specially arranged coordinate system where each normal distribution forms a distinct straight line. Hence, the number of straight lines and the FZI limiting boundary for each HU can be obtained from the probability plot of logFZI. This exactly corresponds to a linear least-squares regression of data where the slope of the regression line is equal to unity. This method is more useful than histogram method because it is easier to identify straight lines visually, although the superposition effects may shift or distort the straight lines to some degree. In the figure 3.3, six distinct straight lines were recognized. Therefore, we decided to group the core plug data into six clusters, corresponding to six appropriate HUs. Normal Probability Plot 6 0.999 0.997 0.99 4 0.98 5 0.95 0.90 Probability 0.75 3 0.50 2 0.25 0.10 1 0.05 0.02 0.01 0.003 0.001 -1 -0.8 -0.6 -0.4 -0.2 0 Data 0.2 0.4 0.6 0.8 Log(FZI) Fig. 3.3 Normal probability plot of log(FZI) with division into 6 homogeneous groups of HUs with constant FZI 38 3.2.3 Ward’s algorithm approach Ward’s algorithm is an analytical technique in hierarchical cluster analysis. In this method, the distance between data points (FZI values) is calculated, initially treating each sample data as a cluster. Next, the two clusters that are closest in distance are merged and the distance of new cluster from other clusters is computed. The process of distance calculation and emerging of clusters is continued until all available data points are emerged or until the required number of clusters is attained. One of the advantages of the Ward’s algorithm over the others is its special treatment of the cluster variances. Clusters are formed so as to minimize the increase in the within-cluster sums of squares of deviations from their mean. The distance between two clusters causes an increase in the sum of squares if the two clusters were emerged. Therefore, each cluster tends to attain a minimum spread around its mean value while having maximum separation from the other clusters. This is exactly what is expected from a HU (Abbaszadeh et al., 1995). Therefore this method for hierarchical clustering was selected for use in this study. The six clusters are provided as input to Ward’s algorithm based on the probability plot (Fig. 3.3) and as such provide a good means to determine an appropriate number of HU’s for a data set. The results of the clustering are presented in Fig. 3.4. The three black dashed lines and a red line shows the possible cutoffs for the proposed divisions into 3, 4, 8, and 6 groups, respectively. We then decided to use six groups (the continuous red line in Fig. 3.4) that are consistent with the results from probability plot in the previous section. 39 Fig. 3.4 Clustering of the FZI – HU data set into six groups, according to the Ward method 3.3 Critical review on determining HU in the Lower Miocene reservoir of Z gas field 3.3.1 Preliminary outcomes The cluster analysis presented in section 3.2 resulted in six clusters. Care was taken to assign each original core plug data set (K & ) the same number as its associated HU. In this way, the HU of each core plug could be identified and plotted on the permeability porosity plot (Fig. 3.5) Because mean FZI values are not calculated from the histogram, the probability plots or Ward’s HU classification algorithm, a plot of z vs. RQI for each HU was constructed (Fig. 3.6). The unit slope lines were drawn for each HU through their data clusters according to the mean value of FZI calculated for each HU at the intercept with z = 1. The mean FZI values were then used to construct the porosity - permeability relationship within each HU using Equation 3.8. Figure 3.7 shows the porosity-permeability crossplot combined with the HUs 40 for all core data. The curves represent the porosity – permeability relationship based on Equation 3.6 using the mean value of FZI for each hydraulic unit. Simple statistics of permeability, porosity and FZI show that the uniform separate groups are unambiguously described by the mean value of FZI (Table 3.2). For these six defined groups of data, each with homogeneous HUs of constant reservoir parameters, we calculated the equations relating FZI to the permeability and porosity using core data. Finally, the permeability that was calculated on the basis of Equation 3.8 with mean values of FZI for each HUs was highly correlated to the core origin permeability (Table 3.2, Fig. 3.8) (Ha Quang and Jarzyna, 2008 a, b). Table 3.2 Simple statistics of permeability, porosity, FZI and the determination coefficients (R2) for the permeability, calculated from the FZI_mean and from the core in 6 HUs. HUs min mean max min mean max min mean max R2 (k_FZI_mean vs. k_core) K (mD) Nr. of data in HU PHI (%) FZI HU1 28 0.02 0.72 2.82 0.07 0.16 0.233 0.095 0.283 0.400 0.728 HU2 58 0.17 9.15 24.33 0.078 0.21 0.251 0.466 0.734 0.971 0.888 HU3 89 9.78 50.75 120.04 0.15 0.24 0.292 0.997 1.379 1.687 0.645 HU4 117 40.470 144.72 358.55 0.203 0.257 0.315 1.733 2.10 2.563 0.743 HU5 214 79.79 445.77 1461.7 0.189 0.26 0.32 2.587 3.51 4.512 0.603 HU6 64 430.07 1458.96 3631.1 0.229 0.27 0.306 4.555 5.85 8.833 0.411 All 0.97 3.3.2 Global Hydraulic Elements Corbett et al. 2003 proposed the rapid and more straightforward approach to plot the porosity and permeability data on the pre-determined global hydraulic elements (GHE) template (Fig. 3.9) which is constructed on the basis of eq. 3.8. A systematic series of a priori FZI values was arbitrarily chosen to define 10 porosity-permeability elements. Only ten were chosen in order to split the wide range of porosity and permeability parameter space into a manageable number of GHEs (Table 3.3). Table 3.3 Global hydraulic elements (GHE) template parameters GHE1 GHE2 GHE3 GHE4 GHE5 GHE6 GHE7 GHE8 GHE9 GHE10 0.0938 0.1875 0.375 0.75 1.5 3 6 12 24 48 41 The series of FZI values (0.0938 – 48) corresponds respectively to the lower boundary of Global Hydraulic Elements (1-10). This allows any core plug to be rapidly classified in terms of GHEs merely by plotting its porosity and permeability values on the template. There is no need to calculate FZI values. The GHE approach also permits selection of representative samples even when core data availability is limited. The core porosity and permeability data from Z gas field was projected on the appropriate GHE template constructed for each HU (Fig. 3.9). It was observed that the study fit the processing model exactly as predicted: In each HU / GHE pair, the close relationship between permeability and porosity was established and those equations were used to calculate K from Φ. Fig. 3.10 shows the relationship between permeability from the core data and permeability calculated from the means of FZIs in GHE. When Φ is available from the comprehensive interpretation of logs, we can construct a continuous log, as k=f(depth) (Jarzyna and Ha Quang, 2009). The GHE results gave approximately the same number of GHEs as the HUs. It was therefore useful to compare the previous conventional approach (Fig. 3.7) with the GHE approach (Fig. 3.9) to show that GHEs are a useful concept, and the number of arbitrary GHEs on the template is probably about right. In the future, GHEs appear to provide an easy, rapid way of classifying core data (An, 2004). 42 10000 1000 100 K [mD] HU6 10 HU5 HU4 1 HU3 HU2 0.1 HU1 0.01 0.1 0.05 0.2 0.15 0.3 0.25 0.35 PHi [fraction] Fig. 3.5 Porosity-permeability crossplot, the hydraulic unit classification of all the core data 10 RQI 1 HFU6 0.1 HFU5 HFU4 HFU3 HFU2 HFU1 0.01 0.01 0.1 PhiZ z 1 Fig. 3.6 z vs. RQI crossplot of all the hydraulic units. The mean FZI values for each hydraulic unit are given by the intercept of the straight lines at z =1 43 10000 1000 K [mD] 100 10 HU1 HU2 HU3 HU4 HU5 HU6 1 0.1 0.01 0.05 0.1 0.15 0.2 0.25 0.3 0.35 PHI [fraction] Fig. 3.7 Dispersion plot of PHI_core vs. K, and the six HUs defined in the area of core origin data 10000 1000 y = 1.33x0.95 K_pre [mD] 100 2 R = 0.97 10 HU1 HU2 1 HU3 HU4 0.1 HU5 HU6 0.01 0.01 0.1 1 10 100 1000 10000 K_core [mD] Fig. 3.8 Dispersion plot and correlation line between the core origin permeability vs. the permeability calculated from the mean values of FZI for H 44 Fig. 3.9 Permeability vs. porosity data on the background of 7 GHE Fig. 3.10 Permeability, K_GHE, calculated on the basis of relationship RQI vs. Φz for 7 GHE and core origin permeability, K_core 45 3.4 Conclusions The hydraulic flow unit technique has been developed and applied to identify the reservoir characteristics. This technique has a wide variety of practical field applications to both cored and uncored intervals/wells. In the study area, 6 HUs were identified based on 570 core plugs data by applying conventional cluster analysis techniques: these included histogram, probability plot and Ward’s algorithm. The calculated permeability using the 6 HUs classification shows very good result. The determination coefficient R2 between the calculated permeability and the actual permeability measured on core plugs was 0.97 (Fig. 3.8), indicates nearly perfect correlation. Using GHE, the reservoir can be divided into 7 distinct GHEs. The calculated permeability using this method resulted in a correlation coefficient of 0.96 (Fig. 3.11), which is slightly smaller than that of when applying 6HUs. However this method, is very useful for a reservoir with limited core plugs data. In fact using this method means we can reduce the amount of core data taken from the reservoir limited datasets and still provide with acceptably accurate results. Because of the slight increase in correlation coefficient from the HU method over the GHE method, and because of the reservoir complexity at Z gas field, the following was decided: 1) all 570 core plugs from the study area would be used, 2) the HU technique would be used with, 3) a model of 6HUs, therefore, is being used for the reservoir simulation that will be discussed in more detail in Chapter 6. To summarize the methodology: The HU technique first identifies the prevailing HU’s in a reservoir using core data and various cluster analysis techniques. A linear multiple regression (LMR) and alternating conditional expectation (ACE) statistical method is then used to infer HU and permeability prediction at logged wells will be discussed in the next chapter. 46 Chapter 4 CORE – LOG INTEGRATION In this chapter, we will apply two statistical approaches: Alternating Conditional Expectation (ACE) and Line Multiple Regression (LMR) to predict six hydraulic flow units (HU_log) from core and log integration. There is no rock types (RT) descriptive data in the study area so unsupervised clustering method such as K means clustering will be used for facies prediction based on log data. The results of the Rock Type (RT) prediction and Hydraulic Units from log (HU_log) will be also compared and presented in this chapter. Before applying the statistical techniques, it is very important that core and log should be depth matched. 4.1 Core - log depth matching Core - log data Core and well log data from wells in the Z gas field (Z72, Z74, Z75, Z76, Z77, Z78, Z79, Z81, Z82, Z84, and D2) in the NE Polish part of the Carpathian Foredeep were available (Fig. 4.1). Laboratory core measurements included effective porosity, (Φe) and absolute permeability, (K), were taken from various depths in the selected wells. The study dataset included 570 core samples from 11 wells. All primary statistical analyses, i.e., basic statistics, histograms, and FZI calculations, were performed on the full data set. After core-depth matching and min-max standardization of the data, 396 core samples from 10 wells were retained for analyses. We omitted the D2 well to verify our correlations with blind tests. The interval 253.10 – 1154 m was cored. Most of samples were obtained in deltaic sandy-muddyshaly deltaic facies. A small number of samples (12) came from turbidites at depths below 1000 m. Three samples were taken from a littoral facies at a depth between 253.10-272.00 m. The primary focus was on 313 samples taken from deltaic sediments at a depth section between 400 - 900 m. The Z gas deposit consists of several gas areas (Fig.2.3). Sediments are of the same age (Sarmatian) and generally come from the same sedimentation environment (deltaic). Because of local variation in sediment type, and the presence of faulting, the data was further subdivided. Lab data were combined with eight logs from eleven wells. These logs included spontaneous potential (SP), natural gamma ray intensity (GR), resistivity log EN16 (short normal) and EN64 (long normal) and EL14 (short lateral) and EL28 (long lateral), neutron porosity (NPHI), and transit time interval (DT). In a few wells, the bulk 47 density (RHOB) was also available however, this log was used only to assist in the recognition of lithology and saturation. Z gas field logs represented various vintages, and from different log manufacturers. To prevent this non-standardization from introducing artifacts in the log analysis results, the number of logs used was further reduced to a selected uniform group. Core – log depth matching Core and log depth matching is an important problem to address. In many cases, data was selected from the continuously cored section, to achieve good correlation between the laboratory measured values and the values measured from field logs. This approach is illustrated in well Z76 (Fig. 4.1), by matching the depth sections with the porosity from the core (PHI_core) and the neutron porosity (NPHI) from the log. Additionally, the gamma ray (GR) log was used to obtain information about the shaliness of rocks, to help establish a shift of the discreet lab data along the continuous log. It was decided to shift the core data along the depth scale, since samples cut from the cores for lab measurements are very small compared with the space covered by log. Consideration was given to the fact that the Sarmatian formation is thinly laminated and that logging results can be influenced not only by the beds penetrated directly by the well bore or core barrel, but also from lateral variation in beds a short distance away from the well bore not sampled by the core barrel. Correlation coefficients (R) calculated for the log estimated NPHI and the core data PHI_core provide a measure of the success of the depth matching. Additionally, the constant slope of regression lines in selected depth sections with depth matching was used independently to control the established correspondence between the NPHI and the PHI_core. A study of depth matching reveled that after depth matching, the correlation coefficient increased from 0.11 to 0.87, Fig. 4.1). Therefore, depth matching was performed on all wells used in the study. 48 922 922 922 924 924 922 926 Depth [m] Depth [m] 924 926 924 926 926 928 928 928 NPHI PHI_core 930 NPHI PHI_core 928 930 930 0.1 0.2 0.3 0.4 0.1 0.2 0.3 0.4 NPHI PHI_core 0.1 0.2 0.3 0.4 0.1 0.2 0.3 0.4 930 NPHI PHI_core Fig. 4.1 Two approaches to depth matching between the core and log data in well Z76. Crosses – lab porosity – PHI_core, sampled irregularly; continuous curves with triangles – NPHI, sampled regularly, 0.25 m. The two horizontal scales of porosity and two vertical scales of depth relate to the matched data sets 49 4.2 Hydraulic Flow Unit Prediction 4.2.1 Methodology In this section, the goal was to prepare petrophysical data for modeling media flow in the pore space domain. The rock formations were divided into homogeneous hydraulic units based upon parameters measured from laboratory tests and log analysis (Fig. 4.2). At the beginning of two branches of the flow chart (the left hand side is based entirely on core data, and the right hand side is based on core and log data) there was a calculation of Flow Zone Indicator (FZI) in cored sections of the wells. Next, is a section to include log data in the relationships that determine the HUs continuously in a full geological profile, using linear multiple regression (LMR) and alternating conditional expectation (ACE) algorithm. Finally, these methods were extended into the uncored parts of the wells, including log analysis. The first part of the analysis was to establish a training stage, using core results from ten wells. The subsequent prediction stage was performed in the eleventh well. The quantities under consideration were the effective porosity from the core, the permeability from the core, the Flow Zone Indicator (FZI), and the Hydraulic Unit (HU). FZI is a mathematically obtained parameter which characterizes media flow in a rock formation better than simple porosity and permeability. Hydraulic Unit (HU) is derived from FZI. HU and FZI have well defined physical interpretations, and they are obtained by simple mathematical transformation of the well known Cozeny-Carman equation. In the second branch of the flow chart (Fig. 4.2), we included well log data. The process starts with simple statistics, histograms, and a linear two-dimensional regression applied to the core data. The statistical approach also includes an Linear Multiple Regression (LMR) to establish the relationship between the core and log data, and a more sophisticated method of data processing, such as the Alternating Conditional Expectation (ACE) algorithm, to improve that statistical relationship. While dividing the FZI data set into uniform hydraulic units, we applied a clustering method and calculated a probability function to confirm the division of the data set. A correlation coefficient (R) and determination coefficient (R2) provided measures of reliability for the determined relationships between the FZI and the HU obtained from core data and estimated from log. The following decision chart (Fig. 4.2) shows how to apply the selected statistical methods: 50 Fig. 4.2 Flow chart for two procedures applied to obtain uniform HUs in the study data 51 4.2.2 Flow Zone Indicator prediction 4.2.2.1 Linear Multiple Regression (LMR) An LMR is an efficient statistical procedure for determining the linear relationship between a dependent variable and several independent variables. In the multivariate case, a regression function is easy to compute but is difficult to visualize in a two-dimensional space. To solve this problem, an equation is needed to illustrate the relationship between the dependent variable, Y=ln(FZI), and the independent variables, Xi, i=1,2, ..., p. This equation shows the p logs measuring p various parameters X1, ..., Xp, b0, b1….bp predictors: Y=b0+b1*X1+b2*X2+...+bp*Xp (4.1) Typically, equation 4.1 is accompanied by a table showing the regression coefficients from the contribution of each independent variable (i.e., p-log) to the dependent variable. Controlling and selecting the proper number and type of variables, such as the number and type of log and other factors, result in sufficient information to describe the dependent variable. LMR assumes that the relationship between the independent variable and each dependent variable is linear. However, that assumption is rarely satisfied in this study: All twodimensional dispersion plots of Y vs. Xi, and linear regression equations Y=a+bi*Xi, were analyzed in a full correlation matrix. Dispersion plots of the natural logarithm of FZI vs. the GR and NPHI, and histograms of the GR and NPHI are presented in Fig. 4.3. Also tested was the normality of all independent variables to explain the departures from the Gaussian curve due to the geological/geophysical nature of the data and the available accuracy of the measurement (laboratory and log). 52 60 40 Mean StDev N Mean StDev N 40 30 20 60.59 11.74 327 30 Frequency Frequency 50 31.97 4.296 327 20 10 10 15 20 25 30 NPHI 35 40 0 45 2 2 1 1 lnFZI lnFZI 0 0 -1 30 40 50 60 GR 70 80 90 50 60 GR 70 80 90 0 -1 -2 15 20 25 30 NPHI 35 40 45 -2 30 40 Fig. 4.3 Dispersion plots of ln(FZI) vs. GR and ln(FZI) vs. NPHI, and histograms of GR and NPHI; Gaussian distributions included in histograms The goal was to construct an efficient function to determine ln(FZI) from the available log or combinations of log. The selected final regression equation was a compromise between the expected high quality of the result (e.g., high accuracy of the calculation of the dependent variable) and the number of available logs. The accuracy of ln(FZI) was determined from selected sets of independent variables, and was based on the correlation coefficient between ln(FZI) calculated from core data and from multiple regression. Linear multiple regression (LMR) - raw and standardized data The starting point for this task was to select all of the data from 10 wells in the deltaic sedimentation environment section after core-log depth matching (313 samples). After calculating the ranking correlations between ln(FZI) and selected log, the best correlation coefficient was observed between ln(FZI) and an apparent resistivity from a short normal device, R_E16N, equal to -0.353. In other words, no simple correlations exist to predict FZI from log data (Table 4.1). Next, an LMR was applied to raw log data, in order to learn more about the relationship between several independent variables (log data) and a dependent variable (ln(FZI_core)), and to establish which log combinations provide the best equation for predicting the dependent 53 variable. The correlation coefficients (R) between ln(FZI) and selected groups of log are presented in Table 4.2. The results showed that the correlations were not high enough to predict ln(FZI) on the basis of raw data from eight different logs. Four groups of wells were selected based on the distance between the wells and tectonic events dividing the study area: G1(Z72-Z74-Z77), G2(Z75-Z84), G3 (Z76-Z81-Z82), and G4(Z78-Z79). Clustering the data was also justified by the facies development of the Ż gas field. All of the data belong to the deltaic facies. Sandy - muddy-shaly thinly bedded rocks are very complex in lithology, and in the deltaic environment of sedimentation, there are tidal, near shore, lagoon, and mixed facies. By basing the grouping on geology (Fig. 2.3), and following the core description, the selected clusters should be be self-consistent (Mysliwiec, 2006a; Matyasik et al., 2007). After applying the LMR prediction of ln(FZI_log) for each group, the correlation coefficients R calculated from the raw data for each group increased, with the best result being R = 0.78 for G4. Table 4.1 Correlation coefficient (R) between ln(FZI) and log Log SP DT GR NPHI E16N E64N EL14 EL28 R -0.198 0.069 -0.171 0.252 -0.353 -0.221 -0.141 -0.019 Table 4.2 Correlation coefficients (R) between ln(FZI) and groups of log with raw data Log R E16N, SP 0.334 E16N, SP, DT 0.380 E16N, SP, DT,GR 0.383 E16N, SP, DT,GR, NPHI 0.439 E16N, SP, DT,GR, NPHI, E64N 0.440 E16N, SP, DT,GR, NPHI, E64N, EL14 0.445 E16N, SP, DT,GR, NPHI, E64N, EL14, EL28 0.446 Equation (4.2) shows the position of selected variables in the raw data regression from eight log for group G3 (Z76-Z81-Z82). 54 Ln(FZI_G3) = -1.51296 + 0.020277*SP - 0.01482*DT - 0.01101*GR + 0.202763*NPHI 0.02721*E16N + 0.027615*E64N - 0.02083*EL14 + 0.189517*EL28 (4.2) To improve the multiple regression between ln(FZI_core) and the log data, we introduced min-max standardization and added extra variables as combinations of log (Table 43). The min-max standardization was done according to the following Equation (4.3): Vst = (Vx – Vmin)/(Vmax-Vmin) (4.3) where Vx: the current value of the variable, Vmin, Vmax: the minimum and maximum values of the variable in the study section. The min-max standardization transformed the raw log data into the range [0 – 1], so that after the transformation 16 data points were lost for each variable. To reduce the number of lost data from the 14 variables, we standardized all of the data from 10 wells before dividing them into four groups. Table 4.3 Selected independent variables used in a linear multiple regression Logs (raw data) SP [mV] DT GR NPHI [μs/m] [API] [dec.] Logs (max-min) standardization SP_st DT_st Combinations of log A1= EL14_st /EL28_st EN16 [omm] GR_st NPHI_st EN16_st A2= A3=DT_st |EN16_st /GR_st EN64_st | EN64 [omm] EL14 [omm] EL28 [omm] EN64_st EL14_st EL28_st A4=DT_st A5=DT_st A6=GR_st /NPHI_st /SP_st /NPHI_st We obtained better results during the training and testing processes after the min-max standardization than we did using raw data. Equation (4.4) presents the result obtained for the group G3 (Z76-Z81-Z82), with the standardized data from eight log_st, and six additional variables A, which are combinations of standardized log (Table 4.3). ln(FZI_G3) = 1.525422 0.77622*NPHI_st + - 0.63514*SP_st 2.363566*E16N_st – + 1.919128*DT_st 1.48731*E64N_st – - 0.90813*GR_st - 0.98887*EL14_st + 2.37722*EL28_st + 0.562993*A1 + 0.598027*A2 + 0.012519*A3_st – 2.00357*A4 – 0.02804*A5 + 0.230311*A6 (4.4) Clustering the data, using a standardization procedure, and adding additional variables improved the results of the estimation of ln(FZI) on the basis of log (Table 4.4). However, the correlation coefficients were still not very high, so it was decided to see if the ACE algorithm would provide a better estimation of ln(FZI). 55 Table 4.4 Correlation coefficients (R) between ln(FZI_core) and the results of LMR and ACE. LMR ACE Number of 8 log 8 log +6 A 8 log +6 A Groups Wells normalized standardized samples raw data standardized data data G1 Z72-Z74-Z-77 89 0.627 0.762 0.899 G2 Z75-Z84 103 0.735 0.838 0.906 G3 Z76-Z-81-Z82 76 0.627 0.747 0.889 G4 Z78-Z79 45 0.783 0.884 0.982 All data (10 wells) 313 0.446 0.507 0.643 4.2.2.2 Alternating Conditional Expectations algorithm (ACE) An LMR provides a good result when there are functional relationships between the independent and dependent variables. However, in this case a basic LMR led to unsatisfactory, unstable results, since there was a complex relationship between the permeability, the porosity, and additional unparameterized factors, such as the specific surface or radii of pores. Also, the relationship between the log and the permeability of the FZI was poorly defined. The non-linear regression technique ACE was considered as an efficient variable selection method for reducing the subset of significant predictors for a considered dependent variable response (Breiman and Friedman, 1985). A very general and computationally efficient nonparametric regression algorithm, ACE, was applied following similar works in petrophysics (for instance Xue et al., 1996). The ACE algorithm estimates the transformations of variables used in a multiple regression, without a prior assumption of a functional relationship between the dependent and independent variables. An optimal transformation minimizes the variance of the linear relationship between the transformed variable and primary variable, both dependent and independent. Using ACE, arbitrary measurable mean-zero transformations were defined. These yielded a maximum correlation between the primary variables and their transformations in the transformed space. There should be minimal error that is unexplained by a regression of the transformed dependent variable and the sum of the transformed independent variables. Also defined was a correlation coefficient between the transformed dependent variable and the sum of transformed independent variables. The coefficient should be maximal. The power of the 56 method lies in its ease of use and its ability to identify and correct for outliers; these outliers should be removed from a constructed relationship without subjective assumptions. Although ACE provided a fully automated approach to estimating optimal transformations, it was decide to also examine using heuristic reasoning based on the experience gained from the data analysis. In fact, the ACE algorithm permitted incorporating a presumed functional form in the model. However, it should be emphasized that the success of the ACE algorithm, like any other regression method, is dependent on the quality of the data and the underlying association between the dependent and independent variables. An optimal relationship between a transformed response variable FZI_tr, and a sum of transformed independent variables (log_tr)s, was derived by minimizing the variance of a linear relationship between them. We estimated the optimal function, f, between the transformed FZI_tr and the transformed log using equation (4.5), below. Finally, we predicted FZI through the reverse transformation, F-1 (4.6). FZI _ tr f ( log i _ tr ) (4.5) i FZI F 1 ( f ( log i _ tr )) (4.6) i FZI_tr is an ACE transformed FZI, and logi_tr is an ACE transformed logi. Alternating Conditional Expectations (ACE) algorithm The ACE algorithm was applied to find a better correlation between ln(FZI_core) vs. ln(FZI_log). Based on the experience of other authors (Xue et al., 1996) who performed a permeability determination from log data, a selection was made of the best transforms of the dependent variable (ln(FZI)) and the independent variables (logs and A1 -A6 combinations of log, Tab. 4.3). Several approaches were investigated and allowing selection of a group of independent predicting factors in the transformed space (Fig. 4.4). Parameters were excluded when the correlation coefficient between the primary and transformed values was too low (Fig. 4.4b). 57 Using the ACE algorithm, it was possible to calculate a correlation between the transformed natural logarithm of FZI, ln(FZI_tr), and a sum of selected transformed independent variables. Equation (4.7) presents this relationship. ln FZI _ tr f ( (log i _ tr Ai _ tr )) (4.7) i 0.6 0.6 0.4 0.4 0.2 A2_Tr EL14_n_Tr 0.2 0 -0.2 0 -0.4 -0.2 -0.6 3 y = 0.0579x - 0.3226x - 1.9846x + 0.7047 R2 = 0.9986 -0.8 y = 19.888x3 + 0.304x2 + 0.5583x - 0.1217 R2 = 0.4616 2 -0.4 -1 -0.6 0 0.2 0.4 0.6 0.8 0 0.05 0.1 0.15 EL14_n Fig. 4.4a Optimal transformation of EL14 0.25 0.3 Fig. 4.4b Optimal transformation of A2 3 3 3 2 y = 0.0444x - 0.3812x + 2.6381x - 2.5387 R2 = 0.995 2 y = 1.10x R2 = 0.81 2 1 1 LnFZI_tr LnFZI_tr 0.2 A2 0 -1 0 -1 -2 -2 -3 -3 -4 -4 -0.5 0 0.5 1 1.5 2 2.5 LnFZI Fig. 4.4c Optimal transformation of ln(FZI) -4 -3 -2 -1 0 1 2 ∑(log_tr) Fig. 4.4d Dispersion plot and correlation for FZI_tr vs. (log i _ tr Ai _ tr ) i The GRACE program (Xue et al., 1996) based on the ACE algorithm was used to generate an optimal correlation between a dependent variable ln(FZI_tr) and multiple independent variables (log _ i tr Ai _ tr ) . This is accomplished through non-parametric transformations of i the dependent and independent variables. Non-parametric implies that no functional form is assumed between the dependent and independent variables, so that the transformations are derived solely based on the data set. The final correlation is determined by plotting the transformed dependent variable against the sum of the transformed independent variables. Ln(FZI_tr) obtained using the ACE algorithm showed a higher correlation with ln(FZI_core) than the results from LMR (Table 4.4). 58 4.2.2.3 Comparison of LMR and ACE algorithm Results from the LMR (ln(FZI_pre_LMR)) and the ACE algorithm (ln(FZI_pre_ACE)) were compared. The correlation coefficient between ln(FZI_core) and ln(FZI_pre_LMR) was calculated as well as the correlation coefficient between ln(FZI_core) and ln(FZI_pre_ACE) (Table 4.4). The regression was calculated for data from well Z76 and for data from group G3. In both cases ln(FZI_pre_ACE) vs. ln(FZI_core) resulted in the higher determination coefficients (Fig. 4.5). The dispersion of data on the plain of ln(FZI_core) vs. ln(FZI_pre) was caused by errors of regression and diversity of the lithology, resulting in an imprecise HU classification in the data set. At the end of the ln(FZI) and FZI estimation procedure (flow chart in Fig. 4.2), it was decided to use the ACE algorithm to obtain a continuous curve of FZI _log vs. depth. The two curves presented in Fig. 4.6.b have greater similarity compared with similar curves in Fig. 4.6a, especially in the section between points 24 and 38. This reinforced the decision to use the ACE algorithm to estimate ln(FZI), and to calculate permeability (K) along the entire well 2.5 2.5 2 2 lnFZI_pre (LMR) lnFZI_pre (LMR) profile, starting with the uncored section. 1.5 1 0.5 1.5 1 0.5 2 R = 0.6603 2 R = 0.5102 0 0 -0.5 -0.5 -0.5 0 0.5 1 1.5 2 2.5 -0.5 0 0.5 lnFZI_core Fig. 4.5a Comparison of ln(FZI_core) to ln(FZI_pre_LMR) for group G3 1.5 2 2.5 Fig. 4.5b Comparison of ln(FZI_core) to ln(FZI_pre_LMR) for well Z76 2.5 2 2 1.5 lnFZI_pre (ACE) lnFZI_pre (ACE) 1 lnFZI_core 1.5 1 0.5 1 R2 = 0.7559 0.5 R2 = 0.7651 0 0 -0.5 -0.5 0 0.5 1 lnFZI_core 1.5 2 2.5 -0.5 -0.5 0 0.5 1 1.5 2 2.5 lnFZI_core Fig. 4.5c Comparison of ln(FZI_core) to Fig. 4.5d Comparison of ln(FZI_core) to ln(FZI_pre_ACE) for group G3 ln(FZI_pre_ACE) for well Z76 59 2.5 lnFZI_core lnFZI_pre(LRM) 2 1.5 1 0.5 Z_76 0 Z_81 Z_82 (a) -0.5 1 6 11 16 21 26 31 36 41 46 51 56 61 66 71 76 2.5 lnFZI_core 2 lnFZI_pre(ACE) 1.5 1 0.5 Z_76 0 Z_81 Z_82 (b) -0.5 1 6 11 16 21 26 31 36 41 46 51 56 61 66 71 76 Fig. 4.6 Results of FZI_core and FZI_pre (predicted from log) for two approaches in the selected section of well Z76; a) LMR; b) ACE; consecutive numbers of data in the depth section are on the horizontal axis 4.2.3 Hydraulic Flow Unit Prediction (HU_log) from FZI_pre_ACE Table 4.4 presents examples of the prediction results, i.e., FZI and HU in well Z76, based on core data (FZI_core and 6 HU_core) and log data using ACE (FZI_pre_ACE and 6 HU_pre_ACE). In general, strong similarity of the results can be seen, but there are a few depth points at which there is no agreement between the two solutions. The training data was not uniform due to a number of HUs in the selected ranges. For instance, with only 10 data points in HU1 after depth matching, there was insufficient representation of the low media flow parts of the formation in the training data set. Therefore, it was decided to exclude data with GR greater than 75 API in the continuous geological profile (in well Z76 for example). This effectively excluded pure shales from the testing data set. Two solutions were examined for predicting hydraulic units from log using the ACE solution for group G3 in a continuous geological profile for well Z76 (Fig. 4.7). Plots were prepared in Schlumberger Petrel® 60 software. Fig. 4.7 shows the GR (left hand side of the figures) and hydraulic units HU (colored bands on the right hand side). The left contour of the GR shows true natural gamma intensity. The colors mark the completion of the GR curve to the right edge. The GR and HUs estimated for the full data set are seen in Fig. 4.7a. A solution excluding pure shales (GR>75API) is is shown in Fig. 4.7b. The included histogram of the GR (Fig. 4.7) shows that part of the geological profile in well Z76 consisted of pure shales. There is a distinct visual similarity between the GR anomalies and the log predicted HUs for the sandier part of the reservoir in well Z76. In other words, visual inspection suggests that low GR anomalies correlate with high HUs for parts of the reservoir Z_76_logfavorable to fluid flow in the pore space. GR (API) Interval : 200. : 1000. 100 160 140 80 100 60 80 40 60 Cumulative Frequency Number of Points 120 40 20 20 0 0 75 API 25. 3201 points plotted out of 3201 Curv e Well GR All Zones Z_76_log 100. Depths Mean Std Dev 200.M - 1000.M 65.585 13.08 65.585 13.08 Fig. 4.7a Prediction results for FZI_pre Fig. 4.7b Prediction results for FZI_pre for well for well Z76, by applying the ACE Z76, by applying the ACE solution for group G3 solution for group G3 for the full excluding pure shales; left colored field – GR geological profile; left colored field – anomaly, right colored steps – log predicted HU; GR anomaly, right colored steps – log bimodal histogram of GR in well Z76 to show predicted HU. how to exclude pure shales 61 Table 4.5 Comparison of results obtained in well Z76; FZI_core – calculated on the basis of porosity and permeability from cores, HU_core – six hydraulic units determined according to FZI_core, FZI_pre - calculated on the basis of log and using ACE algorithm, HU_pre – 6HUs determined according to FZI_log Depth (m) FZI_core 700.25 1.745 701.25 FZI_pre HU_core HU_pre K_core (mD) K_pre(mD) 1.954 4 4 127.3 184.30 4.478 3.627 5 5 1288.94 791.58 704.50 1.302 1.460 3 3 43.71 49.08 704.75 2.154 1.938 4 4 155.39 147.66 705.00 4.119 2.884 5 5 690.26 501.07 705.25 3.534 4.342 5 5 706.85 697.26 705.75 1.570 1.404 3 3 71.52 55.23 706.50 0.722 0.713 2 2 9.4 9.61 709.50 8.607 4.999 6 6 3312.29 1529.93 709.75 2.563 4.174 4 5 293.7 550.77 711.00 2.831 3.406 5 5 419.98 645.27 711.25 3.291 2.933 5 5 440.49 501.07 924.00 3.033 2.318 5 4 379.29 181.82 924.25 1.929 2.775 4 5 126.38 418.39 925.00 3.572 2.548 5 4 471.37 162.89 925.75 1.082 1.285 3 3 22.9 37.21 928.25 2.370 2.459 4 4 228.51 179.36 928.50 2.525 2.753 4 5 190.48 367.85 928.75 2.625 2.886 5 5 172.38 308.15 929.00 3.757 3.529 5 5 452.87 395.29 929.25 2.779 3.943 5 5 318.43 507.94 929.50 3.698 4.459 5 5 619.67 558.19 930.25 4.149 2.899 5 5 529.29 378.64 62 Values of the permeability from cores (k_core) and the permeability predicted using the ACE algorithm from log (k_pre) in well Z76 are presented in table 4.5. Two dispersion plots illustrate the variability of k_core vs. k_pre for group G3 (Fig. 4.8a) and for well Z76 (Fig.4.8b). The determination coefficients are high enough to confirm the correctness of the proposed approach. The gamma ray intensity and the HU in three wells (Z81-Z76-Z82) on 2D seismic profile are presented in Fig. 4.9 on the background of the seismic section. The Hydraulic Units correlate with the seismic horizons. Fig. 4.9 illustrates the 2D section of a continuous static model for fluid flow in a reservoir. 4.2.4 Validation of results The proposed estimation of FZI and HU based on core data and well log was tested in well D2. The predicted values of FZI_pre_ACE were compared with the core origin FZI in well D2 at two cored depth sections (679-687 and 930-940) (Fig. 4.10a). Solutions from the ACE algorithm were used from group G1 to predict FZI. The predicted FZI ranged between 0.1 – 10 were similar to FZI determined from the core data. The validation results were positive, despite a few observed discrepancies. These differences were due to the large inhomogeneity of the reservoir as confirmed by the geological core description. Similar results are presented in Fig. 4.10b for permeability (K). K_pre(ACE) of Group 3 K_pre(ACE) of Z_76 10000 10000 y = 0.8802x 1.0138 R2 = 0.8049 y = 0.9253x 1.0096 R2 = 0.8303 1000 K_core K_core 1000 100 10 100 10 1 1 1 10 100 1000 10000 1 K_pre(ACE) 10 100 1000 10000 K_pre(ACE) Fig. 4.8a Dispersion plot and correlation Fig. 4.8b Dispersion plot and correlation line between K_core and K_pre(ACE) for line between K_core and K_pre(ACE) for the G3 group the Z76 well 63 64 Fig. 4.9 GR (left hand side) and HU (right hand side) presented in wells Z81-76-82 in group G3 on the background of 2D seismic section; HU color scale is the same as in Fig. 13 LnFZI 1 10 750 930 751 931 752 932 753 933 754 755 10 934 935 756 936 757 937 758 938 FZI_pre FZI_pre 759 LnFZI 1 0.1 Depth (m) Depth (m) 0.1 939 FZI_core FZI_core 940 760 Fig. 4.10a Comparison of FZI_pre_ACE (line) and FZI_core (points) in well D2 K [mD] 1 10 K [mD] 100 1000 1 930 752 932 754 934 Depth [m] 756 936 758 938 K_core 1000 K_core K_pre 760 100 Depth [m] 750 10 K_pre 940 Fig. 4.10b Comparison of K_pre_ACE (line) and K_core (points) in well D2 65 4.3 Rock types classification Lithofacies classification is a purely geological grouping of reservoir rocks, which have similar texture, grain size, sorting etc. Each lithofacies indicates a certain depositional environment with a distribution trend and dimension. Petrophysical groups are classified by porosity, permeability, capillary pressure and pore throat size distribution. A Rock Type combines both these classifications by linking petrophysical properties and lithofacies as part of the reservoir rock type definition (Varavur, et al. 2005). Rushing, et al. 2008 developed a workflow process that identifies and compares three different rock types: depositional, petrographic and hydraulic. Each rock type represents different physical and chemical processes affecting rock properties during the depositional and paragenetic cycles. A summary of rock type definition, data source, and evaluation methodology for selected reservoir description and characterization studies of sandstone reservoirs is presented in Table 4.6. Inspection of the compiled references shows that most of the technical literature addressing rock-typing studies does include some or most of the aspects suggested by Archie's definition. Rock type identification is used in well correlation and is also important in 3D facies modeling of the reservoir. In this study, modeling of an HU should be constrained by rock type model to better understand of the geological basis of each hydraulic flow unit. Due to lack of availability of core facies description in this section, the focus will be on prediction of the rock type using unsupervised clustering, or K means clustering for classification of similar rock types (Ha Quang, 2011). 4.3.1 K means clustering background Cluster analysis encompasses a number of different classification algorithms that can be used to organize observed data into meaningful groups. Today, in reservoir characterization and interpretation workflows, the application of clustering techniques is becoming more popular, especially for automatic core, log and seismic facies determination and classification. Unsupervised clustering methods presently available for manipulating core, log and seismic data with good results for predicting reservoir characterization include: Model-based Cluster Analysis (MCA) (Sang, et al., 2002), Self Organizing Map (SOM) (Skalinski et al., 2005), Unsupervised Neural Network (UNN) and K means clustering (Antelo, et al., 2001). 66 Table 4.6 Summary of selected rock types studies and definitions for sandstone reservoirs (Rushing, et al. 2008) Formation/ Location Rock Type Definitions Davies, et al. [1991] Travis Peak sands, East Texas Salt Basin No specific definitions Porras, et al. [1999] Santa Barbara and Pirital Field sands, Eastern Venezuela Basin Lithofacies, petrofacies Davies, et al. [1999] Wilmington Field Pliocene-Age sands No specific definitions Boada, et al. [2001] Santa Rosa Field, Eastern Venezuela Basin Lithofacies, petrofacies Leal, et al. [2001] Block IX sands, Lake Maracaibo Basin, Venezuela Lithofacies, petrofacies Reference Madariage, et al. 2001] Porras, et al. [2001] Solo, et al. [2001] Sandstone/ C4 & C5 sands, Lagunillas Field, Lake Maracaibo Basin, Venezuela Tertiary & Cretaceous sands, Santa Barbara Field, Eastern Venezuela Basin K1sands,Suria & ReformaLibertad Fields, Apiay-Ariari Basin, Columbia Lithofacies, petrofacies Data Sources Evaluation Methodologies Depositional environments, sand body geometry, dimensions from core descriptions Texture, composition, lithology from microscopic imaging No Quantitative porosity-permeability ranges, provided qualitative indicators Physical core descriptions of both large-scale & small-scale features microscopic imaging of texture, composition, lithology, diagenesis Core-based measurement of porosity, permeability; dominant pore throat diameter from mercury-injection capillary pressure data Depositional environments, sand body geometry, dimensions from core descriptions Texture, composition, lithology from microscopic imaging No quantitative porosity-permeability ranges ;provided qualitative indicators Physical core descriptions of both large-scale & small-scale features; microscopic imaging of texture, composition, lithology, diagenesis Core-based measurement of porosity, permeability; dominant pore throat diameter from mercury-injection capillary pressure data Physical core descriptions of both large-scale & small-scale features; microscopic imaging of texture, composition, lithology, diagenesis Core-based measurement of porosity, permeability; dominant pore throat diameter from mercury-injection capillary pressure data Physical core descriptions of both large-scale & small-scale features; microscopic imaging of texture, composition, lithology, diagenesis Core-based measurement of porosity, permeability, electrical properties; dominant pore throat diameter from mercury-injection capillary pressure data No specific definitions Physical core descriptions of both large-scale & small-scale features; microscopic imaging of texture, composition, lithology, diagenesis Core-based measurement of porosity, permeability; dominant pore throat diameter from mercury-injection capillary pressure data No specific definitions Core-based measurement of porosity, permeability; dominant pore throat diameter from mercury-injection capillary pressure data Used fuzzy logic to predict rock types in uncored wells Marquez, et al. [2001] LL-04 sands,Tia Juana Field, Lake Maracaibo Basin, Venezuela Lithofacies, petrofacies Identification of stratigraphic units and lithofacies from log analysis Physical core descriptions of small-scale features; microscopic imaging of texture, composition, lithology Core-based measurement of porosity, permeability; up scaled permeability using NMR log measurements Ali-Nandalal & Gunter [2003] Pleistocene-age sands, Mahogany Field, Columbus Basin, Venezuela Geological facies, Petrophysic al rock types Facies identified using core- and log-based analyses of primary sedimentary structures Core-based measurement of porosity, permeability and electrical properties Shushufindi Field sands, Oriente Basin, Ecuador Geological facies El Furrial Field sands, Venezuela Geological facies, Petrophysic al rock types Napo formation sands, Oriente Basin Ecuador Petrophysical rock type Formation unknown, location unknown Geological facies, Petrophysic al rock types Ohen, et al. [2004] Acosta, et al. [2005] Guo, et al. [2005] Shenawn, et al. [2007] Facies identified using core- and log-based analyses of stratigraphy, structure, depositional environment Core-based measurement of porosity, permeability and water saturation Facies identified using core- and log-based analyses of stratigraphy, structure, depositional environment Core-based measurement of porosity, permeability and water saturation; pore characteristics from mercury-injection capillary pressure data Core-based measurement of porosity, permeability and water saturation; pore characteristics from mercury-injection capillary pressure data Facies identified using log-based definitions of shale content, lithology (density log) Core-based measurement of porosity, permeability and water saturation; pore mercury-injection capillary pressure data 67 In this study, K means clustering is used for rock type classification because the method is very effective when large quantities of data are available. Macqueen (1967) developed the K means clustering algorithm. This algorithm assign a specific number of centers, k, to represent the clustering of N points (k<N). These points are iteratively adjusted so that each point is assigned to one cluster, and the centroid of each cluster is the mean of its assigned points. In general, the K means technique will produce exactly K different clusters of the greatest possible distinction. The main idea behind the K means algorithm is the minimization of an objective function usually taken up as a function of the deviations between all patterns from their respective cluster centers. The minimization of such an objective function is found using an iterative scheme, which starts with an arbitrary chosen initial cluster membership. Further iterations refine the clustering result (Michael, 1999). The algorithm is summarized as follows (Fig. 4.11): 1. Consider each cluster consisting of a set of M samples that are similar to each other: x1, x2 , x3 , . . . , xm, 2. Choose a set of clusters [y1 , y2 , y3 , . . . , yk], 3. Assign the M samples to the cluster, using the minimum Euclidean distance rule, 4. Compute a new cluster so as to minimize the cost function, 5. If any cluster changes, return to step three; otherwise stop, 6. End. Before applying K means clustering it should be noted that some limitations of data distributions occurred such as variation in size, differing densities and non-globular shapes. They key in overcoming this limitation is to use many clusters (find parts of clusters and then apply a merge strategy) (Fig.4.12). Two stages are applied for overcoming the K means limitation noted above. Firstly, the data are divided up into manageable data clusters. The number of clusters should be enough to cover all the different data ranges seen on the log. 15 to 20 clusters appears to be a reasonable number for most data sets. The second step, which is more manual, is to take these 15 to 20 68 clusters and group them into a manageable number of geological facies. This may involve reducing the data to 4 or 5 clusters (IP-3.5 manual, 2009). 4.3.2 Applying K means for the data group 3 (G3: Z76, Z81, Z82) This process mirrors the approach taken in section 4.2 in which 6 HUs were predicted. Grouped wells (G1, G2, G3, G4) were used to predict the rock types in the study area. After testing six logs provided good response for the rock types selected for the K means clustering: GR, DT, NPHI, EN16 and EN64. For testing, data was selected from top horizon 1 (H1) to top horizon 2 (H2) (~ 500 to ~800 m). The matrix crossplots and histograms for each log exhibit a simple normal distribution, which in case of K means clustering will simplify classification of rock types (Fig. 4.13). Firstly, to overcome K means limitations and cover all of clusters, the data was divided into 14 clusters. The 14 “clusters mean” and standard deviation of logs after applying the K means method are presented in table 4.7. Based on the cluster grouping dendrogram (Fig. 4.14a) we can select the number of clusters. Alternatively the cluster randomness plot (Fig. 14.14b) indicates the number of clusters that can be selected in the study area. Finally, it was decided to reduce 14 clusters to 6 clusters to correspond to 6 rock types (6RTs) in the reservoir. Crossplots of 5 logs against 6RTs in the well Z76 are presented in figure 4.15. In figure 4.15a (GR/DT) and 4.15b (GR/NPHI) with GR < 70 the separation between two main groups (RT5, RT6, RT7) and (RT8, RT9, RT10) can be clearly seen. The RT5 can be seen in figure 4.15b with GR <70 and 0.26 < NPHI < 0.42. Separation between RT8 and RT9 is indistinct. RT6 and RT7 can be seen in the crossplot GR/EN16 and GR/EN64 where GR < 70 and EN64 > 3 for RT4 (Fig. 4.15c, d). The results of the rock type selection and GR in wells of group 3 are presented in figure 4.16. 4.4 Relationship between Hydraulic Flow Unit (HU) and Rock Types (RT) To understand the underlying geology controlings petrophysical properties for different rock types, Svirsky, et al. 2004, investigated available results of sieve analysis, thin sections and other special core analyses giving grain size, sorting, pore geometry and mineralogy. This geological and physical background provides vital links between micro-characteristics of the pore space and commonly available log data, which are used for HU prediction in uncored wells. 69 Rock types can be used to link depositional facies and wireline log response. Mikes al et. (2006) presented ideas for the relationship and upscaling between static models (geological models, reservoir models) to dynamic models (flow unit models). Schematic relationships are presented in figure 4.17: Standard facies models can serve as a template for reservoir models, Flow unit and facies are key elements of the reservoir and geological model, respectively, Reservoir models can be represented by a small number of units, whilst preserving all levels of heterogeneity, the spatial distribution of facies/flow units and their hydraulic flow properties, All elements (facies) and boundaries for a specific depositional environment are universal, whilst facies geometries are specific. The transformation of facies into flow units and hence the geological model into a reservoir model is not simple task. The value of this approach is that the reservoir model preserves the spatial distribution of facies. It is especially this spatial distribution that controls flow on a regional scale. This makes the method an efficient tool to quickly model fluid flow through a reservoir, enabling routine modeling and sensitivity analysis (Mikes al et., 2006) . Hydraulic units are related to geological facies (rock types) distributions but do not necessarily coincide with facies boundaries (Abbaszadeh, al et., 1995). Comparison between rock types (RTs) and hydraulic flow units (HUs) is presented in table 4.8. In figure 4.16 GR, 6HUs and 6RTs of the group 3 is presented to show that the best rock type properties of RT5 response to HU5 and HU6. 70 Start Number of cluster K Centroid Distance objects to centroids N No object move group? Y End Grouping based on minimum distance Figure 4.11 The workflow of the K mean clustering 3 Groups 10 Groups 3 Groups 10 Groups 2 Groups 10 Groups Fig.4.12 The limitations and overcomes limitations of the K mean clustering (after Tan, al et, 2004) 71 400 350 Frequency 300 250 Z76 200 150 100 50 0 20 40 60 80 GR (API) 100 Z81 120 700 440 420 600 400 500 Frequency DT (US/M) 380 360 340 320 Z82 400 300 200 300 100 260 40 60 80 GR (API) 100 0 250 120 0.5 0.5 0.45 0.45 0.4 0.4 0.35 0.35 NPHI (NPHI) NPHI (NPHI) 20 0.3 0.25 300 350 DT (US/M) 400 450 700 600 Frequency 280 0.3 500 400 300 0.25 200 0.2 0.2 0.15 0.15 100 0 40 60 80 GR (API) 100 120 250 300 350 DT (US/M) 400 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 NPHI (NPHI) 450 8 7 7 7 600 6 6 6 500 5 4 3 2 5 4 3 2 Frequency 8 EN16 (OHMM) 8 EN16 (OHMM) EN16 (OHMM) 20 5 4 400 300 3 200 2 100 1 1 1 0 40 60 80 GR (API) 100 120 250 300 350 DT (US/M) 400 450 2 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 NPHI (NPHI) 4 EN16 (OHMM) 6 8 8 8 7 7 7 7 400 6 6 6 6 350 4 3 5 4 3 5 4 3 450 Frequency 5 EN64 (OHMM) 8 EN64 (OHMM) 8 EN64 (OHMM) EN64 (OHMM) 20 5 4 300 250 200 3 150 100 2 2 2 2 1 1 1 1 50 0 20 40 60 80 GR (API) 100 120 250 300 350 DT (US/M) 400 450 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 NPHI (NPHI) 2 4 EN16 (OHMM) 6 8 2 4 EN64 (OHMM) 6 8 Figure 4.13 The matrix crossplots of 5 logs (GR, DT, NPHI, EN16 and EN64) and histograms for each curves show data distribution of group 3 (Z76, Z81, Z82) before applying K means clustering 72 Table 4.7 The statistics of the group 3 (Z-76, Z-81, Z-82) for 14 clusters after applying K means clustering Cluster Groups Randomness 20 19.5 Lever cut off to select six Grouping clustersDendrogram Cluster 19 18.5 Randomness Ratio (Low more Random) 18 27 26 25 24 23 22 21 20 19 18 17 16 10 17 16.5 16 15.5 15 14.5 14 13.5 13 12.5 12 11.5 11 10.5 15 13 17.5 2 9 14 12 7 6 11 8 5 1 4 3 10 9.5 Number of Cluster Groups : 6 Clustering Method : Maximum distance between all objects in clusters 2 3 4 5 6 7 8 9 10 11 12 13 14 Number of Clusters (a) (b) Figure 4.14 Applying K mean clustering for the group 3: (a) The cluster grouping dendrogram shows 4 clusters with diffident 4 colors respond (b) The cluster groups randomness 73 Z-values: 6RTs 0.2 96 0.25 104 0.3 112 DT 0.35 NPHI 120 0.4 128 0.45 136 0.5 144 Z-values: 6RTs (b) 0.15 88 (a) 30 40 50 60 70 80 90 30 100 40 50 60 GR RT5 RT6 RT7 RT8 RT9 90 100 Z-values: 6RTs 7.2 9 80 RT10 Z-values: 6RTs F7 F10 F8 F9 F6 F5 1 1.6 2 2.4 3 3.2 4 EN16 5 4 EN64 4.8 6 5.6 7 6.4 8 70 GR 0 0.8 (c) 40 50 60 70 GR 80 90 100 (d) 40 50 60 70 80 90 100 GR Figure 4.15 The cross plots of data in well Z-76 shows 6 rock types (6 clusters) a) GR/DT, b) GR/NPHI, c) GR/EN64, d) GR/EN16 74 Figure 4.16 Comparison between 6 rock types (6RTs) and 6 hydraulic flow unit (6HUs) 75 Figure 4.17 Relationship between geological model, reservoir model and flow unit model (after Mikes D., al et. 2006) 1. Geological model and assignment of its elements 2. Reservoir model and assignment of its elements 3. Micro simulation model, consisting of numerical flow simulation on all flow cell models of the reservoir, yielding effective one- and two-phase permeabilities, and capillary pressures 4. Macro simulation model, consisting of numerical flow simulation on the entire reservoir, yielding production history 76 Table 4.8 Comparison between hydraulic flow units (HUs) and rock types (RTs) in the study Data Hydraulic Flow Units Cores, logs Driver FZI_core (K and Phi) Classification/ Prediction Number Probability plot, Ward’s cluster, ACE and cutoff of GR 6HUs Characterizations Similar fluid flow zone in the reservoirs Rock Types Logs Log properties: shape, value of parameter K means clustering 6RTs Lithologies facies, depositional environments facies (Reservoir and non-reservoir flows) 4.5 Conclusions Effective methods were proposed to calculate the Flow Zone Indicator, FZI, as a continuous function of depth using well log data, and to determine hydraulic units in a reservoir. These definitions of the homogeneous parts of a formation will form the basis for media flow modeling. The study started from core data, since they are relatively easily measured and available, and since they are frequently proposed to reservoir engineers as the only data available to determine permeability. Our goal was to show that the correct combination of well log and core data will provide reservoir engineers with a tool for determining FZI and permeability in a continuous geological profile. Statistical methods that combine core and log data provide an effective and multidimensional tool for reservoir engineers to construct a static model of a reservoir that agrees well with reality. Correctness of statistical data should be taken into account in the defined range of accuracy. The range of accuracy obtained in the study is acceptable, for the given precision of the laboratory measurements and logging measurements. The choice of alternating conditional expectations algorithm (ACE) instead of linear multiple regression (LMR) was based on the assumption that petrophysical parameters belong to a group of fuzzy data. The relationships between the individual factors are known and we can describe them, yet it is very difficult to parameterize most of these relationships. When applying ACE heuristic reasoning was included for the relationships between real parameters, but a “blind” statistical tool was used to determine the optimal transformation of data, treating the data as values without geological meaning. 77 The method proposed for obtaining core origin FZI, for determining a continuous log of FZI on the basis of well log in full geological profile, and for determining homogeneous hydraulic units in a reservoir rock in the study area was tested and validated. Since permeability is an extremely important reservoir parameter for reservoir engineering, the solutions for modeling this parameter described in literature and tested by various authors are of great interest. From a practical point of view, it is very important to develop an effective method for precisely dividing the Sarmatian reservoir into HUs. The permeability predicted from logs in uncored sections of individual wells was correct and could be used in a modeling stage. Statistical methods turned out to be useful, flexible, and effective, and they provide the interpreter with tools that deliver acceptable results. Unsupervised K means clustering is effective for classifying six rock types. The ACE method is good for predicting FZI from core and log. The results obtained in this chapter were used to create a static model of the reservoir for media flow modeling in the next chapter. 78 Chapter 5 STATIC MODELING In chapter 3, using core and well log data we divided the reservoir into six HUs (Tab. 3.2, Fig. 3.7). Integrating both types of data, the FZI_log was predicted using the Alternating Conditional Expectations (ACE) algorithm and then six HU_logs were generated (Fig. 4.5). Six rock type logs (6RTs) were classified using K-mean clustering (Fig. 4.25) in Chapter 4. In this chapter, 2D seismic data and log data were integrated for static modeling by applying geostatistical methods as Sequential Indicator Simulation (SIS) for rock type and hydraulic flow unit modeling and Sequence Gaussian Simulation (SGS) for reservoir properties (PHI, Sw) modeling. This chapter is divided into three parts: 1. A discussion of the basic theory of reservoir modeling, spatial relationships and geostatistical (SIS, SGS) methods. 2. Creating a 3D structure model with four horizons, derivation of Rock Type and hydraulic flow unit models. 3. Generation of property models (PHI, K, Sw and Net to Gross, NTG) constrained by the HU model. 5.1 Reservoir modeling overview 5.1.1 Reservoir modeling workflow Three dimensional reservoir modeling is a broad field of expertise in which geostatistics is one of several key components. The aim of 3D modeling is to provide one or more alternative 3D numerical models to represent those geological, geophysical, and reservoir engineering aspects of the subsurface that help achieve the study goal. These numerical models are used to estimate key reservoir parameters such as original oil in place (OOIP), to predict production performance, and to provide uncertainty statements when needed (Caers, 2005). Generally, we can divide reservoir modeling into two main parts: static modeling and dynamic modeling. There are four steps to reflect the different scales of heterogeneity modeling: 1. to build the structural framework of the reservoir, 2. to model the facies architecture and the distribution of the different facies types, 79 3. to describe the petrophysical properties, 4. to generate a homogeneous model which can be upscaled model for application to fluid flow simulations. In this study, the main modeling workflow is shown in figure 1.1. To create the fluid flow simulation (Chapter 6) we modeled the Hydraulic Flow Units that constrain the rock type model. In the next step, the property model (porosity) was constrained by HU model. 5.1.2 Deltaic facies and spatial relationships It was particularly difficult to generate an accurate 3D distribution of hydraulic flow units (HU) using only geostatistical methods and data only from 7 wells. To improve the HU model, we constrained it by the facies model. The facies model provided additional spatial classification of porosity and permeability, which are the two main parameters influencing FZI. Some basic concepts of deltaic facies and spatial relationship are introduced below. Deltaic facies: As discussed in Chapter 2, the depositional environment in the study area is mainly influenced by deltaic processes. The deltaic lobes have a distinct morphology which reflects the dynamic fluvial and marine processes controlled by variations in climate and paleogeography. This explains why the facies distribution is very complex in the deltaic environment. An example of a deltaic model is shown in figure 5.1a. The overall geometry of a distributary channel deposit is of a low sinuosity sand body elongated parallel to the channel axis. Mud plugs accumulated at the end of flood stages locally compartmentalize the reservoir sands. Sand thicknesses decrease rapidly away from the channel axis. Parallel to the distributary channel, the sand thickness decreases over a distance of 5 km from a maximum of 5 m at the fluvial channel mouth, to 25 cm within coalescent braid bar deposits on the delta front (Lang et al., 2000). Variogram and spatial relationships: In figure 5.1b we can easy see that distance of the point C (unknown point) is closer to point B than point A but the value of parameter in that point (for example PHI) of point C should be closer to point A than point B because both of points C and A belong to the channel. It is clearly explained: 80 - how an unknown point relates to a known point with the facies constrain, - how properties modeling is strongly influenced by facies models. The variogram is the most commonly used measure of spatial correlation for cell based facies and property modeling. Various discussions on variogram and variogram modeling are available in the literature (Clayton, 2002). A variogram is a description of the variation in a property, based on the principle that two points close together are more likely to have similar values than points far from each other (Fig. 5.2 a, b). The concepts of variogram parameters and the variogram ellipsoid around a grid cell are shown in figure 5.3b and 5.3b. The variogram is defined as the expected value, 2(h). Experimentally, the semivariogram is one half of the variogram that for lag distance h is defined as the average squared difference of values separated approximately by h (Eq. 5.1). (5.1) Where: h : lag vector representing separation between two spatial locations, N(h) : the number of pairs for lag h, u : vector of spatial coordinates, z(u) : variable under consideration as a function of spatial location, z(u+h) : lagged version of variable under consideration. 81 (a) CHANEL Nearest known value data point B Unknown point ? C A (b) Fig. 5.1 Deltaic facies model and special relationships control by facies distribution a) reservoir heterogeneity and longitudinal facies variation developed within the eastwest distributary channel of the Neales Delta (modified from Lang et al., 2000) b) how unknown point relates to a known point (Gocad Manual, 2009) 82 Because geological data is usually anisotropic (at least between the vertical and horizontal directions), variograms can be calculated in several different directions. These are commonly created with the major and minor axes in the XY plane and in the vertical direction. However the major axes may not necessarily follow geological layers. Major direction: The major axis or direction defines the direction where the sample points have the strongest correlation. The angle of this major direction can be changed interactively by editing the direction in the search cone. The angle is specified as the clockwise angle from north (in degrees) for the main search direction. Minor direction: The minor search direction is perpendicular to the major direction. Vertical direction: The program searches for sample pairs vertically using the vertical search distance. Orientation is not used when calculating a vertical sample variogram. In the study area, due to the irregular distribution of wells it was difficult to generate a good horizontal variogram model (Fig. 5.2c). In the vertical direction we have a better variogram model with the range about 10 m (Fig. 5.2d). Normally, the range of a variogram could be different in different directions and often vary vertically, therefore in this study we used the same variogram parameters model for all stochastic models (Tab. 5.1). Table 5.1 The grid parameters used for stochastic variogram models Type of function in variogram Exponential Vertical Range Minor Major Azimuth Nugget 10 4000 5000 0 0.01 5.1.3 Geostatistical methods overview Traditional reservoir modeling techniques use variogram and two-point statistics to represent geological phenomena that have complex geometrical configurations. The use of multipoint statistics improved the modeling techniques in recent years, reducing the limitations. The Multiple Point Statistics (MPS) module in Petrel® 2009 for facies modeling release reintroduced multipoint geostatistics, providing a new method to model complex geological features and connectivity. 83 However, in this study, with lack of geological information to create a training image for MPS modeling process, we used two methods that are still the most popular and flexible techniques for geostatistical applications - Sequential Gaussian Simulation (SGS) for continuous variables like porosity and Sequential Indicator Simulation (SIS) for categorical variables like rock type and HU. Some main points for SIS and SGS are reviewed below. Sequential modeling is a general approach to conditional stochastic simulations. Gaussian random models are used in statistics and simulations due to their analytical simplicity. Gaussian models are theoretically consistent models. The Gaussian approach is also related to maximum entropy and correspondingly to the maximum disorder in the data. It is not the best choice when spatial correlations between extremes are of special interest. In that case we can use another nonparametric model like indicator based simulations. (a) (b) (c) (d) Fig. 5.2 Variogram modeling a) concepts of variogram parameters, b) variogram ellipsoid around a grid cell. The values of cells outside the ellipsoid are independent of the value of the current cell (Roxar Manual, 2009), c) horizontal variogram of porosity modeling (7 wells), d) vertical variogram of porosity modeling (7 wells). 84 Sequential Gaussian Simulation (SGS) The basic idea of Sequential Gaussian Simulation (SGS) is similar to kriging. Recall that kriging gives us an estimate of both the mean and standard deviation of the variable at each grid node, meaning that a variable at each grid node can be represented as a random variable following a normal (Gaussian) distribution. Rather than to choose the mean as the estimate at each node, SGS chooses a random deviate from this normal distribution, selected according to an uniform random number representing the probability level. So, the basic steps in the SGS process are: generate a random path through the grid nodes, visit the first node along the path and use kriging to estimate a mean and standard deviation for the variable at that node based on surrounding data values, select a value of random deviate from the corresponding normal distribution and set the variable value at that node to that number, visit each successive node in the random path and repeat the process, including previously simulated nodes as data values in the kriging process. A random path is used to avoid artifacts induced by walking through the grid in a regular fashion. We include previously simulated grid nodes as “data” in order to preserve the proper covariance structure between the simulated values (Bohling, 2005). Sequential Gaussian simulation generates a set of equally probable realisations of a 3D porosity field. Each simulated realisations also different share the same global distributions and spatial correlation. The differences between realisations characterise the variability and uncertainty in the model. Sequential Indicator Simulation (SIS): The sequential indicator simulation (SIS) includes all data available within a neighborhood, including the original data and all previously simulated values. The objective is to generate a joint realization of the random variables at the unsampled locations. The sequential simulation approach requires simulation of a prior distribution at each unsampled location. The SIS method is similar to Sequential Gaussian Simulation (SGS), expecting that indicator kriging is used to build up a discrete cumulative density function (CDF) for the individual 85 categories at each case and the node is assigned a category selected at random from this discrete CDF. Very briefly, an indicator representation for a categorical variable such as facies would be formulated as: 1 i (u ; k ) 0 (5.2) where: 1, when facies k present at uα, 0, otherwise, and we have one indicator variable for each of the K different facies. We can then use kriging (based on indicator semivariograms) to produce a set of facies membership probabilities at each grid point, build up a cumulative distribution function (CDF) from the probabilities, and select a facies as random from the CDF (Bohling, 2005). Sequential Indicator Simulation results are realisations of facies distributions. A set of equally probable realisations can be post processed to obtain statistical inference used directly in further simulation steps as independent realisations. 5.2 Structure modeling Structure modeling is the first step in the reservoir modeling workflow (Fig. 1.1). In the Petrel® system structure modeling is subdivide into three main steps (Petrel® Manual, 2009): Fault modeling: defining the faults in the geological model which form the basis for generating the 3D grid. Pillar Gridding: generating the grid representation as the base of all models. Making Horizons: building the zones in the reservoir model and then creating the layers in the zones. In this study, we have a 2D seismic survey with 25 lines which were recorded in 2002 (Fig. 5.3a). As the first stage in processing the miss-tie between all lines was corrected. 86 5.2.1 Miss-tie correction for 2D seismic survey A simple, but effective algorithm based upon weighting value assignments using a variance criterion was used. It assumed miss-tie values to be random variables. The algorithm satisfies the following requirements: after correction the miss-ties should be reduced to a minimum after the error adjustment, the method should be applicable to any survey configuration. There should be a way to define a weighting factor to good lines (reference lines) in comparison to lines showing data of poor quality (Petrel® Manual, 2009). An example of the result before and after correction the miss-tie between two lines is presented in figure 5.3b. 5.2.2 Horizons picking and creating the 3D grid Horizon picking In this study we only focused on picking the Miocene horizons in the lower part of borehole profiles and also ignored some small faults that were mentioned in some geological setting documents (Chapter 2). After the miss-tie correction, four horizons - H1, H2, H3 and H4 were mapped on the basis of the phase of maximum amplitude on 25 seismic lines corresponding to well markers (well tops ) Top_7, Top_12, Top_15 and Top_17 (Fig. 5.4a) (Gas Field Z-L, Report, 2005). 4 surfaces that were created by gridding corresponding to 4 horizons are presented in figure 5.4b. We can see that 7 wells (Z75, Z76, Z78, Z79, Z81, Z82 and Z84) are located at the central part of the study area, where there is a structural high, coinciding with good reservoir potential for the gas flow (Table 6.1 in Chapter 6). Another group of 3 wells (Z72, Z72, Z77) is located in the eastern part of the study area. Creating the 3D grid The next step of structure modeling is to create 3D grid blocks within the reservoir base volume of 4 horizons. The volume of reservoir is divided into 3 zones: Zone1, Zone2 and Zone3 between each horizon (H1-H4). Initially, in order to capture all of the reservoir property down to fine level, a very fine model was created with the grid block size being 50m 87 x 50m in the X and Y direction. In the vertical direction - Z, we used grid size about 1 meter as the layer scale. So, the reservoir was divided into 189 layers. This grid method led to a grid that was 170x168x198 in size containing 8934354 grid blocks in total. Since it was assumed that reservoir contains no important faults, the gridding process was simplified and any faults were not included. Basic parameters of the 3D grid are shown in Table 5.2. At the final step, on the basis of the check-shot data of 7 wells in the central part the study area the model were converted from time (ms) to depth (m) (Fig. 5.4b). In order to keep better correlation between seismic and log data and also focus on the central part where better reservoir is potential in the next step 7 wells were used for rock type and properties modeling. Table 5.2 The 3D grid parameters Zone Average thickness (m) Number of layers Number of cells Zone1 (H1-H2) 58 54 2436642 Zone2 (H2-H3) 55 44 1985412 Zone3 (H3-H4) 100 100 4512300 Total 213 198 8934354 88 (a) Before After (b) Fig. 5.3 2D seismic data in the study area a) 25 lines 2D seismic survey and 10 wells location b) the miss-tie correction for the 2D seismic survey 89 H1 H2 H3 H4 (a) H1 H2 H3 H4 (b) Fig. 5.4 Horizon picking a) four horizons interpreted basically on 25 lines of 2D seismic and the well-tops of main interpreted horizons: H1(Top_7), H2(Top_12), H3(Top_15) and H4(Top_17), b) the 3D structure mode in the study area; four surfaces were created from 4 horizons H1, H2, H2 and H4 and 10 well locations 90 5.3 Rock Type (RT) and Hydraulic Flow Unit (HU) modeling Several new methods are presented in the literature discussing use of 3D seismic data to improve facies modeling. But only a few papers discuss using 2D seismic with facies modeling. Shuguang, et al. (1999) presented conditional 3D simulation of lithofacies with 2D seismic data using a new methodology and the corresponding GSLIB-type program to integrate 2D average information, such as vertically averaged lithofacies proportions, into estimation/simulation of 3D lithofacies distributions. Seismic information, available in 2D, is used in the cokriging of the 2D average lithofacies proportions. Results showed that even with limited well data, the input vertical lithofacies proportions (which carry the seismic information) is honored quasi-exactly. Three steps were taken: co-located cokriging is used to incorporate 2D seismic information into the estimation of vertical lithofacies proportions, indicator kriging is used to derive the facies conditional probabilities at each of the 3D simulation grid node; this 3D indicator kriging uses the hard well data and the previously estimated vertical proportions, p-field simulation algorithm is used to draw simulated lithofacies indicators from the previously obtained distributions. In this study, facies modeling on the basis of 7 wells lead us to unexpected results so in this part we also integrated log information from 7 wells and seismic data from 25 lines of 2D seismic to improve the facies modeling. The results of facies modeling are used to constrain HU modeling and 3D property modeling. The main steps used in this study are shown in the workflow below (Fig. 5.5) (Ha Quang and Jarzyna, 2010). 91 Fig. 5.5 The workflow for reservoir modeling in this study 92 HU_MODEL 2D SEISMICS SGS: Sequence Gaussian Simulation SIS: Sequence Indicator Simulation UNN: Unsupervised Neural Network ACE: Alternating Conditional Expectation K.C.A: Kozeny-Carman equation modified by Amaefule PHI_MODEL SGS SIS PHI_log SIS RT_MODEL Facies_UNN UNN Attributes PSEUDO 3D SEISMICS Mis-tie Correction Horizon Picking Surfaces mapping Up scaling Amplitude to 3D Grid Amplitude Modeling SGS AMPLITUDE MODEL ACE ECLIPSE K_MODEL HU_log GR_cutoff FZI_log Upscaling K.C.A RT_log K-mean A_logs LOGS FZI_mean HU_core Ward’s FZI_core CORES 5.3.1 Conversion of 2D seismic to pseudo 3D seismic Nowadays, 3D seismic is the most common and effective method for reservoir structure interpretation (horizons and faults) and properties prediction (porosity, permeability, fluid saturation). Rapid improvement of the strong hardware and software in the petroleum industry allowed us to visualize and extract 3D objects (geobody) directly from 3D seismic and improve reservoir understanding, detect anomalies, and define facies model. In the study area the 2D seismic survey has a density of 25 lines (Fig. 5.3a) and this dense grid can be leveraged to overcome some of the limitations of 2D seismic data. At the first step of modeling we converted the 2D survey to a pseudo 3D seismic cube using the Sequence Gaussian Simulation (SGS). The main steps for converting the 2D seismic survey to a pseudo 3D seismic cube are shown in the workflow (Fig. 5.5, left column). The volume of the 3D seismic cube was limited to the reservoir interval study - between top of horizon H1 and bottom of horizon H4 (Fig. 5.6a). The 2D seismic profiles were upscaled into the 3D grid (Fig. 5.6b, 5.7a) using the 2DSeis2Seis3D plug-in of Petrel® (SLB, plug-in, 2008). Then, SGS geostatistical technique was applied to interpolate all data in three directions. The result is named the pseudo 3D seismic cube and the amplitude model is shown in figure 5.6c. The amplitudes in both the pseudo 3D seismic and the 2D seismic were compared for validation purposes. The results are shown in figure 5.7. This indicates clearly that the pseudo 3D seismic is well matched to the 2D seismic. This pseudo 3D seismic will be useful for extracting attributes and geobodies for seismic stratigraphic classification in the next sections. 93 H1 H4 (a) (b) (c) Fig. 5.6 Convertion from 2D seismic to pseudo 3D seismic cube using SGS a) 2D seismic survey - top and bottom of the reservoir, b) 2D seismic profiles - uplscaled to 3D grid, c) pseudo 3D seismic cube after applying the SGS method. 94 The block average of 2D seismic (a) 2D seismic Pseudo 3D seismic (b) Fig. 5.7 Pseudo 3D seismic cube a) the block average data of 2D seismic after upscaling into 3D grid, b) comparison between 2D seismic profiles and the pseudo 3D seismic cube. 95 5.3.2 Rock Type modeling constrained by seismic facies model First we modeled Rock Type because the outcome of this modeling narrowed the range of possible porosity and permeability models and because the Hydraulic Flow Unit properties can vary significantly between each rock type. Three methods for the 3D rock type modeling are applied in this study responding to the workflow in figure 5.5: 1. seismic facies extraction volume, 2. seismic facies classification using Unsupervised Neural Network (UNN) – deterministic statistic method, 3. Rock Type modeling using Sequential Indicator Simulation (SIS) - stochastic method. Applying K mean clustering method to the RT well log data allows us to derive 6 classes, as summarized in section 4.3. 5.3.2.1 Seismic facies extraction volume Seismic surveys typically provide exhaustive coverage of the reservoir volume, but unlike wells, seismic provides only indirect information on reservoir properties and at a much coarser scale and resolution. The scale of observation can usually be estimated, typically on the order of 10 to 100 ft in the vertical and 100 to 1000 ft in the horizontal, depending on the reservoir depth. The difference between the seismic data scale and the scale at which geocellular models are built can be one order of magnitude or more in the vertical dimension. Such a difference is of critical importance when simulating fluid flow: this is why a geocellular model is needed. Seismic facies and seismic attributes Seismic facies analysis is an example of other than structural information derived from seismic. Seismic facies analysis is an useful tool for reservoir characterization. A seismic facies unit can be defined as a sedimentary unit which is different from adjacent units in its seismic characteristics. Parameters that should be taken into consideration in the seismic facies analysis are as follows: reflection amplitude, dominant reflection frequency, reflection polarity, interval velocity, reflection continuity, reflection configuration, abundance of reflections, geometry of seismic facies unit, and relationship with other units (Roksandić, 2006). 96 Interpretation of seismic facies data may be either direct or indirect. The purpose of the direct interpretation is to find out geological causes responsible for the seismic signature of a seismic facies unit. So, the direct interpretation may be aimed at predicting lithology, fluid content, porosity, relative age, overpressure shales, type of stratification, geometry of the geological body corresponding to the seismic facies unit and its geological setting. The indirect interpretation is intended to reach some conclusions on depositional processes and environments, sediment transport direction, and some aspects of geological evolution (transgression, regression, subsidence, uplift, erosion) (Roksandić, 2006). Seismic attributes and their applications as geological indicators for reservoir interpretation are presented in Table 5.3. Generally, frequency attributes relate to bed thickness, wave scattering, and absorption. Time attributes relate to structure and amplitude attributes relate to stratigraphy (Chopra, et al., 2005). In this study, several seismic attributes were extracted from the pseudo 3D seismic volume (Fig. 5.11). Table 5.3 Seismic attributes corresponding to reservoir characterizations (Chopra, et al., 2005) Volume Attributes Bedding Continuity Lithology, Porosity Structure Thickness Envelope Instantaneous Phase Instantaneous Frequency Average Weighted Frequency Semblance Waveform Difference Spectral Decomposition Simple Difference Relative Acoustic Impedance Sweetness 97 Seismic facies extraction volume We were able to interactively blend multiple seismic volumes, isolate areas of interest, and then extract what is visualized into a 3D object called a geobody. As the geobody was extracted, the interpreter assigned a geological template to the geobody, assigning the geobody to a geological class providing the body with instant geological meaning. Geobodies could be included directly in the 3D geological model, bridging the gap between geophysics and geology (Petrel® Manual 2009). The histogram of the “amplitude envelope” attribute reveals the distribution of the geobodies inside the volume. This is achieved by visual inspection and interactive filtering of the rendered volume (Fig. 5.8a, b). The resulting geobody representing the deltaic environment in the study area is shown in figure 5.8c. We can observe the geobody/facies is close to the concept of the deltaic environment outlined in chapter 2. The main direction of channel system in this interval from northwest to southeast with one delta lobe being very clearly imaged in the west part. In figure 5.8c we show that the G3 group of wells (Z76, Z81, Z82) and G4 group of wells (Z78, Z79) were drilled in the west part of the study area in an area having a very complex facies distribution. Wells of G2 group (Z75, Z84) were drilled in the channel system (middle part) and wells of G1 group (Z72, Z74, Z77) were drilled in the east part of the study area where there is the delta lobe facies. The well Z72 is at the extreme edge of the delta lobe. Due to low vertical resolution of the pseudo 3D seismic the result achieved using this methodology provided a simple picture of the facies distribution in the study area. We could not clearly differentiate the facies distribution in the west part of the study area. In addition the distribution of the six rock types (chapter 4) was sub optimal. In order to improve discrimination within the facies model we apply Unsupervised Neural Network (UNN) for seismic facies classification. 98 Data filter (a) (b) Delta plain Channel system Delta lobe Fault (?) (c) Fig. 5.8 Channel system and delta lobe (geobody) facies from pseudo 3D seismic extraction a) envelope histogram before filtering, b) envelope histogram after filtering, c) model of deltaic environment in the study area. 99 5.3.2.2 Seismic facies classification Artificial Neural Network (ANN) is becoming more popular as a tool utilized in the field of reservoir characterization. The neural network techniques are applied for well log and seismic facies classification and characterization of reservoir properties. For seismic facies analysis Unsupervised Neural Network (UNN) is most commonly used. This technique does not require prior knowledge of object to be classified and is based on pattern recognition and a self-organization program which decides how to separate some attributes into classes. The neural network program then classifies the data, based on the amplitude and shape characteristics of the attributes that are input. Supervised Neural Networks (SNN) are applied for classification of facies and for predicting reservoir properties (porosity, permeability) from the combination of various samples based on seismic attributes including acoustic impedance and well log data. In this approach well points are considered as training points and the statistical methods are used to derive the relationships between the attributes and well log data. Carrillat and Valles (2005) have showed clearly eight steps for seismic facies classification workflow (Fig. 5.9): 1. generation of the seismic texture attribute cubes, 2. running a classification in an unsupervised mode (A) and inspection of the results, 3. running a supervised classification with user-defined selection of training points (B), and examination of these points in N-dimensional space to ensure adequate clustering and minimal overlap between the training points, 4. evaluation of the supervised classification output, using (i) checking the result in 3D and comparing the results against the seismic/attribute cubes used and, (ii) the uncertainty analysis of the classification with probability cubes, 5. the removal/incorporation of training points/seismic attribute cubes prior to making a revised classification, 6. final tuning of the neural network parameters, 7. visualization and interpretation of the data and the analysis of facies associations and structures in 3D space, 100 8. finally, the calibration of the results against well data allows establishing a deterministic link between seismic facies and lithology. In this study, due to the paucity of wells for training data and limitations in the quality of the pseudo 3D seismic we did not observed a good correlation between seismic attributes and log data. Therefore we applied the Unsupervised Neural Network method for seismic facies classification to improve the results from this area of investigation. UNN technology is shown in figure 5.10 where we input the seismic attributes for UNN processing, the output result being the facies classes. In order to reduce noise in the seismic data and also to focus on the reservoir interval (between H1 and H4), all of the attributes were sampled into the 3D grid. After testing UNN with several seismic attributes we identified four attributes that correlated well with facies: these are Amplitude (raw data), Relative acoustic impedance, Envelope, and Sweetness (Fig. 5.11). There are different attributes here and in Fig. 5.11. Fig. 5.9 The workflow for seismic facies classification using ANN method (Carrillat and Valles, 2005) 101 Attributes input UNN processing Facies classes Fig. 5.10 Application of the Unsupervised Neural Network technology for seismic facies classification Relative acoustic impendence Sweetness Envelope Root mean square (RMS) Fig. 5.11 Seismic attributes generated from pseudo 3D seismic amplitude for UNN processing 102 Because we used the UNN method for the facies classification, the main consideration in choosing the number of classes/facies was focused to balance the level of real geological units with the statistical outcomes. Ideally, each chosen facies should be geologically significant and yet we need to have enough data to allow reliable inference of the required statistics for reservoir modeling. In practice, it is difficult to support more than four different facies from the data. The facies must have clearly different petrophysical properties or spatial features that make them easy to model. There is no benefit to separate the data according to facies that do not lead to distinguishable flow properties (Clayton, 2002). Now, we met the same problem discussed in section 4.4 - to overcome limitation of unsupervised clustering method - unknown number of classes for the primary input. The same problem in the K-mean clustering method had been discussed when we derived HU from well log data. At the first step we divided the data into 10 classes/facies for UNN processing to recognize a maximum number of classes/facies in the data. The names and percentage (%) of data of each facies are shown in figure 5.12a and the 3D model result named 10Facies_UNN is displayed in figure 5.13a. In the next step, on the basis of the histogram of 10 facies (Fig. 5.12a) we decided to reduce the number of facies to 6 facies (Fig. 5.12b) to secure correspondence with 6 Rock Types from log data. In this case, in a very simple way we could combine the first four facies (0 + 1 + 2 + 3) with smaller percentage (%) into the facies 4 and keep the same next five facies with higher percentage (%). Figures 5.12c, d, e, f showed the crossplots between two seismic attributes corresponding with the colors of 10 and 6 facies. At the final step, in order to reduce the influence of noise in the pseudo 3D seismic data we smoothed the 6Facies_UNN model. The 6Facies_UNN model after smoothing is displayed in figure 5.13b. This result showed clearly the delta lobe and the channels system that was similar to the geobody facies model in figure 5.8. But this model shows more details for each facies distribution, both laterally and vertically. A comparison between 10Facies_UNN model and 6Faceis_UNN model in cross sections cut through 3 wells (Z81 – Z76 – Z82) from upper to lower part is also presented in figure 5.13. 103 5.3.2.3 Rock Type modeling Seismic data are often ideal for constraining numerical facies models of object based or training image based algorithms. The large scale information provided by seismic correlates better with large scale facies variations (proportions) than with fine scale petrophysical properties. Directly constraining porosity model to seismic is difficult because porosity measurements from cores or well logs often show a low collocated correlation with the seismic data. Before applying a geostatistical algorithm for constraining facies models to seismic, a calibration of the information content of seismic on facies presence is obtained. From the 3D seismic data, for calibration purposes, one retains only seismic observations made along the well paths that contain logs or have been cored. Next, one estimates the probabilistic relationship between facies presence and seismic data from the pair of seismic and well data (Caers, 2005). Seismic waves respond to large scale geological events within the reservoir. Geological events can be seen directly from seismic attributes. The result outcome using seismic extraction volume and seismic facies classification in this study has proved high usefulness for rock type modeling. The 6Facies_UNN (Fig. 5.13b) showed very good distribution close to our understanding of deltaic depositional environment. However, it was built automatically and only on the base of pseudo 3D seismic data. Manual interpretation may improve this model. In figure 5.13b we observe the lateral facies distribution which we assume as good for horizontal (X and Y directions) scale because of the relatively high density of the 2D seismic data. In figure 5.14a, however, we need to consider that for the vertical (Z direction) scale the resolution is not sufficient due to complex geological structure of the thin interbedded sandyshaly layers in the study area. We can easily recognize that if the thickness of facies is smaller than seismic vertical resolution (~10 – 20 m) that facies cannot be displayed in the facies model. Figures 5.14 shows the comparison between seismic based 6Facies_UNN and the 6RT_log results. In the vertical direction the seismic facies have a low resolution on the order of 10’s of meters. Conversely the log based Rock Type facies results has a very high resolution on the order of 1 meter or so. For this reason, to improve the 3D Rock Type model in vertical Z direction, after the seismic facies classification step, we applied the stochastic method Sequential Indicator Simulation (SIS) for Rock Type modeling (the workflow in Fig. 5.5). 104 (a) (b) (c) (d) (e) (f) Fig. 5.12 Reduction from 10Facies_UNN to 6Facies_UNN on the basis of the facies histogram a) 10Facies_UNN histogram, b) 6Facies_UNN histogram, c) Crossplot of Amplitude & Relative acoustic impedance for 10Facies_UNN, d) Crossplot Amplitude & Relative acoustic impedance for 6Facies_UNN, e) Envelope & Relative acoustic impedance for 10Facies_UNN, f) Envelope & Relative acoustic impedance for 6Facies_UNN. 105 (a) Delta plain Delta lobe Mouth bar (b) Fig. 5.13 3D Facies classification using UNN method a) 3D 10Facies_UNN and cross section b) 3D 6Facies_UNN and cross section 106 The SIS is a kriging-based stochastic method mentioned in section 5.1.3. Application of that method to the rock type distribution provided a result which honored at the well data at each well location. In order to improve in horizontal resolution caused by uneven spatial distribution of wells the 6Facies_UNN is used for rock type conditioning in this step. The resulting model 6RT_SIS is shown in Fig. 5.16a. The cross section in the figure 5.14b shows a good match between the 6RT_SIS model and the 6RT_log model. For the purpose of comparison in three dimensions the 6Facies_UNN model was filtered to display only facies F8 which imaged delta lobe distribution very clearly. In addition, the 6RT_SIS model was filtered to display 3 rock types (RT7, RT8, RT9). This showed more detail in the Rock Type distribution and while honoring the log data at the respective well locations (Fig. 5.15). The 4 rock type maps responding to 4 surfaces (H1, H2, H3 and H4) show that rock type distribution varies with depth as shown in figure 5.17. 5.3.3 Hydraulic Flow Unit modeling constrained by Rock Type model Six hydraulic flow units (6HUs) have been defined in the study area (Chapter 4) and upscaled into the 3D grid as facies data. Once again, the Sequence Indicator Simulation (SIS) method was applied for 3D HFU modeling. In order to control distribution of HU, the 6RT_SIS is used for constraining the HU model. The resulting model 6HU_SIS is shown in Figure 5.16b. A comparison between 6RT_SIS and 6HU_SIS in 3D window is shown in figure 5.16. On the different surfaces (H1, H2, H3 and H4) in figure 5.17 we can see that edges of the rock type and hydraulic flow units do not match exactly but the overall depositional trend is similar. The trend of rock type 10 (RT10/shale) is close that of to HU1 and the trend of the good reservoir units (HU6, HU5) are close to the trend of facies RT7, RT8. 107 108 (d) (b) Matching with RT_log Not Matching with RT_log Fig. 5.14 Comparison 6Facies_UNN and 6RT_SIS; a) 6Facies_UNN and 6RT_log, b) 6Facies_UNN histogram, c) 6RT_SIS and 6RT_log, d) 6RT_SIS (blue) and 6TR_log (green) histogram. (c) (a) (a) (b) Fig. 5.15 Comparison between seismic facies classification (6Facies_UNN) and 3D rock type model (6RT_SIS) a) 6Facies_UNN model filtered to display only faceis F8_UNN showing clearly delta lobe distribution in the study area, b) 6RT_SIS model filtered to display 3 Rock Types (RT7, RT8, RT9) showing in more details rock type distribution and honoring wells location. 109 (a) (b) Fig. 5.16 Comparison between the 6RT_SIS and 6HU_SIS a) 3D 6RT_SIS, b) 3D 6HU_SIS 110 6RT_SIS 6HU_SIS (H1) (H2) (H3) (H4) Fig. 5.17 Comparison of 6RT_SIS and 6HU_SIS in 4 horizons: H1, H2, H3 and H4 111 5.4 Property modeling constrained by HU model The final step of static modeling is property modeling (Fig. 5.5). As discussed in section 5.1.3, Sequence Gaussian Simulation (SGS) is commonly used in property modeling. In this study, 3D porosity (3D PHI) models are generated corresponding to the six hydraulic flow units model (6HU_SIS) using the SGS method. Due to very complex variation of permeability in the reservoir model we should first model porosity and then permeability. Generally, permeability modeling will be done using SGS and co-kriging with the porosity model. In this project with advantages of applying the hydraulic flow unit method we can directly calculate the 3D permeability model from the 3D porosity model using Kozeny- Carman equation 3.8 in Chapter 3 (Ha Quang and Jarzyna, 2011a,b). 5.4.1 Porosity and permeability modeling a) Porosity modeling without any conditions or constraints Porosity was calculated on the basis of logs from 7 wells and then upscaled into the 3D grid. The SGS method was applied to interpolate porosity into the 3D model, without any conditions or constraints so the result was strongly influenced by variogram parameters. b) Porosity modeling constrained by 3D Hydraulic Flow Unit model (PHI_HU) Following the workflow in this study (Fig. 5.5), the final porosity model was constrained by the 3D Hydraulic Flow Unit model using the SGS method and the same variogram parameters (Table 5.1). 3D porosity model (PHI_HU) is shown in figure 5.18a where porosity distribution responds to 6 HUs, high porosity values correlate with HU6s, HU5s and HU4s and low porosity with HU1s, HU2s and HU3s. c) Permeability calculation using Kozeny-Carman equation (K_HU) In the 3D model we also applied equation 3.8 to calculate permeability model (K_HU) on the basis of 3D 6HU_SIS (Fig. 5.16b), 3D PHI_HU (Fig. 5.18a) and FZI_mean (Table 3.2). The result is shown in figure 5.18b and distributions of PHI_HU and K_HU are presented in figure 5.19. The cross plot of 3D PHI_HU vs. 3D K_HU shows very clearly the relationship between K and PHI corresponding to each HU (Fig. 5.20). The picture in this figure can be compared with the plot in figure 3.7 (Chapter 3). There is observed correlation between permeability from logs (K_log) and permeability calculated for 3D model (Fig. 5.21). We can see that permeability of HU1 is slightly smaller than K_log but in other HUs the correlation is excellent. 112 (a) (b) Fig. 5.18 Comparison between 3D PHI_HU model and the 3D K_HU model a) porosity model constrained by HU model (PHI_HU) b) permeability model calculated from the porosity model using Kozeny-Carman equation (K_HU) 113 Porosity Permeability (H1) (H2) (H3) (H4) Fig. 5.19 Comparison between PHI_HU model and the K_HU model; results presented on various surfaces 114 Fig. 5.20 Cross plot of 3D PHI_HU versus 3D K_HU; shown very clear relationship between K and PHI responding to each HU Fig. 5.21 Cross plot between permeability from logs (K_log) and permeability calculated for 3D_K_HU model from the mean values of FZI for HU 115 5.4.2 Water saturation and Net to Gross modeling a) Water saturation modeling An improved technique for modeling the initial reservoir hydrocarbon saturation is presented. In contrast to the Leverett J- function approach (5.3), this methodology (hereby termed flowunit-derived initial oil saturation or FUSOI) determines the distribution of the initial oil saturation from a measure of the mean hydraulic radius, referred to the flow zone indicator (FZI). In the FUSOI approach, capillary pressure parameters (Pc), irreducible water saturation (Swir), pore-entry pressure (Pd), and pore-size distribution index (, derived from the Brooks and Corey (1966) model (Eq. 5.4), are correlated to the FZI. Subsequent applications of these parameters then permit the computation of improved hydrocarbon saturations as functions of FZI and height above the free water level (FWL). This technique has been successfully applied in the Mississippian Aux Vases Sandstone reservoirs of the Illinois Basin (USA) (Udegbunam & Amaefule, 1996). Amaefule (1995) showed that the Leverett J function should only be used to correlate capillary pressure data for samples from the same flow unit (Fig. 5.22). The Pc and Sw data are very satisfactorily matched by Brooks and Corey’s equation and the best correlations of Swir, and Pd are shown as equation 5.5, 5.6 and 5.7 (Udegbunam & Amaefule, 1996). However, in this particular gas reservoir, due to lack of data for reservoir rock properties and uncertainty about the free water level (FWL), in order to correlate Sw to the hydraulic flow units, one again, we used the SGS method for water saturation modeling and constrained it by the 6HU_SIS model. The Sw model was shown in figure 5.23 made it easy to calculate gas saturation model (Sg) by equation 5.8. 116 Sg = 1 – Sw (5.8) Where: J(Sw) = Leverett J function (dimensionless), Sw = water saturation for a given capillary pressure, Swir = irreducible water saturation, Pc = capillary pressure at any height above the free water level, Pd = pore-entry pressure, pore-size distribution index, = contact angle, = interfacial tension. Fig. 5.22 Typical capillary pressure curves for different flow units fitted with the Brooks and Corey’s equation (Udegbunam & Amaefule, 1996) 117 Fig. 5.23 Water saturation model constrained by Hydraulic Flow Unit model (6HU_SIS) b) The 3D Net to Gross model (NTG) The principal use of cut-offs is to delineate net pay, which can be broadly described as the summation of those depth intervals through which hydrocarbons are (economically) producible. In the context of integrated reservoir studies, net pay has an important role to play both directly and through a net-to-gross pay ratio. Net pay demarcates those intervals that are the focus of the reservoir study. It defines an effective flow thickness that is pertinent to the identification of flow units, that identifies target intervals for well completions and stimulation programs, and that is needed to estimate permeability through the analysis of well test data. The net to-gross pay ratio is input directly to volumetric computations of hydrocarbons in place and then to "static" estimates of reserves. It is a key indicator of hydrocarbon connectivity, and it contributes to the initializing of a reservoir simulator and then to "dynamic" estimates of reserves. Unfortunately, there is no universal definition of net pay nor is there general agreement on how it should be delineated. For this reason, net pay has been incorporated within integrated reservoir studies in many different ways that have not always been fit for purpose. In particular, there is no generally accepted method for 118 quantifying net pay cut-offs, without which net pay cannot be delineated (Worthington and Cosentino, 2003). Normally, in sandstone reservoir we can use Vsh, Sw, PHI and K from logs data for reservoir cut-off values. In this study with some advantages of applying HU method we directly calculated 3D NTG model on the basis of 3D HU model. Histograms of 6HUs and NTG models are presented in figure 5.24 and parameters used to calculate 3D NTG (Fig. 5.25) are shown in table 5.4. Table 5.4 The values cut-off reservoir basic on 6 HUs 6HUs Cutoff value for 3D NTG HU1 HU2 HU3 HU4 HU5 HU6 0 0.4 0.5 0.7 0.8 1 6HUs Histogram Net to gross histogram Fig. 5.24 Comparison between 6HUs histogram and Net to Gross (NTG) histogram 119 Fig. 5.25 Net to gross model based on 3D HU model 5.5 Conclusions In this chapter one of possible methodologies for incorporating 2D seismic information and well logs in stochastic simulations of rock type modeling was presented. A main point in this approach is the ability to simulate the 3D rock type at a fine scale while accounting for larger scale 2D seismic derived information and then constrain to 3D HU modeling in the study area. Geostatistical modeling using the Sequence Indicator Simulation (SIS) and the Sequence Gaussian Simulation (SGS) methods were used in this study for rock type modeling. Specially, with high density distribution of 2D seismic data, the stochastic method (SGS) was a good choice to convert the 2D seismic survey into pseudo 3D seismic cube. The geobodies that were extracted from the pseudo 3D seismic show a simple picture of deltaic facies distribution and the channel systems prograding from northwest to southeast. Applying this solution, the results mainly depended on the density of 2D seismic lines and also quality of seismic data. 120 Seismic traces contain information about facies changes and contrasts. Cluster analysis was the approach that had potential to classify seismic trace shapes into meaningful facies. In this study the Unsupervised Neural Network (UNN) was used to classify the pseudo 3D seismic attributes into six facies (6Facies_UNN). In order to improve result seismic attributes were resampled into the 3D grid before clustering and the final facies model also needed to be smoothed. Applying deterministic-statistic (seismic extraction volume, UNN) and stochastic (SIS) methods and then combining them we improved Rock Type modeling. The result (6RT_SIS) showed better rock type distribution in lateral (X, Y) and vertical (Z) directions and then the model was used to constrain the hydraulic flow unit modeling. Comparing the rock type model and HU model we observed that although the borders did not match the trends of HU were similar to the rock type distribution. It means that the 3D HU model was not only dependent on reservoir properties (K, PHI) but also was controlled by paleostratigraphy (facies or rock type). The porosity model was constrained by the 6HU_SIS model obtained using SGS method. The PHI_HU model showed porosity distribution corresponding to 6HUs, with higher porosity value in HU6, HU5 and HU4 and lower porosity in HU1, HU2 and HU3. The 3D permeability model (K_HU) was calculated using Kozeny-Carman equation on the basis of 3D PHI_HU and FZI_mean of each HU. In each HU a good correlation between permeability log (K_log) and 3D permeability model (K_HU) was clearly visible which confirmed the advantage of applying the HU method in this study. Due to lack of SCAL data for reservoir rock parameters the 3D NTG model was calculated directly from 3D HU and it again showed the advantages of HU methods. The static models created in this chapter will be upscaled for reservoir simulation in the next chapter. 121 Chapter 6 HISTORY MATCHING UNDER HYDRAULIC FLOW UNIT CONTROL Static modeling presented in the previous chapters provides an effective method to characterize reservoirs using cores, logs, and seismic data. However, they do not account for the dynamic data such as pressure, and oil and gas production, which are an important goal of the petroleum industry. Engineering decisions cannot be based on static models that do not match historical production data. Historical production data provide a direct observation of the ultimate modeling goal: reservoir flow performance. Any method for integrating production data into 3D static models will call upon a reservoir simulator or dynamic modeling. Theoretically, an integration of additional data into reservoir model should lead to a more realistic model with reduced uncertainty range. In reality however, flow simulation requires the knowledge of many additional reservoir parameters that sometimes simply are not available. In such situations flow simulation becomes closer to an art rather than an engineering science. In this chapter, the static models in Chapter 5 need to be constrained to dynamic data obtained from historical production data. To achieve this, we used the ECLIPSE® 100 software for history matching. The workflow used was as follows: up scaling of the Petrel® geological model to a reservoir model for input to Eclipse®, specifying the reservoir initial condition, running the model to extract reservoir performance data on history matching control by HU model, discussion. Due to lack of initial condition data in the reservoir interval we used some default parameters in Petrel® software for simulation case such as relative permeability (Kg, Kw). Before discussing history matching under HU control, some main points of history matching will first be reviewed. 122 6.1. History matching under Hydraulic Flow Unit control 6.1.1 History Matching Overview Ideally, reservoir models should match the observed dynamic behavior of the reservoir within some accepted tolerance. To check the model's consistency with dynamic data, flow simulation is required. However, flow simulation is CPU-demanding and is only rarely applied directly to the high-resolution geocellular model. Some form of upscaling, also termed model coarsening, is required to reduce the number of grid cells and make flow simulation feasible. Even if upscaling were not required, the geocellular model built solely from well-log and seismic data rarely matches the production data. The process of adjusting/perturbing an initial reservoir model to match production data is commonly known as history matching. The history matching step is iterative, whereby the initial model is perturbed many times, either manually or automatically, to achieve a satisfactory history match (Caers, 2005). The purpose of history matching is not just to match historical results to current results, but rather to produce models that can be used to forecast reservoir performance within some accepted tolerance. The goal is to produce reservoir models that have an improved prediction power over models that do not match history. It is the combination of all data matching, including production data plus a correct geological model that makes a good reservoir model. Because there is no objective measure to gauge the degree by which a model will predict future performance accurately, the best one can be done is to build reservoir models that match all relevant reservoir data, including production data, and reflect as realistically as possible the subsurface heterogeneity. History matching, therefore, should never be an isolated task left to the sole ability of the reservoir engineer. Instead, production data should be considered as one additional, important piece of information in model building. One should take care in integrating this information jointly with all other relevant data. History matching is called for at various stages of the reservoir modeling task depending on the amount of production data available. In general, there are two main stages: 123 Stage 1: Adjusting Large/Field-Scale Structural Elements At this stage, one is merely interested in determining the large-scale "plumbing" of the reservoir correctly. Usually, this involves adjusting the following elements of a reservoir model: position and transmissibility of the main faults, depth and strength of the fluid contacts (water/oil/gas), overall reservoir pore volume, fluid properties [pressure/volume/temperature (PVT)] and rock compressibility, for all compartments: average permeability, porosity, and saturations, relative permeability curves. These properties are often adjusted manually or interactively. Automatic procedures are rarely applied at this stage. Stage 2: Adjusting Local Reservoir Properties Stage 1 may provide a good overall field match (total production, rates, and pressure), but it may not provide a satisfactory local, well-by-well match. This would involve the following adjustments: variable transmissibilities along the fault surfaces, local facies proportions and their spatial distribution, including relative permeability curves per facies, porosity and permeability distributions within facies. Before any history matching exercise, whether manual or automatic, a sensitivity study investigating which component (global or local) of the reservoir model has greatest influence on production data is required. Regardless of the efficiency of any particular history matching method, if one does not perturb the components ranked as most sensitive to production data, a history match may never be achieved. Such a study may also expose the shortcomings of the initial geocellular model. For example, if an inappropriate variogram based approach for modeling permeability is used in a strongly channeled reservoir, then regardless of how much one perturbs that variogram, a history match may never be achieved. Sensitivity can be evaluated from sound engineering expertise and insight into the particular situation at hand; that in sight can be confirmed by means of experimental design-type studies. 124 6. 1.2. History matching under Hydraulic Flow Unit control Traditional method: Normally, a two stage approach is used to produce consistent petrophysical properties, i.e., porosity and permeability, with the underlying geological description. This goal is achieved through two independent processes, which are later combined to produce the final result by a filtering process. The first process is the simulation of geological description. The second process is the simulation of porosity. The permeability description can be obtained through correlation between porosity and logarithm of permeability once the final porosity description is produced. To combine the results of these two processes, i.e., geological and porosity simulations, a filtering process is applied. The filtered porosity at a grid block is obtained by first examining the type of geological unit at that location and then selecting the porosity value from the corresponding realization. Once the filtered porosity is known, the permeability value at that location can be determined using the suitable correlation. This approach has produced consistent result, between the petrophysical properties and the underlying geological description (Kelkar al el., 1997) (Figure 6.1). Fig. 6.1 Schematic diagram of two stages approach (Kelkar al el., 1997) History matching under Hydraulic Flow Unit control: Figure 6.2 shows in more detail the four main stages used to integrate from geophysical to geological data into static models. Note that for each HU we have good correlation between porosity and permeability (Fig. 6.3). The final results will be upscaled and input to Eclipse® for 125 126 Fourth stage Fig. 6.2 The four main stages to integrate geophysical and geological data into reservoir modeling with HU method in the study (modified from Kelkar al el., 1997) Third stage KHU6 HU6 HU6 RT10 F10 Second stage KHU5 HU5 HU5 RT9 F9 First stage KHU4 HU4 HU4 RT8 F8 KHU3 HU3 RT7 F7 HU3 KHU2 &K Correlation HU2 Filtering by SGS method HU2 Filtered Permeability RT6 Filtered porosity F6 1 2 3 4 5 6 Porosity simulation for each HU KHU1 HUs HU1 HU1 HU2 HU3 HU4 HU5 HU6 HU simulation for each Rock type HU1 Geological simulation RT5 RT5 RT6 RT7 RT8 RT9 RT10 Rock type simulation F5 Geophysical simulation flow simulation. The workflow of history matching is controlled by hydraulic flow unit method displayed in figure 6.4. The procedure workflow is as follows: 1. data input: static model (PHI, K, NTG), PVT data, and initial reservoir condition, and historical production data, 2. flow simulation using program ECLIPSE® 100, 3. history matching criteria, which either require some adjustment of the model parameters, 4. model adjustment includes: change of size grid upscaling, relative permeability function(s), Net to Gross cut off, 5. Production forecast for gas volume production. (b) 10000 1000 K [mD] 100 10 HU1 HU2 HU3 HU4 HU5 HU6 1 0.1 0.01 0.05 0.1 0.15 0.2 0.25 0.3 (c) 0.35 PHI [fraction] (a) Fig. 6.3 Hydraulic Flow Unit control a) Cross plot K vs. PHI responding to 6HUs (core data), b) 6HUs section of 3 wells (Z81, Z76 and Z82), c) Permeability section of 3 wells (Z81, Z76 and Z82). 127 core Logs Seis. HU PHI Production & PVT data NTG K ECLIPSE Adjust Model HISTORY MATCHING No Yes Production Forecast Fig. 6.4 The workflow of history matching under Hydraulic Flow Unit control (modified from Mikhail, 1997) 128 6.2. Upscaling Many reservoir flow simulators cannot directly and effectively handle the size of grids used in geological models. Such models can easily contain as many as 10 million cells, whereas single CPU simulations will only run in reasonable time with models of the order of 100,000 cells. Furthermore, grids used in geological models are often unsuitable for simulation due to geometric problems such as inside-out cells. Upscaling is the process of creating a coarser (lower resolution) grid based on the geological grid which is more appropriate for simulation. While this necessitates the omission of much of the geological models fine detail, the result is intended to preserve representative simulation behavior. In Petrel®, upscaling is split into two steps (Petrel® 2010, Manual): Scale up Structure: define the new layering scheme (numbers and shapes of layers) of the simulation grid, Scale up Properties: populate grid properties, such as porosity and permeability, based on those in the fine grid. Scale up Structure: Table 6.1 shows the parameters of scale up grid structure. The number of cells must therefore be reduced by up scaling structure. In this study, we tested many grids and a bigger grid with bock sizes of 300x300x99 gave better results and also increased the speed of simulation in Eclipse®. Table 6.1 Upscaling structure models for Eclipse® simulation Models Grid parameters Grid size Total cells Fine scale 50x50x189 267x169x198 8934354 Model 1 100x100x189 133x84x198 2212056 Model 2 200x200x99 66x42x99 274428 Model 3 300x300x99 44x28x99 121968 129 Scale up Properties: Properties from one grid can be transferred to another grid of a different resolution or orientation using the Scale up properties process. This is usually done in the context of building a simulation model from a geological model, where the simulation model has been coarsened and reoriented for flow simulation. Some quantities, such as porosity and water saturation are easy to upscale, because they may be averaged arithmetically. Averaging can be based on volume or on number of cells. Volume-weighted averaging will weight source property values by the volume contributed by the fine cell, whereas cell count averaging will give equal weight to all source cells involved in the average (regardless of variations in their volumes). Cell count averaging is not available when using the all intersecting cells sampling method because a fine grid cell will contribute to multiple coarse grid cells when using this sampling method, so volume-weighting must be used to ensure its overall contribution respects fine grid volumes (Petrel® 2010, Manual). In figures 6.7 a, b, c, d there are shown 3D_HU and 3D_PHI upscalings into grid 300x300x99 (Model3 – Table 6.1). However, the 3D permeability model (3D_K) is much more difficult to upscale. There are many papers discussing how to make an optimal upscaled permeability model for reservoir simulation. In this study, due to complex HU distribution, we used the Flow Based Tensor Upscaling method for upscaling permeability with some of the input parameters shown in figure 6.5. A comparison between fine scale and the result after downscaling of permeability is show in figures 6.6 and 6.7. 130 Fig. 6.5 The Process Dialog for Flow-Based Tensor Upscaling Flow-based upscaling involves performing a numerical pressure simulation on the block of fine cells coinciding with each coarse cell to determine a representative coarse cell permeability. The process will calculate I, J, and K or X, Y, and Z permeabilities from input as permeability in the I, J and K directions, net-to-gross and porosity. If the Output full tensor permeabilities option is used, then off-diagonal terms in the permeability tensor (in the IJ, IK and JK directions) will also be calculated. It is assumed that the system is symmetrical, i.e. IJ = JI. Tensor upscaling methods always include a fine grid cell in its entirety (or not at all) in the calculation of each coarse grid cell value. As such, small intersections can cause more cells than expected to be incorporated, particularly when using the all intersecting cells sampling method. When doing tensor upscaling between closely aligned grids, always use the cells with center inside target cell or Zone-mapped layers sampling methods to avoid unexpected "grazing" intersections (Petrel® 2010, Manual). 131 (1) (2) (3) (4) (5) Fig. 6.6 Comparison of scale of permeability in well Z76 (1) Log scale, (2) Log upscaled in 189 layers (static model scale), (3) Ki, (4) Kj, (5) Kk - permeability downscaled in 99 layers by Flow Based Tensor Upscaling method (dynamic scale). 132 Static models Dynamic models (a) HU (50x50x198) (b) HU (300x300x99) (c) PHI (50x50x198) (d) PHI (300x300x99) (e) K (50x50x198) (f) K (300x300x99) Fig. 6.7 Comparison of static scale results (a, c, e) and dynamic scale results (b, d, f) 133 6.3 Reservoir initial condition The conditions in the reservoir that the simulator used to calculate the pressure and phase saturations in every grid block during initialization should be defined. Each fluid region in the reservoir may contain a number of different, unconnected initial condition regions. For each region we must specify a reference depth and corresponding pressure, gas-oil contact depth and water contact depth that depending on which phases in the reservoir. Make fluid model: Fluid models are used by the simulator to define how physical properties of the fluid such as density and viscosity vary with pressure and temperature. Fluid models may also define how the initial conditions in the simulator are to be calculated, by specifying the fluid contacts, pressure, and compositional variation with depth. In the reservoir study, due to lack of initial condition data we can take the average value of initial pressure and temperature measured in some wells from Table 6.2 - Pi = 60 bar and Ti = 35oC and also unclearly information for a gas-water contact (GWC); so we assumed that it is below 900m (almost below reservoir section). The gas components taken from the report (Geological documentation of Z-L gas field) are shown in Table 6.3. Table 6.2 Pressure and temperature measurement results (Geological documentation of Z-L gas field) Wells Z74A Z74A Z74A Z74A Z74A Z74A Z74A Z74A Z75 Z75 Z76 Z76 Z76 Z76 Z76 Z76 Z76 Z76 Z82 Z82 Depth [m] 650 552.5 873.5 734 783.5 765.5 734 783.5 778.5 617.5 739.5 702.5 686 646.5 612.5 593 560 563 549.5 487 Pressure [bar] 62.8 54.6 89.2 76.3 82 79.9 46.7 51.3 76.9 61.2 70.3 67.9 64.9 61.1 56.5 55.2 52.6 35.7 51.7 37.7 Temperature [oC] 33.65 34.75 34.85 32.85 30.85 30.05 - 134 Table 6.3 Gas components (Geological documentation of Z-L gas field) Components CH4 C2H6 C3+ C02 N2 3 H2S[mg ] He[%] [%] 96.98 0.109 0.015 0.000 2.896 0.000 0.000 Making rock physics functions: Petrel® software includes some functions of saturation or pressure used in simulation that represent the physical parameters of fluids and rocks, or interaction between rocks and fluids. Saturation functions are in tables showing relative permeability and capillary pressure versus saturation. These tables are used to calculate: the initial saturation for each phase in each cell, the initial transition zone saturation of each phase, fluid mobility to solve the flow equations. Unfortunately, without lab data for the relative permeability and capillary pressure curves data we had to use default values for saturation function. Figure 6.8 shows gas and water relative permeability in sand and shaly-sand (Petrel® 2010, Manual). Fig. 6.8 Gas (Kg) and water (Kw) relative permeability curves in sand and shaly-sand reservoir (Petrel® 2010, Manual) 135 Rock compaction functions are in tables showing pore volume multipliers versus pressure, or a single rock compressibility value used by the simulator to calculate the p ore volume change. Creating rock compaction functions also creates a transmission multiplier versus pressure curves. Figure 6.9 shows parameters of rock compaction function where we also used porosity and HU model. In this case, we assume that all cells with the similar properties (PHI) form a region of the grid, and a separate rock compaction function is created for each hydraulic flow unit. Fig. 6.9 Rock compaction function responding with porosity model and each hydraulic flow unit 136 6.4 History matching and discussions Historical production data (observed data) In this study for history matching simulation only 3 wells (Z74A and Z75 and Z76) which have gas flow production between H1 and H4 horizon were used. The perforation intervals and time starting produce and status of each well are presented in Table 6.4. Table 6.4 Perforation intervals and time starting produce and status of Z74A, Z75, and Z76 wells Perforation (m) Time starting produce Status Top Bottom Thickness [DD/MM/YYYY] (10/01/2010) Z74Ad 781 786 5 09/01/2006 production Z74Ag 723 744 21 09/01/2006 shut-in Z75d 656 660 4 06/01/2009 shut-in Z75g 541 571 30 06/01/2009 shut-in Z76 560 566 6 12/01/2008 production Wells “g” means upper completion, “d” means lower completion The observed historical gas flow rate and bottom hole pressure (BHP) data are shown in figure 6.9 and Table 6.5 (Appendix A). In order to show more clearly advantages of applying HU method for reservoir simulation we also run Eclipse® with static models by using traditional method. 6.4.1 Results Traditional method with two stages (Fig. 6.1): Following the work flow in figure 6.1 means that we only use facies model constrains to property models (K, PHI) and then make up scaling of the data and input them into Eclipse® for history matching. The results of history matching simulation in the form of gas flow rate and bottom hole pressure, BHP, are displayed in figure 6.11. HU method with four stage (Fig. 6.2): After testing history matching for several various grid sizes (table 6.1), NTG models, and some functions of back-oil fluid model (PVT) we decided to run the simulation with main parameters as follows: 1. grid: 3D grid size (300x300x99) with: Ki, Kj, Kk, PHI_HU and NTG, 137 2. functions: Back-oil fluid model (PVT) controlled by hydraulic flow units model, 3. strategies: Development strategy from 09/01/2006 to 1/10/2010 with two option of production control mode: Reservoir volume and Gas control. The result of the first case simulation with reservoir volume control is shown in figure 6.12 and the results of the second case with gas control are presented in figure 6.13. In both cases of simulation for 3 wells we see that the gas flow rate production was well matched with historical production data (Figs 6.12a, 6.13a) and the BHP simulation is higher than observed data. Only in the Z74Ag well at lower part of perforations the results of simulation are closed to observed data (Figs 6.12b and 6.13b). 6.4.2 Discussion Comparison of simulation results between traditional method and applying HU method in figures 6.11 and 6.12 shows that HU method give better results than traditional method. Especially BHP results in HU method are close to observed data even in this case parameters as relative permeability (Kg, Kw) and rock compression were used as default ones. We also considered the following reasons as influencing the history matching results: from 3 wells used as input to Eclipse® simulation only two wells - Z75 and Z76 were included to static modeling (Chapter 5); the Z74A well (close to well Z74) was excluded so we can use this well for validation of static model in history matching, in the reservoir interval, due to very thin and interbedded shaly - sandy layers we observe high water saturation from well log data, somewhere over 80%, and the gas water contact (GWC) was also unclear, so we decided exclude water saturation model in Eclipse® simulation. 138 (a) (b) Fig. 6.10 Historical production data for 3 wells: Z74A, Z75 and Z76 a) Gas Flowrate, b) Bottom Hole Pressure (BHP) 139 (a) Gas Flowrate (b) Bottom Hole Pressure Fig. 6.11 History matching results (line colors) and observed data (dot points) of traditional method 140 (a) Gas Flowrate (b) Bottom Hole Pressure Fig. 6.12 History matching results (line colors) and observed data (dot points) by HU method with reservoir volume control 141 (a) Gas Flow rate (b) Bottom Hole Pressure Fig. 6.13 History matching results (line colors) and observed data (dot points) by HU method with gas control 142 6.5 Conclusions The static models (Chapter 5) were upscaled and simulated using flow simulator Eclipse®. History matching was performed by manually adjusting a few reservoir model parameters through a trial-and-error procedure. Manual history matching run the simulation model for the historical production period and then compared the results with known field performance. After the comparison was made, history matching results of three cases from 3 wells (Z74, Z75 and Z76) showed good results for gas flow rate for a few years production. Comparison of the results between the traditional method (two stages) and HU method (four stages) showed some distinct advantages of the HU method: at each HU we have very good correlation between porosity and permeability that is good for classification of cells in fluid flow for Eclipse® simulation, from 6HUs distribution we can easy control NTG model for Eclipse® simulation cutting off HU1 with lower PHI and K and also reducing number of cells in 3D grid and increasing CPU processing, rock compaction function can respond to porosity model (PHI) and each hydraulic flow unit. 143 Chapter 7 CONCLUSIONS AND RECOMMENDATIONS The main area of concentration for this study is integrating multiple datasets including wireline logs, core data, production data, and other geological and geophysical data for building static 3D reservoir property models. In particular, the concept of using hydraulic flow units was explored in an attempt to simplify and improve the quality of the 3D static model of a gas reservoir. To test this concept, the Eclipse simulation engine was run against the 3D reservoir model. The results of these simulations was compared to actual reservoir performance by history matching. The case study presented shows the methods used to integrate data from various scales; from microscopic (core plugs) through mesoscopic (logs) to megascopic (seismic). Because a 3D seismic survey was not available, a pseudo 3D cube was created from a dense grid of 2D lines. Stochastic and deterministic geostatistical methods were combined using Petrel software to generate static models which were then were used for upscaling to the Eclipse simulation engine. 7.1 Conclusions The case study presented was performed in the Z gas deposit belonging to a group of Miocene gas reservoirs in the northern part of the Carpathian Foredeep of Poland. The Z gas field reservoir depositional environment was deltaic. The water depth was shallow nearshore marine and the sediments were laid down in an area where tidal influences resulted in a very complicated facies distribution. The reservoir is thus comprised of interbeded sandy- shaly layers which not only vary in thickness, but also in lateral extent. Despite the challenges presented by this complex environment of deposition, a range of geophysical, geological, and geostatistical modeling and reservoir simulation techniques was used to reach the following conclusions: Hydraulic flow unit classification using core data: 1. The hydraulic flow unit technique has been developed by oil industry researchers and as such is routinely applied to the problem of identifying reservoir characteristics. This technique has a wide variety of practical field applications to both cored and uncored intervals/wells. In this study, the data from 570 core plugs (PHI_core and K_core) from ten wells was classified into six Hydraulic Flow Units (6HUs) by applying 144 conventional cluster analysis techniques including histogram, probability plot and Ward’s algorithm. Hydraulic flow unit prediction using core and log data: 2. Statistical methods turned out to be useful, flexible, and effective, because they provide the interpreter with tools that deliver acceptable results. Two methods were tested. These were Linear Multiple Regression (LMR) and Alternating Conditional Expectations (ACE). Both methods were used to solve the problem of integrating core and log data to calculate the Flow Zone Indicator (FZI.) The FZI was in turn used to divide the reservoir in each well into Hydraulic Flow Units (HUs). The applicability of the two methods was tested by comparing the correlation coefficient of both sets of theoretical transformations against the actual reservoir parameters. This revealed that the optimal ACE transformations of dependent (FZI) and independent (logs) values improved the correlation of the FZI from core and well log data and the results were superior to those obtained by the LMR algorithm. Rock type classification using log data: 3. Without core facies descriptions, the six Rock Types (RT) were classified by K mean clustering method. After applying techniques to overcome the limitations of this method, the resulting classification was used as background to control the distribution of the HUs. Hydraulic flow unit modeling using well log and 2D seismic data: 4. In recent years the Multiple Point Statistics (MPS) included in Petrel has shown its ability to improve geostatistical modeling techniques. However, in this study the traditional technique using variogram and two-point statistics such as Sequence Indicator Simulation (SIS) and the Sequence Gaussian Simulation (SGS) methods was applied to build a geostatistical model with good results. In particular, a pseudo 3D seismic cube was created from a high density grid of 25 lines of 2D seismic data using SGS high speech computers and the powerful Petrel seismic interpretation software simplified the extraction of geobodies from the pseudo 3D seismic cube. This accelerated the process of mapping the deltaic facies revealed a channel system 145 flowing from northwest to southeast, with general progradation of the delta front to the southeast. 5. Unsupervised Neural Network (UNN) was used to classify the pseudo 3D seismic attributes (RMS, Envelope, and Sweetness) into six seismic facies (6Facies_UNN). Rock type modeling using stochastic methods (SIS) helped to overcome the limitation of the seismic scale by combining well log (6RTs) with seismic facies (6Facies_UNN). 6. Six hydraulic flow units were modeled using the SIS method constrained by the six rock types model. This was helpful in controlling the HUs distribution in three dimensions. This was very significant because of the paucity of well control in the study area. Since the method incorporated the facies distribution extracted from the pseudo 3D cube it meant that the 3D HU model outcome was now not only dependent on petrophysical properties (K, PHI) but also was controlled by paleostratigraphy. 7. To reduce uncertainty in the applied geostatistical methods, and to maximize the many advantages of applying the hydraulic flow unit technique, the 3D permeability model was directly calculated from the 3D porosity model constrained by HU model (using Kozeny-Carman equation). The results show excellent correlation between the permeability as measured from log data and calculated 3D permeability. As a bonus, the 6HUs distribution enabled us to easily generate a net to gross ratio (NTG) model by using cutoff value for each HU. A comparison of deterministic (6Facies_UNN) and stochastic methods (6RT_SIS) shows clearly that the stochastic method gives high resolution results (log scale) which are honored at the well location. Because the stochastic method uses a random number generator to generate equiprobable results away from well control, it is an appropriate method for improving the abstract concept of reserves and overall probability of reservoir characterization. Deterministic methods are more accurate but cannot predict reservoir character away from well control. Both methods have their uses. The stochastic method is an effective predictor for statistical estimation of the reservoir character away from well control, however one should be aware of its limitations. In particular the stochastic approach should not be used to predict the actual reservoir characteristics when considering where to locate a new well. 146 History matching controlled by HU model: 8. The final statistic models (K, PHI and NTG) were upscaled and then input to Eclipse for simulation and history matching. With the HU method controlling the static models, even with lack of reservoir engineering parameters and initial condition data, the results showed a good match for gas flow rate and slightly higher bottom hole pressure (BHP). By comparing history match results with and without HU control we can clearly see that some advantages: for example because the HU’s encapsulate a set of reservoir parameters into discrete package, it is very easy to test the effect of different NTG models and rock compaction factors on each individual HU. 7.2 Recommendations The study workflow shows the advantages and disadvantages of the integrated reservoir analysis approach to combining cores, logs and 2D seismic data for geological modeling which can then be applied to production history matching via Eclipse simulation. However, in order to get better results in static and dynamic modeling using the HU method, the points listed below should be considered: Core data: 1. Core plug porosity and permeability data (K_core, PHI_core) are parameters critically necessary for calculating the FZI that is key to the HU method in the study. However, after core depth matching by using PHI_core and Neutron porosity log (NPHI or PHI) we found that about 30% of core data was unusable because it could not be reliably depth registered. Since core data, very expensive to acquire, is essential to the task at hand, and must be depth matched to be used effectively, we recommend that at a minimum both Gamma ray log (GR_log) and Gamma ray core (GR_core) should always be run. In addition the FMI log should be run across the reservoir, as this gives an excellent false color image of the borehole for direct comparison to the recovered core. 2. No core facies descriptions were available. Even with the advantages of K mean clustering for rock types classification, it is still only an unsupervised learning method without a training data set. Due to complex facies distribution in deltaic environment core facies description is very useful to validate the results. Therefore we recommend to preserve core facies description with the other core material. 147 3. The HU approach in this project was used to perform a complex reservoir characterization. However, due to lack of SCAL data some important parameters such as water saturation, relative permeability, capillary pressure and others were excluded, and therefore we had to use default values in the Eclipse simulation. We strongly recommend that SCAL data be available to improve Eclipse results. Log data: 4. The log data was sampled at 0.25m/point. This coarse sample interval introduced several problems: a. The resolution is incapable of resolving the fine detail in the very thin sand shale layers in the study , b. It is too coarse also for core and log depth matching, c. For a static 3D reservoir geology model, we should start with the highest resolution logs possible. They can always be upscaled later to a coarser sample interval suitable for a dynamic reservoir model. 5. A log sampling interval of 0.10m/point or less (as used in the Dipmeter) is recommended. If the available LAS files are not sampled appropriately, then they can be resampled at a higher resolution. Seismic data: 6. Misties in amplitude, time and phase are a common problem with 2D seismic surveys, especially when derived from different sources or if of different vintages. Petrel software is able to correct misties, but we still recommend correcting 2D seismic misties by using a professional seismic processing software that is designed specifically for that task. 7. In the presented case study the high density of 2D seismic lines enabled the conversion of 2D seismic to the pseudo 3D seismic cube. This technique is a very helpful step to enable 3D seismic facies classification. The results can be used as a conditional model for HU. The presented work flow can be applied to the upper part of the study area (in the Z gas field) or other reservoirs in Poland where dense 148 networks of 2D seismic data commonly available. Of course, ideally one would want to use a 3D survey, if available. 8. Time to depth conversion in this study based on the check shots from only 7 wells provides a result of limited credibility. Statistical methods: 9. Several statistical methods were used for HU classification and prediction based on core and log data including Histogram, Probability plot, Ward’s clustering, LMR, ACE (chapter 3 and 4). Other methods such as Principal Component Analysis (PCA), Discriminant Analysis (DA), and Supervised Neural Network (SNN) might improve the results of the study but are beyond the scope of this investigation. 10. 6HUs classification used in the study gave good correlation between porosity and permeability in each HU. However, it is recommended to apply the GHE method as quick look analysis tool, to reduce the number of cores in the new wells is recommended. Geostatistical methods: 11. Traditional geostatistic methods based on variogram analysis for trend detection produced static 3D reservoir property models when combined with well log and 2D seismic data. However, the newly developed Multiple Point Statistic (MPS) module in Petrel might improve stochastic facies modeling, especially in very large and complex deltaic environments. Uncertainty analysis: 12. Stochastic models are based on statistical results from multiple applications of random numbers (variogram models and conditional distributions). No two runs will produce the same result, therefore we need to quantify our confidence in the results by using uncertainty analysis for both static and dynamic modeling. 13. Because of the limitation of the 2D seismic grid, we were not able to reliably include faults in our seismic interpretation. Therefore faults were excluded from the static 3D property model. If fault barriers are present in reality, but not mapped, that might introduce some mismatches in the history matching phase of the project. Since the pseudo 3D cube is based on the 2D data, it cannot be used like a regular 3D cube to 149 map the faults in more detail. Similarly, using the pseudo 3D cube for detailed horizon picking and inversion would probably not be effective. Generally speaking, a 3D survey over a producing field would cost less than a dry hole, Therefore we recommend to shoot a 3D survey over any producing field which has the upside potential to justify the cost of such a survey. 150 REFERENCES 1. Abbaszadeh, M., Fujii, H., and Fujimoto, F., Permeability Prediction by Hydraulic Flow Units - Theory and Applications, Paper SPE 30158 presented at the 1995 SPE Petrovietnam Conference, Hochiminh, Vietnam, 1–3 March. 2. Amaefule, J. O., Altunbay, M., Tiab, D., Kersey, D. G., and Keelan, D. K., 1993, Enhanced reservoir description: using core and log data to identify hydraulic (flow) units and predict permeability in uncored intervals/wells, Paper SPE 26436, 205 – 220. 3. Amaefule, J.O., 1995, A Flow Unit Approach on Integrated Reservoirs, in World Expo 1995. Sterling Publications, London, pp. 25–31. 4. An L. H., Innovative neural network approaches for petrophysical parameter reduction, PhD. Thesis, 2004. Heriot-Watt University Institute of Petroleum Engineering, UK. 5. Antelo R., Aguirre O., Andina S.A. (Repsol YPF), 2001, Permeability calculations from clustering electrofacies technique for the petrophysical evaluation in La Peña and Tundy oil fields, Paper SPE 69400. 6. Archie G.E., 1942, The Electrical Resistivity Log as an Aid in Determining Some Reservoir Characteristics, Published in Petroleum Transactions, AIME, Volume 146, 1942, Paper SPE 942054 p. 54-62. 7. Babadagli T., Al-Salmi S., 2004, A Review of Permeability - Prediction Methods for Carbonate Reservoirs Using Well-Log Data, Paper SPE 87824. 8. Bear, J., 1972, Dynamics of Fluids in Porous Media, Elsevier, New York, 1972. 9. Bohling, 2005, Stochastic simulation and reservoir modeling workflow: C&PE 940, http://people.ku.edu/~gbohling/cpe940/Simulation.pdf. 10. Breiman, L. and Friedman, J. H., 1985, Estimating Optimal Transformations for Multiple Regression and Correlation, Journal of the American Statistical Association (September), p. 580 -598. 11. Brooks R.H. and Corey A.T., 1966, Hydraulic Properties of Porous Media. Hydrology Papers, Colorado State University. Ft. Collins, No.3, March. 151 12. Caers J., 2005, Petroleum geostatistics: SPE - An interdisciplinary approach to topic in Petroleum engineering and Geosciences. 13. Carman P.C., 1937, Fluid Flow through Granular Beds, Trans. Inst. Chem. Eng. Vol.15, 1937, pp. 150-167. 14. Carrillat A. and Valles B., 2005, From 3D Seismic Facies to Reservoir Simulation: An Example From the Grane Field, Mathematical Methods and Modeling in Hydrocarbon Exploration and Production (Iske A. & Randen T., eds), Springer, 2005, VOL 7, p. 301-338. 15. Chopra S., Marfurt K. J., 2005. Seismic attributes – A historical perspective, Geophysics vol. 70 (5), p. 3SO-28SO. 16. Clayton V. Deutsch, 2002, Geostatistical reservoir modeling, Oxford University press, 2002, p. 32-152. 17. Corbett P.W.M. and Potter D.K., 2004 - Petrotyping: a basemap and atlas for navigating through permeability and porosity data for reservoir comparison and permeability prediction: Paper prepared for presentation at the international symposium of the Society of Core Analysts held in Abu Dhabi, UAE, 5-9 October, 2004, p. 1-12. 18. Corbett P.W.M., Ellabard Y., Mohhammed K., 2003, Global Hydraulic ElementsElementary Petrophysics for Reduced Reservoir Modeling. EAGE 65th Conference and Exhibition. Stavanger. F-26. 19. Documentation of wells: Z75, 79 and 84, Archive of POGC Warsaw, Department in Sanok, Biuro in Jasło (in Polish). 20. Dziadzio P., 2000, Sekwencje depozycyjne w utworach badenu i sarmatu w SE części zapadliska przedkarpackiego. Przegląd Geologiczny, v. 48, 1124 – 1138 (in Polish). 21. Dziadzio P., Liszka B., Maksym A., Staryszak G., 1997, Środowisko sedymentacji utworów miocenu autochtonicznego w brzeżnej strefie Karpat a interpretacja geologiczno-złożowa w obszarze Husów-Albigowa-Krasne. Nafta-Gaz, v.9, 40074014. (in Polish). 22. Ebanks W. J., 1987. Flow unit concept - integrated approach for engineering projects. Abstract presented June 8, during the roundtable sessions at the 1987 American Association of Petroleum Geologists Annual Convention. 152 23. Ebanks, W.J. Jr., Scheiling, M.H., Atkinson, C.D., 1992. Flow units for reservoir characterization. In: D. Morton-Thompson, A.M. Woods (Eds.), Development Geology Referrence Manual, Amer. Assoc. Petrol. Geol. Methods in Exploration Series No. 10, p. 282-284. 24. Frazier, D. E., 1974, Depositional episodes: their relationship to the Quaternary stratigraphic framework in the northwestern portion of the Gulf basin. Austin, Bureau of Economic Geology, Geological Circular 71-1. The University of Texas at Austin. 25. Geological Documentation of the Z-L gas field by Bosak B., 2007. Appendix no 5, 2007, Archive of POGC Warsaw, Department in Sanok, Biuro in Jasło (in Polish). 26. Gocad 2009 Manual, Pradigm modeling Software. 27. Gunter, G.W., Finneran, J.M., Hartmann, D.J., Miller, J.D., 1997, Early Determination of Reservoir Flow Units Using an Integrated Petrophysical Method, paper SPE 38679 prepared for presentation at the 1997 SPE Annual Technical Conference and Exhibition held in San Antonio, Texas, 5-8 October, 1997. 28. Ha Quang M. and Jarzyna J., 2008a, Dane petrofizyczne w jednostkach o jednakowych własnościach hydraulicznych do modelowania przepływów w złożu gazu w zapadlisku przedkarpackim, Pierwszy Polski Kongres Geologiczny, 26-28 czerwca, Kraków, p. 72-73. 29. Ha Quang M. and Jarzyna J., 2008b, Preliminary analysis of petrophysical parameters of the Miocene sandy-shaly formation in the Carpathian foredeep as an introduction to Eclipse® (Schlumberger) modeling, Prace INiG, 2008 nr 150 wyd. konf. Geopetrol2008, p. 209-214 (in English). 30. Ha Quang M. and Jarzyna J., 2010, Application of artificial neural networks for properties modeling using well logs and 2D seismics in the Carpathian Foredeep gas field. Prace INiG; 2010 nr 170, wyd. konf. Geopetrol2010, p. 209-212 (in English). 31. Ha Quang M. and Jarzyna J., 2011a, Hydraulic Flow Unit classification for geological modeling of the shaly-sandy reservoir using core, log and 2D seismic data, an example of the Miocene formation. Workshop, held 26 - 30 June 2011 in Auberge Saint Antoine, Quebec City, Canada. 153 32. Ha Quang M. and Jarzyna J., 2011b, Integrating of Core and Logs and 2D Seismic Data to Improve 3D Hydraulic Flow Unit Modeling. Proceedings of the 73rd EAGE Conference & Exhibition incorporating SPE EUROPEC 2011 in Vienna, Austria. 33. Ha Quang M., 2011, Deltaic Facies Classification Using Log data, International conference of young geologists HERĽANY 2011. 34. Hear C. L., Ebanks W. J., Tye R. S. and Ranganatha V., 1984, Geological Factors Influencing Reservoir Performance of the Hartzog Draw Field, Wyoming. Journal of Petroleum Tech., August 1984, p. 1335-1344. 35. IP-3.5 Manual, Interactive Petrophysics Software, 2009. 36. Jarzyna J. and Ha Quang M., 2010, Classification of rock formation on the basis of hydraulic features for porous media mobility evaluation, Klasyfikacja formacji skalnych na podstawie ich zdolności hydraulicznych dla oceny możliwości ruchu mediów w przestrzeni porowej. Jubileuszowa konferencja prof. Wacława M. Zuberka Geofizyka w Geologii i Górnictwie, Sosnowiec-Zawiercie, 15-17 września 2010 r. 37. Jarzyna J. and Ha Quang M., 2009, Podział skały zbiornikowej na jednostki o jednakowych właściwościach hydraulicznych w celu dokładniejszego wyznaczania przepuszczalności podczas modelowania przepływów w złożu gazu, Przegląd Geologiczny, v.57, nr.11, 996-1003 (in Polish, Abstract in English). 38. Karnkowski P., 1999, Oil and Gas Deposits in Poland. Geosynoptics Society “GEOS”, Cracow, Poland. 39. Kelkar B. G., Gamble R. F., Kerr. D. R., Thompson L. G., Shenoi S., 1997, Application of artificial intelligence to reservoir characterization, An interdisciplinary approach, Final report, The University of Tulsa. 40. Kozeny, J., 1927, Uber Kapillare Letung des Wassers im Boden, Sitzungsberichte: Royal Academy of Science, Vienna, Proc. Class I (1927) v. 136, p. 271-306. 41. Lang S.C., Kassan J., Benson J.M., Grasso C.A., 2000, Applications of modern and ancient geological analogues in characterization of fluvial and fluvial-lacustrine deltaic reservoirs in the cooper basin: APPEA Journal, 2000, p. 392-416. http://www.appea.com.au/publications/appea-journal.html. 154 42. MacQueen J. B., 1967, Some methods for classification and analysis of multivariate observations, Proceedings of the 5th Berkeley Symposium on Mathematical Statistics and Probability, 1, p. 281–297. 43. Martini E., 1971, Standard Tertiary and Quaternary Calcareous Rannoplankton Zonation. Proceedings of the II Planktonic Conference, Roma, 1870, 2, p. 739-785. 44. Mastalerz K., Aleksandrowski P. (Pracownia Geologiczna Explora), współpraca: P. Pasek, M. Stadtműller (Geofizyka Kraków sp. z o.o.), Wrocław-Kraków, 2004, Analiza sedymentologiczna i strukturalna osadów miocenu w otworach Z75 i Z76 na podstawie danych z pomiarów upadomierza sześcioramiennego SED (Halliburton), Sedimentological and structural analysis of the Miocene sediments in wells Z75 and Z76 on the basis of dipmeter measurements SED (Halliburton), Archive of Geofizyka Kraków sp. z o.o. (in Polish). 45. Matyasik I., Mysliwiec M., Lesniak G., Such P., 2007, Relationship between Hydrocarbon Generation and Reservoir Development in the Carpathian Foreland, Chapter 22 - Frontiers in Earth Science: Thrust Belt and Foreland Basin From Fault Kinetics to Hydrocarbon System, Springer. 46. Michael K. Ng, 1999, A note on constrained k-means algorithms, Pattern Recognition 33 (2000), p. 515-519. 47. Mikes D., Barzandji O.H.M., Bruining J., Geel C.R., 2006, Upscaling of small-scale heterogeneities to flow units for reservoir modeling, Marine and Petroleum Geology 23 (2006), p. 931–942. 48. Mikhail Y. T., 1997, Integrated reservoir description for boonsville (Texas) field using 3-d seismic, well and production data, Master thesis, University of Tulsa, 1997. 49. Morrison D.F, 1990, Multivariate statistical methods. PWN, Warsaw, Poland, p. 590. 50. Myśliwiec M., 2004, Mioceńskie skały zbiornikowe zapadliska przedkarpackiego. The Miocene Reservoir Rocks of the Carpathian Foredeep. Przegląd Geologiczny, v. 52, 581-592 (in Polish, Abstract in English). 51. Myśliwiec M., 2006a, Typy mioceńskich skał zbiornikowych złoża gazu ziemnego Żołynia-Leżajsk, a objętościowe metody szacowania zasobów. Types of the Miocene reservoir rocks (Żołynia-Leżajsk gas field) and the methods of the gas reserves estimation, Nafta-Gaz, v. 62, no 4, p. 139-150 (in Polish, Abstract in English). 155 52. Myśliwiec M., 2006b – Żołynia – Leżajsk – stare złoża, nowe zasoby. Nafta-Gaz, v. 62, nr 3, p. 97-105. (in Polish, Abstract in English). 53. Myśliwiec M., Madej K., Bys I., 2004, The Miocene gas fields discovered in the Rzeszów area, Carpathian Foredeep, on the base of the Direct Hydrocarbon Indicators, Przegląd Geologiczny, v. 52, no 7, p. 501–506 (in Polish, Abstract in English). 54. Olszewska B., 1999, Biostratygrafia neogenu zapadliska przedkarpackiego w świetle nowych badań mikropaleontologicznych. Prace Państwowego Instytutu Geologicznego, 168, p. 9-28 (in Polish, Abstract in English). 55. Oszczypko N., 1999, Przebieg mioceńskiej subsydencji w polskiej strefie zapadliska przedkarpackiego. Prace Państwowego Instytutu Geologicznego, 168, p. 209-213 (in Polish, Abstract in English). 56. Oszczypko N., Krzywiec P., Popadyuk I., Peryt T., 2006, Carpathian Foredeep Basin (Poland and Ukraine); Its Sedimentary, structural and Geodynamic Evolution. In: Golonka J. and Picha F. (Eds.), The Carpathians and their Foreland: Geology and Hydrocarbon resources. AAPG Memoir, 84, p. 293 – 350. 57. Oszczypko, 1999, From remmant oceanic basin to collision-related foreland basin - a tentative history of the Outler Carpathians. Geol. Carpathica, 50 (special issue). 58. Petrel® 2009 and 2010 Manual, Schlumberger Software. 59. Rögl F., 1996, Stratigraphic Correlation of Paratethys Oligocene and Miocene. Mitt Ges Geol Bergbaust ud Öster, p. 65-73. 60. Roksandić M., 2006, Seismic facies analysis concepts: Gephysical Processing Vol.26, p. 382-398. 61. Roxar Manual, RMS modeling software, 2009. 62. Rushing J.A., Newsham K.E., Blasingame T.A., 2008, Rock Typing - Keys to Understanding Productivity in Tight Gas Sands, Paper SPE 114164. 63. Sang H. L., Arun K, and Akhil D. G., SPE, 2002, Electrofacies characterization and permeability predictions in complex reservoirs, Paper SPE 78662. 64. Shuguang M., Journel A.G., 1999, Conditional 3D simulation of lithofacies with 2D seismic data, Computer & Geosciences 25, p. 845-862. 156 65. Skalinski M., Gottlib-Zeh S., Moss B., 2005, Defining and Predicting Rock Types in Carbonates-an Integrated Approach using Core and Log Data in Tengiz Field, SPWLA 46th Annual Logging Symposium. 66. SLB, Plug-in, 2008, Petrel plus-in upscaling 2D seismic into 3D grid, http://slb.ru/sis/plugins/Petrel2008.Plugins.2DSeis2Seis3D.zip. 67. Śmist P., 2003 – Zmiany w litologiczno-facjalnym wykształceniu osadów miocenu w rejonie Żołynia-Chałupki Dębniańskie-Grodzisko Dolne. Raport Przedsiębiorstwa Usług Laboratoryjnych i Geologicznych Sp. z o.o., Oddział w Jaśle, Archive of POGC Warsaw, Department in Sanok, Biuro in Jasło (in Polish). 68. Svirsky D., Ryazanov A., Pankov M., Corbett P.W.N., Posysoyev A., 2004, Hydraulic Flow Units Resolve Reservoir Description Challenges in the Siberian Oil Fields. Paper SPE 87056, p. 1-15. 69. Tan, Steinbach, Kumar, 2004, Introduction to Data Mining, Steven F. Ashby Center for Applied Scientific Computing Month DD, 1997 (Internet presentation, only: http://www.gersteinlab.org/courses/545/07-spr/slides/DM_cluster-tan.ppt). 70. Udegbunam, E. and Amaefule J.O., 1996, An Improved Technique for Modeling Initial Reservoir Hydrocarbon Saturation Distributions: Applications in Illinois (USA) Aux V. Oil reservoirs, JPT, V. 21. 71. Varavur S., Shebl H., Salman S.M., Shibasaki T., and Dabbouk C., 2005, Reservoir Rock Typing in a Giant Carbonate, Paper SPE 93477. 72. Worthington, P.F. & Cosentino L. 2003. The role of cut-off in integrated reservoir studies, Paper SPE 84387. 73. Xue G., Datta-Gupta A., Valko P., and Balsingame T., 1996, Optimal Transformations for Multiple Regression: Application to Permeability Estimation from Well Logs, SPE 35412 presented at the Improved Oil Recovery Symposium, Tulsa, Ok, 21 April 1996. 157 LIST OF FIGURES AND TABLES FIGURE Page Chapter1 1.1 The main workflow in this study for integration of static and dynamic 2 models Chapter2 2.1 General overview of the Carpathians and Carpathian Foredeep 5 (Oszczypko, 2006) 2.2 Schematic map showing the submarine deposits diversity in the eastern 6 part of the Carpathian Foredeep (Myśliwiec, 2004a) 2.3 Location of Z gas-field with cored and logged wells; tectonic elements 8 marked in red (after Myśliwiec, 2006b and Myśliwiec et al 2004 2.4 The result of grain size analysis of powdered sandstone taken at 677m, 14 the depth of the Z75 well 2.5 Photomicrographs of a polished section impregnated with blue resin, 14 from a sample obtained at a depth of 855.20m in well Z75; crossed nicoles (left – magnification 14X, right – 120X) (Documentation of well Z75) 2.6 Scanning microscope photomicrographs of a polished section from a 15 sample obtained at a depth of 855.20m in well Z75; magnification 600X (Documentation of well Z75) 2.7 Well logs (GR and NPHI) showing vertical facies distribution; 19 a) an example of turbidities from the depth section of 925 – 1025 m in the Z84 well; low value of GR and rather low NPHI, sandstones dominate over shales, b) a typical parasequence coarsening upward in deltaic sediments in the Z75 well, c) sediments of shallow shelf form the top of the Sarmatian succession (GR and NPHI are very close to one another), sediments are shaly and have high neutron porosity 2.8 RMS amplitude map and GR log for each well shows deltaic facies 20 158 distribution in the study area 2.9 Overview sequence stratigraphy in the study area based on 2D seismic 23 data a) seismic facies classification based on seismic attributes (Chapter 5), b) balancing section a) c) some basic sequence stratigraphy interpretation, d) successive progradation of delta lobes deposits an offlap succession with a clinoform geometry (after Frazier 1974) 2.10 Comparison of the result of well log correlation; 24 a) cross section correlation Z-81 – Z76 - Z-82 based on GR and NPHI (after Mastalerz, et al., 2004), b) cross section basic on seismic facies classification (Chapter 5) Chapter3 3.1 Various parameters used in defining geologic flow units (Ebanks et al., 30 1992); four flow units are defined on the basis of lithofacies, pore types, porosity, and permeability crossplots, capillary pressure measurements, and gamma-ray log response (after Ebanks et al. (1992) 3.2 Histogram of porosity (a), permeability (b), FZI (c) for 570 core data 37 measurements 3.3 Normal probability plot of log(FZI) with division into 6 homogeneous 38 groups of HUs with constant FZI 3.4 Clustering of the FZI – HU data set into six groups, according to the 40 Ward method 3.5 Porosity-permeability crossplot, the hydraulic unit classification of all 43 the core data 3.6 z vs. RQI crossplot of all the hydraulic units. The mean FZI values for 43 each hydraulic unit are given by the intercept of the straight lines at z =1 3.7 Dispersion plot of PHI_core vs. K, and the six HUs defined in the area 44 of core origin data 159 3.8 Dispersion plot and correlation line between the core origin 44 permeability vs. the permeability calculated from the mean values of FZI for HU 3.9 Permeability vs. porosity data on the background of 7 GHE 45 3.10 Permeability, K_GHE, calculated on the basis of relationship RQI vs. 45 Φz for 7 GHE and core origin permeability, K_core Chapter4 4.1 Two approaches to depth matching between the core and log data in 49 well Z76. Crosses – lab porosity – PHI_core, sampled irregularly; continuous curves with triangles – NPHI, sampled regularly, 0.25 m. The two horizontal scales of porosity and two vertical scales of depth relate to the matched data sets 4.2 Flow chart for two procedures applied to obtain uniform HUs in the 51 study data 4.3 Dispersion plots of ln(FZI) vs. GR and ln(FZI) vs. NPHI, and 53 histograms of GR and NPHI; Gaussian distributions included in histograms 4.4 a) Optimal transformation of EL14 58 b) Optimal transformation of A2 c) Optimal transformation of ln(FZI) a) d) Dispersion plot and correlation for FZI_tr vs. (log _ i tr Ai _ tr ) i 4.5 a) Comparison of ln(FZI_core) to ln(FZI_pre_LMR) for group G3 59 b) Comparison of ln(FZI_core) to ln(FZI_pre_LMR) for well Z76 c) Comparison of ln(FZI_core) to ln(FZI_pre_ACE) for group G3 d) Comparison of ln(FZI_core) to ln(FZI_pre_ACE) for well Z76 4.6 Results of FZI_core and FZI_pre (predicted from log) for two 60 approaches in the selected section of well Z76; a) LMR; b) ACE; consecutive numbers of data in the depth section are on the horizontal axis 4.7 a) Prediction results for FZI_pre for well Z76, by applying the ACE 61 solution for group G3 for the full geological profile; left colored 160 field – GR anomaly, right colored steps – log predicted HU b) Prediction results for FZI_pre for well Z76, by applying the ACE solution for group G3 excluding pure shales; left colored field – GR anomaly, right colored steps – log predicted HU; bimodal histogram of GR in well Z76 to show how to exclude pure shales 4.8 a) Dispersion plot and correlation line between K_core and 63 K_pre(ACE) for the G3 group b) Dispersion plot and correlation line between K_core and K_pre(ACE) for the Z76 well 4.9 GR (left hand side) and HU (right hand side) presented in wells Z81- 64 76-82 in group G3 on the background of 2D seismic section; HU color scale is the same as in Fig. 13 4.10 a) Comparison of FZI_pre_ACE (line) and FZI_core (points) in well 65 D2 b) Comparison of K_pre_ACE (line) and K_core (points) in well D2 4.11 The workflow of the K mean clustering 71 4.12 The limitations and overcomes limitations of the K mean clustering 71 (after Tan, al et, 2004) 4.13 The matrix crossplots of 5 logs (GR, DT, NPHI, EN16 and EN64) and 72 histograms for each curves show data distribution of group 3 (Z76, Z81, Z82) before applying K means clustering 4.14 Applying K mean clustering for the group 3: 73 (a) The cluster grouping dendrogram shows 4 clusters with diffident 4 colors respond (b) The cluster groups randomness 4.15 The cross plots of data in well Z-76 shows 6 rock types (6 clusters) a) GR/DT, 4.16 b) GR/NPHI, 74 c) GR/EN64, d) GR/EN16 Comparison between 6 rock types (6RTs) and 6 hydraulic flow unit 75 (6HUs) 4.17 Relationship between geological model, reservoir model and flow unit 76 model (modified from D. Mikes al et. 2006) 1. Geological model and assignment of its elements 161 2. Reservoir model and assignment of its elements 3. Micro simulation model, consisting of numerical flow simulation on all flow cell models of the reservoir, yielding effective oneand two-phase permeabilities, and capillary pressures 4. Macro simulation model, consisting of numerical flow simulation on the entire reservoir, yielding production history Chapter 5 5.1 Deltaic facies model and special relationships control by facies 82 distribution a) reservoir heterogeneity and longitudinal facies variation developed within the east-west distributary channel of the Neales Delta (modified from Lang et al., 2000), b) how unknown point relates to a known point (Gocad Manual, 2009) 5.2 Variogram modeling 84 a) concepts of variogram parameters, b) variogram ellipsoid around a grid cell. The values of cells outside the ellipsoid are independent of the value of the current cell (Roxar Manual, 2009), c) horizontal variogram of porosity modeling (7 wells), d) vertical variogram of porosity modeling (7 wells). 5.3 2D seismic data in the study area; 89 a) 25 lines 2D seismic survey and 10 wells location, b) the miss-tie correction for the 2D seismic survey 5.4 Horizon picking 90 a) four horizons interpreted basically on 25 lines of 2D seismic and the well-tops of main interpreted horizons: H1 (Top_7), H2 (Top_12), H3 (Top_15) and H4 (Top_17), b) the 3D structure mode in the study area; four surfaces were created from 4 horizons H1, H2, H2 and H4 and 10 well locations 5.5 The workflow for reservoir modeling in this study 92 5.6 Convertion from 2D seismic to pseudo 3D seismic cube using SGS; 94 a) 2D seismic survey - top and bottom of the reservoir, 162 b) 2D seismic profiles - uplscaled to 3D grid, c) pseudo 3D seismic cube after applying the SGS method 5.7 Pseudo 3D seismic cube 95 a) the block average data of 2D seismic after upscaling into 3D grid; b) comparison between 2D seismic profiles and the pseudo 3D seismic cube 5.8 Channel system and delta lobe (geobody) facies from pseudo 3D seismic 99 extraction; a) envelope histogram before filtering, b) envelope histogram after filtering, c) model of deltaic environment in the study area 5.9 The workflow for seismic facies classification using ANN method 101 (Carrillat and Valles, 2005) 5.10 Application of the Unsupervised Neural Network technology for seismic 102 facies classification 5.11 Seismic attributes generated from pseudo 3D seismic amplitude for UNN 102 5.12 Reduction from 10Facies_UNN to 6Facies_UNN on the basis of the 105 facies histogram; a) 10Facies_UNN histogram, b) 6Facies_UNN histogram, c) Crossplot of Amplitude & Relative acoustic impedance for 10Facies_UNN, d) Crossplot Amplitude & Relative acoustic impedance for 6Facies_UNN, e) Envelope & Relative acoustic impedance for 10Facies_UNN, f) Envelope & Relative acoustic impedance for 6Facies_UNN 5.13 3D Facies classification using UNN method 106 a) 3D 10Facies_UNN and cross section, b) 3D 6Facies_UNN and cross section 5.14 Comparison 6Facies_UNN and 6RT_SIS; 108 a) 6Facies_UNN and 6RT_log, b) 6Facies_UNN histogram, a) c) 6RT_SIS and 6RT_log, d) 6RT_SIS (blue) and 6TR_log (green) 163 histogram 5.15 Comparison between seismic facies classification (6Facies_UNN) and 109 3D rock type model (6RT_SIS) a) 6Facies_UNN model filtered to display only faceis F8_UNN showing clearly delta lobe distribution in the study area, b) 6RT_SIS model filtered to display 3 Rock Types (RT7, RT8, RT9) showing in more details rock type distribution and honoring wells location 5.16 Comparison between the 6RT_SIS and 6HU_SIS 110 a) 3D 6RT_SIS b) 3D 6HU_SIS 5.17 Comparison of 6RT_SIS and 6HU_SIS in 4 horizons: H1, H2, H3 and H4 111 5.18 Comparison between 3D PHI_HU model and the 3D K_HU model; 113 a) porosity model constrained by HU model, b) permeability model calculated from the porosity model using KozenyCarman equation 5.19 Comparison between PHI_HU model and the K_HU model; results 114 presented on various surfaces 5.20 Cross plot of 3D PHI_HU versus 3D K_HU; shown very clear 115 relationship between K and PHI responding to each HU 5.21 Cross plot between permeability from logs (K_log) and permeability 115 calculated for 3D_K_HU model from the mean values of FZI for HU 5.22 Typical capillary pressure curves for different flow units fitted with the 117 Brooks and Corey’s equation (Udegbunam & Amaefule, 1996) 5.23 Water saturation model constrained by Hydraulic Flow Unit model 118 (6HU_SIS) 5.24 Comparison between 6HUs histogram and Net to Gross (NTG) histogram 119 5.25 Net to gross model based on 3D HU model 120 Chapter 6 6.1 Schematic diagram of two stages approach (Kelkar al el., 1997) 125 6.2 The four main stages to integrate geophysical and geological data into 126 reservoir modeling with HU method in the study (modified from Kelkar 164 al el., 1997) 6.3 Hydraulic Flow Unit control 128 a) Cross plot K vs. PHI responding to 6HUs (core data), b) 6HUs section of 3 wells (Z81, Z76 and Z82), c) Permeability section of 3 wells (Z81, Z76 and Z82) 6.4 The workflow of history matching under Hydraulic Flow Unit control 129 (modified from Mikhail, 1997) 6.5 The Process Dialog for Flow-Based Tensor Upscaling 131 6.6 Comparison of scale of permeability in well Z76 132 6.7 (1) Log scale; (2) Log upscaled in 189 layers (static model scale); (3) Ki, (4) Kj, (5) Kk - permeability downscaled in 99 layers by Flow Based Tensor Upscaling method (dynamic scale) Comparison of static scale results (a, c, e) and dynamic scale results (b, 133 d, f) 6.8 Gas (Kg) and water (Kw) relative permeability curves in sand and shaly- 135 sand reservoir (Petrel 2010, Manual) 6.9 Rock compaction function responding with porosity model and each 136 hydraulic flow unit 6.10 Historical production data for 3 wells: Z74A, Z75 and Z76 139 a) Gas Flowrate, b) Bottom Hole Pressure (BHP) 6.11 History matching results (line colors) and observed data (dot points) of 140 traditional method 6.12 History matching results (line colors) and observed data (dot points) by 141 HU method with reservoir volume control 6.13 History matching results (line colors) and observed data (dot points) by 142 HU method with gas control TABLE Page Chapter 2 2.1 Top of the Precambrian in study area 6 2.2 Stratigraphy of the Miocene formation (Rögl, 1996; Martini, 1971) 10 165 2.3 2.4 Important correlation horizons (event markers) distinguished in the Z75 and Z76 wells in the selected depth section according to its geological core description (after Mastalerz, et al., 2004) Genetic sequences, SG, and other stratigraphic intervals 25 26 distinguished in the Miocene succession in the Z75 well (after Mastalerz, et al., 2004) 2.5 Genetic sequences and other stratigraphic intervals distinguished in 27 the Miocene succession in the Z76 well (after Mastalerz, et al., 2004) Chapter 3 3.1 Permeability correlations developed on the basis of pore and grain 33 properties (modified after Babdagli and Al-Slmin, 2004) 3.2 Simple statistics of permeability, porosity, FZI and the determination 41 coefficients (R2) for the permeability, calculated from the FZI_mean and from the core in 6 HUs. 3.3 Global hydraulic elements (GHE) template parameters 41 Chapter4 4.1 Correlation coefficient R between ln(FZI) and log 54 4.2 Correlation coefficients between ln(FZI) and groups of log with raw 54 data 4.3 Selected independent variables used in a linear multiple regression 55 4.4 Correlation coefficients (R) between ln(FZI_core) and the results of 56 LMR and ACE. 4.5 Comparison of results obtained in well Z76; FZI_core – calculated on 62 the basis of porosity and permeability from cores, HU_core – six hydraulic units determined according to FZI_core, FZI_pre calculated on the basis of log and using ACE algorithm, HU_pre – 6HUs determined according to FZI_log 4.6 Summary of selected rock types studies and definitions for sandstone 67 reservoirs (Rushing, et al. 2008) 4.7 The statistics of the group 3 (Z-76, Z-81, Z-82) for 14 clusters after 73 applying K means clustering 4.8 Comparison between hydraulic flow units (HUs) and rock types 77 (RTs) in the study 166 Chapter 5 5.1 The grid parameters used for stochastic variogram models 83 5.2 The 3D grid parameters 88 5.3 Seismic attributes corresponding to reservoir characterizations 97 (Chopra, et al., 2005) 5.4 The values cut-off reservoir basic on 6 HUs 119 Chapter 6 6.1 Upscaling structure models for Eclipse simulation 129 6.2 Pressure and temperature measurement results (Documentation of Z- 134 L gas field) 6.3 Gas components (Documentation of Z-L gas field) 135 6.4 Perforation intervals and time starting produce and status of Z74A, 137 Z75, and Z76 wells 6.5 Historical production data for history matching simulation (Appendix A) 157 167 LIST OF ABBREVIATIONS ACE: Alternating Conditional Expectation ANN: Artificial Neural Network BHP: Bottom Hole Pressure BVI: Bulk Volume Irreducible Water CDF: Cumulative Density Function DA: Discriminant Analysis FWL: Free Water Level FZI: Flow Zone Indicator GHE: Global Hydraulic Elements GSLIB Geostatistical Software Library GWC: Gas Water Contact HU: Hydraulic Flow Unit LMR: Linear Multiple Regression MCA: Model Based Cluster MFS: Maximum Flooding Surfaces MPS: Multiple Point Statistics NTG: Net To Gross OOIP: Original Oil In Place PCA: Principal Component Analysis PVT: Back Oil Fluid Model RHOB: The Bulk Density RMS: Root Mean Square RQI: Reservoir Quality Index RT: Rock Types SCAL: Special Core Analysis SEM: Scanning Electronic Microscope SGS: Sequence Gaussian Simulation SIS: Sequential Indicator Simulation SNN: Supervised Neural Networks SOM: Self Organizing Map UNN: Unsupervised Neuron Network 168 APPENDIX A Table 6.5 Historical production data for history matching simulation Wells [-] Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ad Z-74ag Top Bottom [m] [m] 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 781 786 723 744 Qg Presure [103nm3] [MPa] 162.125 6.69 1529.181 6.34 1534.069 5.96 1639.6 5.72 1700.605 5.54 1495.481 5.16 1680.524 4.9 1599.296 4.64 1627.63 4.38 1563.286 4.08 1626.239 3.79 1455.28 3.63 1482.482 3.46 1405.91 3.43 976.21 3.65 1315.673 3.41 1214.373 3.36 1065.866 3.34 1107.32 3.23 1065.642 3.2 1062.075 3.14 996.39 3.09 979.864 3.19 933.339 3.02 564.734 4 663.921 3.91 657.605 3.86 683.731 3.6 649.732 4.16 608.734 3.88 606.785 3.83 489.339 3.75 411.331 3.75 340.133 3.83 234.236 3.75 263.551 3.88 219.418 3.76 239.911 3.76 211.77 2.62 19.704 0 128.039 6.69 Qw [Kg] 0 375 780 855 1250 775 1160 1120 1080 670 1360 1315 746 827 620 1669 1089 898 971 400 1244 919 870 813 300 491 621 12599 19141 33013 84965 123775 186410 211521 199335 243467 251921 320876 329643 34201 0 Days Time [-] [DD/MM/YYYY] 4 9/1/2006 31 10/1/2006 30 11/1/2006 31 12/1/2006 31 1/1/2007 28 2/1/2007 31 3/1/2007 30 4/1/2007 31 5/1/2007 30 6/1/2007 31 7/1/2007 31 8/1/2007 30 9/1/2007 31 10/1/2007 23 11/1/2007 31 12/1/2007 31 1/1/2008 29 2/1/2008 31 3/1/2008 30 4/1/2008 31 5/1/2008 30 6/1/2008 31 7/1/2008 31 8/1/2008 20 9/1/2008 31 10/1/2008 30 11/1/2008 31 12/1/2008 31 1/1/2009 28 2/1/2009 31 3/1/2009 30 4/1/2009 31 5/1/2009 30 6/1/2009 31 7/1/2009 31 8/1/2009 30 9/1/2009 31 10/1/2009 30 11/1/2009 4 12/1/2009 4 9/1/2006 169 Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-74ag Z-75d 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 723 656 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 744 660 1304.111 1275.55 1393.805 1410.978 1221.825 1352.397 1253.835 1209.746 1066.111 1128.999 1023.15 1044.447 1088.593 967.846 970.746 913.511 811.253 847.597 772.387 743.64 681.573 643.237 584.338 348.396 501.288 480.42 479.007 456.16 356.743 308.649 298.301 204.283 153.998 19.16 126.179 93.125 10.014 26.119 95.244 91.73 96.753 44.589 91.561 107.224 99.716 444.235 6.32 6.08 5.84 5.62 5.6 5.38 5.22 5.2 5.07 4.84 4.87 3.84 4.46 4.39 4.32 4.25 4.1 4 3.97 3.88 3.79 3.81 3.66 3.81 3.76 3.73 3.73 3.6 3.71 3.74 3.71 3.74 3.95 4.95 4.03 4.56 4.58 4.09 4.11 4.16 4.2 4.2 4.3 4.34 4.34 5.4 724 1372 1515 4130 4755 7710 8170 9440 9400 11720 13414 19543 23561 19620 23270 23838 23243 27793 27840 30789 32920 40529 41524 28260 40790 56205 62144 68746 60454 67405 85567 67199 53358 28417 115768 59950 7770 47480 75820 86560 62610 29040 82616 78020 75935 0 31 30 31 31 28 31 30 31 30 31 31 30 31 30 31 31 29 31 30 31 30 31 31 20 31 30 31 31 28 31 30 31 30 6 30 24 3 9 31 30 31 15 31 31 30 15 10/1/2006 11/1/2006 12/1/2006 1/1/2007 2/1/2007 3/1/2007 4/1/2007 5/1/2007 6/1/2007 7/1/2007 8/1/2007 9/1/2007 10/1/2007 11/1/2007 12/1/2007 1/1/2008 2/1/2008 3/1/2008 4/1/2008 5/1/2008 6/1/2008 7/1/2008 8/1/2008 9/1/2008 10/1/2008 11/1/2008 12/1/2008 1/1/2009 2/1/2009 3/1/2009 4/1/2009 5/1/2009 6/1/2009 10/1/2009 11/1/2009 12/1/2009 1/1/2010 2/1/2010 3/1/2010 4/1/2010 5/1/2010 6/1/2010 7/1/2010 8/1/2010 9/1/2010 6/1/2009 170 Z-75d Z-75d Z-75d Z-75d Z-75d Z-75d Z-75d Z-75d Z-75d Z-75d Z-75g Z-75g Z-75g Z-75g Z-75g Z-75g Z-75g Z-75g Z-75g Z-75g Z-75g Z-75g Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 Z-76 656 656 656 656 656 656 656 656 656 656 541 541 541 541 541 541 541 541 541 541 541 541 560 560 560 560 560 560 560 560 560 560 560 560 560 560 560 560 560 560 560 560 560 660 660 660 660 660 660 660 660 660 660 571 571 571 571 571 571 571 571 571 571 571 571 566 566 566 566 566 566 566 566 566 566 566 566 566 566 566 566 566 566 566 566 566 938.399 899.973 854.949 548.64 316.657 370.409 278.502 221.221 283.789 37.726 525.425 1059.605 954.162 888.097 813.133 667.383 595.539 482.994 387.384 382.991 338.968 369.476 138.05 501.928 413.863 400.916 360.82 166.391 242.555 279.552 271.229 196.072 125.213 41.673 43.634 36.54 33.652 40.812 36.233 14.136 69.266 47.836 43.042 5.37 4.56 4.56 2.99 2.94 2.99 2.99 2.99 2.99 2.99 3 4.42 3.96 3.96 3.61 3.56 3.61 3.61 3.61 3.61 3.61 3.61 4.69 4.37 4.24 4.24 3.75 3.75 3.65 3.53 3.53 3.53 3.37 3.33 3.37 3.37 3.37 3.37 3.37 3.37 3.37 3.37 3.37 0 1235 932 2846 29187 50009 53988 58642 92049 15544 0 0 1380 2372 4747 4290 5489 4234 3891 4117 3970 4454 70 10 500 33 0 280 1091 3558 3703 2824 6301 5179 5014 5528 4480 4503 2323 1278 7822 9501 9349 31 31 30 31 30 31 31 28 31 6 15 31 31 30 31 30 31 31 28 31 30 31 9 31 28 31 30 24 30 31 31 25 31 30 31 31 28 31 22 6 31 31 30 7/1/2009 8/1/2009 9/1/2009 10/1/2009 11/1/2009 12/1/2009 1/1/2010 2/1/2010 3/1/2010 4/1/2010 6/1/2009 7/1/2009 8/1/2009 9/1/2009 10/1/2009 11/1/2009 12/1/2009 1/1/2010 2/1/2010 3/1/2010 4/1/2010 5/1/2010 12/1/2008 1/1/2009 2/1/2009 3/1/2009 4/1/2009 5/1/2009 6/1/2009 7/1/2009 8/1/2009 9/1/2009 10/1/2009 11/1/2009 12/1/2009 1/1/2010 2/1/2010 3/1/2010 4/1/2010 6/1/2010 7/1/2010 8/1/2010 9/1/2010 Note: Z74 and Z74A are the different wells, and both are dual-completed, equipped with packers and two tubing strings; “g” means upper completion; “d” means lower completion (or horizon) 171