Uncertainty quantification for reservoir geomechanics

Uncertainty quantification for reservoir geomechanics

Accepted Manuscript Uncertainty quantification for reservoir geomechanics Leonardo Cabral Pereira, Marcelo S´anchez, Leonardo Jos´e do Nascimento Guim...

1MB Sizes 1 Downloads 406 Views

Accepted Manuscript Uncertainty quantification for reservoir geomechanics Leonardo Cabral Pereira, Marcelo S´anchez, Leonardo Jos´e do Nascimento Guimar˜aes PII: DOI: Reference:

S2352-3808(16)30090-9 http://dx.doi.org/10.1016/j.gete.2016.11.001 GETE 51

To appear in:

Geomechanics for Energy and the Environment

Received date: 1 November 2015 Revised date: 4 November 2016 Accepted date: 6 November 2016 Please cite this article as: Pereira LC, S´anchez M, Guimar˜aes. Uncertainty quantification for reservoir geomechanics. Geomechanics for Energy and the Environment (2016), http://dx.doi.org/10.1016/j.gete.2016.11.001 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

*Highlights

1

HIGHLIGHTS

2 3



Reservoir geomechanics incorporating uncertainty quantification

4



Non-probabilistic methods, evidence theory for handling epistemic uncertainties.

5



Analyses via analytical solutions of injection pressure for fracture propagation and

6

reservoir compaction

7



Comparison of deterministic, probabilistic and non-probabilistic methods

8



Effect of adding increasingly information on uncertainty reduction

1

*Manuscript (pages & lines numbered) Click here to view linked References

1

SUBMITTED TO:

Geomechanics for Energy and the Environment

2 3 4

DATE:

5

TITLE:

4th November 2016

Uncertainty Quantification for Reservoir Geomechanics

6 7 8 9 10 11 12

AUTHORS: Leonardo Cabral Pereira1, Marcelo Sánchez*2, Leonardo José do Nascimento Guimarães3 AFFILIATIONS:

13 14 15 16 17

1

18 19 20 21 22 23 24 25 26 27 28 29 30

Dr Marcelo Sánchez Associate Professor Zachry Department of Civil Engineering Texas A&M University College Station Texas 77843-3136, USA Telephone: (+1) 979 862 6604 Fax: (+1) 979 862 7696 E-mail: [email protected]

Petrobras - Petróleo Brasileiro S.A, Brazil Zachry Department of Civil Engineering, Texas A&M University, College Station, US 3 Federal University of Pernambuco, Recife, Brazil 2

*CORRESPONDING AUTHOR:

Keywords: reservoir geomechanics, epistemic uncertainty, analytical solution, non-probabilistic methods, uncertainty quantification.

1

31

ABSTRACT

32

Geomechanics plays a critical role in oil and gas reservoir problems. One of the challenges related

33

to reservoir geomechanics is that information about sediments and rock properties is generally

34

very limited. Furthermore, these geomaterials are typically highly heterogeneous. Therefore, a

35

formal framework for dealing with the uncertainties associated with this type of problem is much

36

required. Uncertainty quantification can be performed either via probabilistic or non-probabilistic

37

methods. The latter approach is generally better suited for handling epistemic (or reducible)

38

uncertainties, which are the more usual in geomechanical reservoir problems. A non-probabilistic

39

framework based on the evidence theory was adopted in this work for dealing with the

40

uncertainties related to the material properties. The contribution focuses on how additional

41

information can be incorporated in a consist fashion in the analysis to reduce the uncertainties.

42

Two analytic solutions associated with typical reservoir geomechanics problems, such as the

43

estimation of the fracture pressure and subsidence, were adopted as the case studies. These

44

application cases shown that the evidence theory was able to reduce the uncertainties associated

45

with these problems when additional information about material properties became available.

46

Another advantage of the adopted framework is that experts’ opinion related (for example) to the

47

reliability of the experimental data gathered from different sources could be explicitly incorporated

48

in the study. It is shown that this type of approach can be instrumental in assisting the decision-

49

making process.

50

2

51

1.

52

Reservoir geomechanics encompasses aspects related to rock mechanics, structural geology and

53

petroleum engineering. Some typical geomechanical problems associated with petroleum and gas

54

production are, amongst others: pore pressure prediction, well design, reservoir compaction, CO2

55

(or H20) injection to enhance oil production, reactivation of geological fault, and hydraulic

56

fracturing. A common characteristic to these problems is that information about sediments and

57

rocks properties is generally scarce. These are also highly heterogeneous geological materials, so

58

significant variations of their geomechanical properties are anticipated. The combination of these

59

factors implies that huge uncertainties are generally associated with problems involving reservoir

60

geomechanics. Therefore, it is essential to develop proper frameworks for handling the

61

uncertainties associated with this kind of analysis.

62

Uncertainty can be defined as the lack of exact knowledge, regardless the cause of this ignorance

63

[1]. Each decision (or set of decisions) is associated with several factors and thus it is highly

64

affected by the uncertainties involved in a particular problem [2]. Uncertainties are typically

65

present in all the stages of a reservoir geomechanics project, from the adoption of simplifying

66

assumptions, to the selection of the constitutive models to be used in the analysis, and the adoption

67

of the input parameters. The need to identify the different sources and types of uncertainties has

68

become a subject of increasing interest in recent years. Different classifications of uncertainties

69

can be found in the literature (e.g., [3]; [4]; [5]). The definitions presented below have been adopted

70

in this work.

71

The dual nature of uncertainties can be described as [6]:

72

INTRODUCTION

i) aleatory uncertainty, associated with random behavior of the system, and

3

73

ii) epistemic uncertainty, associated with the lack of knowledge about a system and/or its

74

properties.

75

Uncertainty type i) above is also known as: stochastic, or type A, or irreducible uncertainty; while

76

uncertainty type ii) is also called subjective, or type B, or reducible uncertainty. In reservoir

77

geomechanics the uncertainties are generally associated with the lack, or limited, information

78

available in terms of material properties and initial stresses. Therefore, it can be considered that

79

these uncertainties are manly epistemic in nature.

80

Both probabilistic and non-probabilistic methods can be used to quantify uncertainties. Non-

81

probabilistic approaches are more appropriate when dealing with epistemic uncertainties (e.g. [6]).

82

A methodology based on the non-probabilistic evidence theory (e.g. [6], [7]) was adopted in this

83

work to handle the uncertainties related to reservoir geomechanics problems. An aspect of interest

84

in this kind of problem is how to reduce the uncertainties associated to problem under study by

85

incorporating any additional information that may become available during the analysis. To study

86

this aspect two relatively simple analytical solutions associated with the problems of fracture

87

propagation pressure and reservoir compaction were selected as application cases. These equations

88

were solved deterministically as well as using classical probabilistic and non-probabilistic

89

methods.

90

The paper is organized as follows. First, the main aspects related to the analytical solutions adopted

91

in this work are discussed. Second, the basic components of the evidence theory adopted for the

92

uncertainty quantification analyses are introduced. Third, the main components and outputs of the

93

application cases are discussed. Finally, the main conclusions of this work are presented.

94 95

4

96

2

97

This section presents the main components and equations associated with the two application cases

98

analyzed in this paper: fracture propagation pressure, and reservoir compaction.

99

2.1.

ANALYTICAL SOLUTIONS

Injection above the fracture propagation pressure

100

Water injection is one of the most common methods to maintain and/or enhance the production of

101

petroleum reservoirs. This technique is very effective to assist oil extraction once the reservoir

102

production by primary mechanisms is exhausted. It is also a convenient production strategy in the

103

early stages of the field exploitation to enhance oil production. During the water injection process,

104

the injectivity can decline because of specific rock and fluids features, geometry of the injectors

105

and producers wells, precipitation of salts, or by the presence of solid particles in the injection

106

water. One of the best ways to avoid injectivity loss is to inject above the fracture propagation

107

pressure [8].

108

A semi-analytical model to predict fracture propagation during water flooding processes was first

109

proposed by Hagoort et al. [9]. However, the thermally induced stresses (one of the most relevant

110

factors) was not considered in this model. Consequently, the transfer of heat between the injected

111

fluid and the formation was not discussed. Afterwards, a three model regions considering the

112

behavior of water flow together with the fracture mechanics was developed by [10]. This model

113

has been widely cited and applied in problems related to reservoir geomechanics (e.g. [11]; [12];

114

[13]). The formulations proposed by de Souza et al. [8] and Perkins and Gonzalez [10] to calculate

115

the reservoir fracture pressure is presented as follows.

116

2.1.1

Fracture pressure calculation

5

117

The fracture pressure in vertical wells for the case of a predominant normal fault regime

118

considering the effect of the penetrating fluid, and thermally induced stress changes, can be

119

calculated through (e.g. [14]; [15]):

120

Pfrat

  h   TPfrat   max  (3 h   H  eTa Pe   T )   TPfrat  (1    eTa ) 

(1)

121

where  h is the minimum horizontal stress;  H is the maximum horizontal stress; eTa is an elastic

122

parameter calculated as  (1  2 (1   )) , where α and υ are the Biot and Poisson coefficients,

123

respectively; Pe is the reservoir initial pressure;  is the pore pressure factor in tensile failure

124

criterion;  T is the tensile strength of the reservoir rock, and  TPfrat is the thermally induced

125

stress increment because of the temperature difference between the injected and formation fluids,

126

which can be calculated as follows:

127

 TPfrat 

 (Tr  Ti ) E ff 1 

(2)

128

where β is the thermal expansion coefficient; E is the Young modulus; Tr is the reservoir

129

temperature, Ti is he injected fluid temperature; and ff is the elastic term form factor, which

130

typically ranges between 0.5 and 1, depending on the temperature front [10].

131

The total vertical stress can be calculated as:

132

 v   w LA   r Sot (3)

6

133

where  w is the water unit weight;  r is the lithostatic (total) unit weight; LA is the water depth;

134

and Sot is the overburden thickness. The minimum total horizontal stress can be estimated by the

135

following expression:

136

 h  K0 ( v  Pe )  Pe

(4)

137

where K0 correlates the minimum horizontal stress with the vertical stress, which is also known as

138

trust coefficient (also known as coefficient of earth pressure at rest). Note that K0 is used (strictly)

139

to relate the horizontal effective stress with the vertical effective stress. In this work K0 was also

140

used with the stresses definitions adopted in equations (1) and (4), which are based on Biot’s

141

stresses.

142

2.1.2

143

To illustrate the use and significance of the previous equations a simple application example is

144

presented in this section. Typical parameters and properties for oil reservoirs were adopted in this

145

analysis. They are listed in Table 1. Note that a quite simple approach was adopted in this paper

146

to calculate the stress state and to estimate some model parameters. The technique suggested in

147

this paper could also be coupled with more advanced methods (e.g. [16]) to deal with uncertainties

148

in reservoir geomechanics.

149

The calculated fracture pressure based on these properties and using the equations presented in the

150

previous section is equal to 60.42 MPa. This is a deterministic analysis and then unable to quantify

151

the uncertainties associated with this problem. In Section 4.1, this case is solved using the evidence

152

theory incorporating the quantification of uncertainties in the analysis.

153

Application case

<< Insert Table 1 here >>

7

154

2.2

155

The terms subsidence and compaction have different meanings. Compaction is related to the

156

thickness reduction in a given formation, while subsidence refers to a downward movement of the

157

surface. The phenomenon of subsidence occurs over a much larger area than the compaction one,

158

which is associated with the reservoir thickness. The pore pressure reduction during production

159

leads to important changes in the reservoir and adjacent rocks. Such changes in pore pressure affect

160

directly the compaction and subsidence phenomena, which in turns significantly impact on the

161

fluid flow in the reservoir.

162

Several authors have discussed the relevance of the understanding the phenomena associated with

163

reservoir compaction and subsidence during oil fields exploitation (e.g. [17]; [18]; [19]; [20]). In

164

particular, Ferreira [21] conducted an extensive review on cases of subsidence around the world

165

and compared some analytical models with numerical solutions using a finite difference

166

commercial code. In the following section the equations to estimate the compaction and subsidence

167

of an oil reservoir under depletion are presented.

168

2.2.1

169

The vertical displacement field proposed by Geertsma [17] considers the geometry of the disk-

170

shaped reservoir as illustrated in Figure 1.

Reservoir compaction

Vertical displacement calculation

171 172

<< Insert Figure 1 here >> The vertical displacement is calculated via:

cm hR  C ( Z  1) (3  4 )C ( Z  1)    2 2 1/2 2  [1  C ( Z  1) ] [1  C 2 ( Z  1) 2 ]1/2  2CZ  (3  4    )  p [1  C 2 ( Z  1) 2 ]3/2 

uz (0, z )   173

8

(5)

174

where (see Figure 1) Z = z/c; C = c/R and ε = -1 when z > c, and ε = +1 when z < c; being c is the

175

distance between the free surface and the point under analysis The elastic constant cm is defined

176

by:

cm 

177

 (1  2 ) 2G(1   )

(6)

178

where G is the shear modulus.

179

Note that this equation provides the vertical displacement at different positions, therefore it is possible to

180

estimate the subsidence (i.e. the displacement at the seafloor or groundfloor level, depending if the case

181

under study corresponds to an off-shore or an on-shore case, respectively), and the compaction (as the

182

difference in displacements between the top and the bottom of the reservoir). Also in this case a relatively

183

simple model was selected to show how the selected approach can deal with uncertainties in reservoir

184

geomechanics related problems. The proposed approach is independent of the adopted equation and

185

certainly more complex models can be used in future works.

186

2.2.2

187

An example of application using the properties listed in Table 2 is presented in this section to

188

illustrate the use of the Equation 5.

Application example

189

<< Insert Table 2 here >>

190

Figure 2 presents the variation of the vertical displacement with depth for a reservoir located at a

191

depth of 2900 meters. Note that the model predicts that the top of the reservoir will move down

192

while the reservoir base will move up. The reservoir thickness reduction is around 2.7 m (i.e.

193

displacement at the top ~ +1.40 m, minus the displacement at the bottom ~ -1.3 m), while the

194

subsidence is in the order of 20 centimeters.

9

195

This analysis is deterministic and again is not possible to quantify the uncertainties associated with

196

it. The inference of new sampled data for the reservoir compaction problem and their impact on

197

the uncertainty quantification is discussed in Section 4.2.

198

<< Insert Figure 2 here >>

199

3.

200

Uncertainty quantification (UQ) is a key component when analyzing problems with limited data

201

and/or high variability of the available information. In the last few years the development of new

202

and more efficient numerical techniques, the increase in computational power, and the massive

203

use of parallel computing, have facilitated the incorporation of UQ analyses in routinely studies.

204

As mentioned before, there are two main types of uncertainties: aleatory and epistemic

205

uncertainties. The probability theory is commonly adopted to evaluate both types of uncertainties.

206

However, recent studies have shown that probabilistic methods may not be the most convenient

207

ones for analyses involving epistemic uncertainties. Alternative approaches, based for example on

208

evidence theory seem more appropriate for handling reducible uncertainties [6].

209

The prior probability distribution (called ‘prior’ henceforth) of a variable is a basic information

210

required in analysis based on probabilistic methods. The “Principle of insufficient reason” [22] is

211

generally used when this information is not available. This implies that when the density functions

212

associated with the model parameters are not known, uniform distribution functions could be

213

assumed. This assumption is valid for random variables, but may be inaccurate when dealing with

214

epistemic ones.

215

An alternative approach to quantify epistemic uncertainties is based on the evidence theory. This

216

theory is particularly useful in those cases where there is not enough information to quantify the

217

uncertainty with a known density function. In the evidence theory the range of the variables is

UNCERTAINTY QUANTIFICATION

10

218

defined by intervals or sets of possible values. The definition of a range (or a set) of probable

219

values linked with a given variable has three important implications:

220

i) it is possible to incorporate in the analysis experimental information from different

221

sources that can be weighted according to the user confidence on particular test results;

222

ii) the principle of insufficient reason is not enforced. Furthermore there is not a pre-

223

established structure for the distribution of the input variables. Initial probabilities can be

224

arranged in sets, without having pre-established assumptions about the probabilities of

225

individual events;

226

iii) the additivity axiom is not mandatory. Therefore, the complementary sets of

227

probabilities measures do not have to add up to one. When the additivity axiom is

228

considered in analysis based on the evidence theory this framework converges to the

229

traditional probabilistic method [6].

230

By comparing the probability and the evidence theories it is possible to see that the benefit of the

231

latter is that it allows a less restrictive description of the uncertainties. However, a drawback of the

232

evidence theory is that uncertainty analyses are computationally more demanding than the ones

233

based on probabilistic method. The additional computational effort comes mainly from the

234

numerical optimization process needed to calculate the belief and plausibility functions. In the

235

following section some background information related to the evidence theory is discussed, more

236

details can be found elsewhere (e.g. [23], [24], [7]).

237

3.1

238

The two core concepts in evidence theory are: i) the basic probability assignment (BPA) or

239

weights, and ii) the focal elements. The BPA for a given set is the probability that can be assigned

240

to it. This probability cannot be decomposed into additional probabilities for subsets of the original

Belief and plausibility functions

11

241

set. The focal elements are those sets that have nonzero BPAs. This then leads naturally to a clear

242

distinction between an evidence space for a set of possible outcomes and a probability space for a

243

set of possible outcomes. The basic probability assignment is different from the classical definition

244

of probability. It is defined by mapping over an interval in which the basic assignment of the null

245

set is ‘0’ and the summation of basic assignments in a given set is ‘1’ [6].

246

The first work in evidence theory was done by Shafer [25], as an expansion of the one previously

247

performed by Dempster [26]. In a finite discrete space, the evidence theory or the Dempster-Shafer

248

theory (DST), can be interpreted as the generalization of the probability theory for the case in

249

which the probabilities are assigned to sets of values. When the evidence is sufficient to assign

250

probabilities to single events, the Dempster-Shafer model falls into the traditional probabilistic

251

formulation.

252

Let’s consider a model defined by:

253

y  f (x)

(7)

254

where x  [ x1 , x2 , x3 ,..., xnX ] is the vector of the input variables and y  [ y1 , y2 , y3 ,..., ynY ] is the

255

vector associated with the model results. Thus, the uncertainty in y can be estimated in the

256

framework of the probabilistic theory by the Cumulative Distribution Function (CDF).

257

prob( y  y)    ( f (x) | y)d X dX  i S1 ( f (x) | y) / nS n

P

(8)

258

where y is the expected value (i.e. the different outputs associated with each ones of the possible

259

realizations); nS is the number of samples;  ( f (x) | y) = 1 if f (x)  y and  ( f (x) | y) = 0 if

260

f (x)  y . Furthermore, the probability theory requires that the uncertainty associated with x

261

should be defined by a density function (e.g. normal, triangular or uniform). To characterize the

12

262

uncertainty in x is necessary to establish a probability space (XP; XP’;mPX) where XP is the sample

263

space of possible values of x; XP’ is the set of subsets of XP chosen appropriately (Sigma-algebra);

264

and mPX is a probability factor.

265

Conversely the evidence theory does not impose a rigid structure for the characterization of the

266

uncertainties. It is a flexible approach. The uncertainty in y is estimated by means of two

267

uncertainty functions, the belief and plausibility functions as follows [6]:

268

269

BelY ( y  y) 

PlY ( y  y) 

m

A Ay



A Ay  0

EY

( A)

(9)

mEY ( A)

(10)

270

At an intuitive level, belief is a measure of the amount of information that indicates that a statement

271

is true, and plausibility is a measure of the amount of information that indicates that a statement

272

could be true. Central to this concept is the idea that all the probability distributions that do not

273

violate the assumed properties of the evidence space are under consideration. The ‘ball-box’

274

analogy of the Dempster-Shafer theory can assist to a simple interpretation of this technique. In

275

this model is assumed that n unmarked steel balls are distributed among k boxes A1,. . . , Ak, with

276

mi representing the fraction of balls in Ai. The boxes are placed in a bigger box U and are allowed

277

to overlap. The position of each ball within its box is unspecified. In this model, the granular

278

distribution A describes the distribution of the balls among the boxes. Note that the number of balls

279

in Ai is unrelated to the ones in Aj. It is important to distinguish between the number of balls in Ai

280

and the ones in Aj, because these fractions might overlap (i.e. the boundary of the boxes are

281

penetrable). Now, given a region Q in U, it is possible to ask the following question: How many

282

balls are in Q? Because the information regarding the position of each ball is incomplete, the

13

283

answer to this question will be (in general) interval-valued. The upper bound can readily be found

284

by visualizing Q as an attractor, for example, a magnet. Under this assumption, the proportion of

285

balls drawn into Q is given by Eq. (10), which is the expression for plausibility in the Dempster-

286

Shafer theory. Similarly, the lower bound results from visualizing Q as a repeller. In this case, the

287

lower bound is given by Eq. (9); which is the expression for belief in the Dempster-Shafer theory.

288

The combination of the intervals defined as uncertain values will outline the upper and lower

289

bounds of the response as described in the previous paragraph. The definition of the belief and

290

plausibility functions for the model outputs can be done through an optimization process, as

291

discussed in the following section. More details can be found elsewhere (e.g. [24], [6], [7]).

292

3.2

293

The belief and plausibility functions are defined for the input variables (x). For each input variable

294

it is necessary to specify intervals (and the associated bounds) and the corresponding weights (i.e.

295

BPAs). To calculate the belief and plausibility measures it is necessary to calculate the minimum

296

and maximum of the response function for each interval cell combination. For example, if a case

297

has 2 input variables and each of these is defined by 3 intervals, the number of combination cells

298

is equal to 9. This implies that epistemic analyses require the implementation of an optimization

299

algorithm to find the maximum and minimum for all the possible combinations considered. More

300

details about the implementation of typical optimization methods to perform Dempster-Shafer

301

analysis can be found elsewhere (e.g. [27], [28]; 23]; [7]). The intervals and their associated BPAs

302

are then propagated to obtain cumulative distribution belief and plausibility functions (i.e. CBF

303

and CPF, respectively). As mentioned above, belief is the lower bound on a probability estimate

304

that is consistent with the evidence, and plausibility corresponds to the upper bound on a

305

probability estimate that is consistent with the evidence.

Optimization algorithm

14

306

In this work the approach discussed in Cabral et al. [7] to compute the CBF and CPF is adopted,

307

which consists of the main following steps:

308

1) Combine intervals of each input variable into combination cells.

309

2) For each cell, two optimization problems have to be solved to find the lower (lb) and

310

upper (ub) bounds of the response.

311

3) Sort the lower bounds of the output response intervals obtained in Step 2 in ascending

312

order.

313

More details about this optimization process are presented in Cabral et al. [7].

314

3.3

315

To establish the type of analysis to be performed it is important to pay attention not only to the

316

availability of data but also to the origin and reliability of this information. There are two critical

317

aspects related to the combination of evidence obtained from different sources. The first one is

318

related to the type of evidence involved in the study and the second aspect is based on how to

319

handle conflicting evidence. Figure 3 presents four different types of evidence that impact on the

320

choice of how such evidences must be combined. They are called: inclusive evidence, consistent

321

evidence, arbitrary evidence and disjointed evidence [29].

Types of evidence

322

<< Insert Figure 3 here >>

323

The inclusive evidence can be represented by subsets that are contained in a larger subsets, which

324

in turns can be part of a larger one (Figure 3a). For example, this type of evidence can be related

325

to the case where the information for a given set is obtained from a larger period of time by

326

reducing the intervals of evidence. In the case of the consistent evidence, at least one element is

327

common to all subsets of the sample space (Figure 3b). The arbitrary evidence corresponds to the

328

situation where there is no element in common for all the subsets (Figure 3c), although there may 15

329

be some subsets of elements in common. Finally, the disjointed evidence is associated with the

330

case in which for any two subsets of the sample space there are no elements in common. (Figure

331

3d).

332

Each of these possible configurations of evidence from different sources has different implications

333

on the level of conflict associated with each situation. It is intuitive that for the case of disjointed

334

evidence, all sources of evidence are in conflict. For the case of arbitrary evidence, there are some

335

agreements between arbitrary sources, but there is no consensus on any of the elements. The

336

consistent evidence implies an agreement between at least one subset of evidence. Inclusive

337

evidence represents a situation in which every subset is included in a larger one, i.e., it is

338

guaranteed that there is a correlation between the evidences. However, there is inherent conflict in

339

the additional evidence supported by the larger subsets.

340

The traditional probability theory cannot deal with inclusive, consistent or arbitrary evidences,

341

without applying the definition of probability density functions for all the elements of the sets.

342

Furthermore, it is not possible to express the level of conflict between these sets of evidences. The

343

following examples show how the evidence theory is more flexible when dealing with the

344

scenarios discussed above. Furthermore, in the evidence theory, there are several ways to

345

incorporate imminent conflicts in the process of combining multiple sources of information. In the

346

next section, the application of this method is conducted based on the two problems discussed in

347

Section 2.

348 349

4.

350

In reservoir geomechanics there is a tendency to develop increasingly sophisticated constitutive

351

models and coupled formulations. This is a very valuable effort, because the behavior of sediments

DECISIONS MAKING PROCESS UNDER UNCERTAINTY

16

352

and rock involved in this type of problems is generally very complex. However, it seems also very

353

important to assess the impact of the variation of material parameters in the analysis and the

354

associated uncertainties. This section is aimed at studying decision making cases incorporating

355

uncertainty quantification via the evidence theory and adopting simple solutions, as the ones based

356

on analytic or semi-analytical solutions presented in Sections 2.1 and 2.2.

357

Decision-making is certainly the most important task of a manager and it often tends to be rather

358

difficult. The type of approaches used to make the decision ranges between two extreme cases.

359

One of them corresponds to the deterministic solutions (as the ones presented in the previous

360

Sections), where specific (determined) values are assigned to the different parameters and then it

361

is not possible to quantify the uncertainties. The other extreme corresponds to the pure uncertainty

362

analysis, where there is not previous information about the parameters associated with the case

363

under study. Those studies in which the risk can be assessed lie between these two extreme types

364

of analyses [30].

365

During reservoir operation it is possible that additional information related to the material

366

properties become available. It is of great interest to study how this new information may impact

367

on the uncertainty studies. In the following section an analysis under pure uncertainty (i.e. lack of

368

input data) is discussed first. Then, the incorporation of increasingly new input data in the analyses

369

and the corresponding effect in the uncertainty quantification is discussed.

370 371 372

4.1

373

The lack or limited of material data is a recurring problem associated with reservoir geomechanics.

374

One of the main reasons related to it has been the high cost associated with the gathering of the

Lack of input data

17

375

data from the field. Furthermore, the reduction of the geomechanical data gathered via seismic

376

logs or cores is quite complex and require high technical expertise.

377

In this section the calculation of the fracture pressure is used as an example of the decision making

378

process incorporating uncertainty quantification in a case in which the input data is quite scarce.

379

Afterwards, the discussion focuses on the comparisons between the results obtained via probability

380

and evidence theories. In this example, three experts were consulted to define the ranges of

381

parameters to be used in the UQ analysis. These experts defined the weights required for the use

382

of the evidence theory. The ranges of the parameters were defined by the material properties found

383

in analogous reservoirs. The uncertainty parameters are: Young’s modulus, Poisson’s coefficient,

384

Biot’s coefficient and thrust coefficient. The same weight value for each expert opinion was

385

assumed, therefore the BPAs are equal to 1/3. Table 3 presents the experts opinions related to these

386

parameters.

387

<< Insert Table 3 here >>

388

The variables quoted above are considered epistemic and the evidence theory is a proper

389

framework for the treatment of this kind of uncertainty. Figure 4 presents the CBF and CPF of the

390

fracture pressure calculation using the parameters shown in Table 3.

391

One of the main dilemmas associated with the use of the probability theory for uncertainty

392

quantification of epistemic variables is related to the choice of the probability density functions.

393

Figure 3 shows three CDF's associated with the choice of an uniform probability density function

394

(in dark blue), normal (in light blue) and a distribution that considers the BPAs as weights (in

395

green) in the respective intervals.

396

<< Insert Figure 4 here >>

18

397

This figure can be used to analyze the injection pressure that need to be applied for a given

398

reservoir to induce the fracture propagation. These plots can be used in different ways, for example

399

let’s assume that 0.50 is the largest possible probability that an injection project manager would

400

like to take to assure that injection pressure is equal or above the fracture pressure propagation and

401

in this way maintain a good injectivity. Considering first the CDFs (i.e. the ones obtained with the

402

probabilistic method), the needed increment in the injection pressure is around 69.63 MPa if

403

uniform or normal density functions are assumed (note that only for a probability of 0.5 these two

404

solutions predict the same pressure, for other probabilities the injection pressures estimated by

405

these two functions are different), while the other CDF predict a bit lower value of injection

406

pressure, around 68.65 MPa. As for the evidence theory, it is important to remember that Bel(U) is

407

the smallest possible probability that could occur for U over the set of indicated distributions, and

408

Pl(U) is the largest possible probability that could occur for U over the set of indicated

409

distributions. Therefore looking at the CBF and CPF curves, one may conclude that based on the

410

existing evidences it is believed that the fracture will propagate with a confidence of 0.5 if the

411

injection pressure is above 63.74 MPa. However the available evidence also informs that it is

412

plausible that an injection pressure as high as 75.02 MPa may be necessary to induce the

413

propagation of the fracture in this reservoir (also in this case with a confidence of 0.5).

414

It can be seen that the evidence theory provides more complete information to assist the decision

415

maker, with a range of variation of the possible values of injection pressures that will induce the

416

fracture propagation when a give probability is set. This interval of variation of the injection

417

pressure is a clear indicator of the uncertainties associated with the problem under study and with

418

the available data. The example discussed in the next section shows how the uncertainty can be

19

419

reduced when more information about the geomechanical properties of the reservoir is added into

420

the analysis.

421

4.2

422

Table 4 corresponds to a possible scenario for information availability over time for reservoir

423

compaction problem analyzed in Section 2.2. At the beginning, there was no information about

424

the mechanical properties of the reservoir. The parameters used in this first analysis were estimated

425

from the data observed in similar reservoirs. Then, information obtained from the first well drilled

426

became available. This data was obtained from log correlations. Results from laboratory tests were

427

available afterwards. Finally, another well was drilled and therefore more information about

428

reservoir properties was available.

New data inference

429

<< Insert Table 4 here >>

430

A crucial point at this stage of the analysis is to define how this new information is incorporated

431

in the process of uncertainty quantification. It is also very important to have the opportunity of

432

weighting the available information based on the reliability of this data. As discussed previously,

433

the evidence theory is flexible and allows defining probabilities by sets, which is very convenient

434

for handling the information associated with this type of problem. In the evidence theory the expert

435

can decide how to measure the significance of the existing information coming from different

436

sources (i.e. by the definition of the BPAs or weights). The weighting of the available data is based

437

on the reliability of the existing data, experts’ judgment, and previous experience.

438

In Table 4 it can be observed that for the analysis identified as ‘T0’, the BPA assigned to the

439

different properties is equal to 1. This is because there is only one source of information (i.e. the

440

experience from previous studies similar to this one). In the analysis coded as ‘T1’ (i.e. the case in

441

which information from the first well became available), a BPA of 0.8 was assigned to the data 20

442

coming from the new well, and 0.2 was assigned to the T0 data. This implies that according to the

443

expert opinion, the information from this reservoir is more reliable (i.e. higher weight value) than

444

the one coming from previous similar studies. It can also be observed that the reservoir parameters

445

coming from the well vary in a smaller range. Also, according to the experts’ opinion, the

446

information gathered in the laboratory and incorporated in the analysis ‘T2’is more reliable than

447

the one related to the previous studies, but significantly less dependable than the data associated

448

with the borewell. The weights assigned to the previous studies, well, and lab data are: 0.1, 0.7 and

449

0.2; respectively (note that the sum of the BPAs has to be always equal to 1). A similar reasoning

450

is followed in analyses T3, when data coming from another well is incorporated in the analyses.

451

Figure 5 shows the changes in the curves associated with the uncertainty quantification analyses

452

when new information is incorporated at different stages of the analysis. The flexibility of the

453

evidence theory for new data inference is clear in the evaluation of the results. It is not surprising

454

that new information about the material parameters reduces the uncertainty range associated with

455

the estimated subsidence. Initially, when few information was available (i.e. based mainly on

456

previous experiences in similar reservoirs), the range of possible subsidence values was quite wide,

457

i.e. between 0.05 and 0.9 m. After incorporating information from the first well, the uncertainty

458

decreased and the calculated subsidence was between 0.05 to 0.7 meters. Once the information

459

from the laboratory test became available, a similar trend was observed with a further reduction of

460

the uncertainty range. The minimum and maximum possible subsidence in this new study was in

461

the range between 0.05 and 0.52 m, respectively. Finally, after the information from the drilling of

462

the second well was incorporated, the uncertainty reduced even more, and the possible subsidence

463

values were between 0.05 to 0.18 m.

464

<< Insert Figure 5 here >>

21

465

It is worth mentioning that the use of the evidence theory in this case allowed the inference of new

466

data without the requirement of defining priors, as it is necessary in classical probability analyses.

467

It was observed that the curve related to the probability theory considering uniform distribution

468

kept constant during the subsequent analyses, because the range that defines the minimum and

469

maximum values of the parameters were not modified as new information became available, and

470

consequently the uniform distribution did not change (Figure 5).

471

5.

472

This work focused on the uncertainty quantification in problems related to reservoir geomechanics.

473

These types of problems are characterized by the limited (and in some occasions nil) information

474

associated with material properties. To quantify the uncertainties a formal framework based on the

475

evidence theory was adapted to oil reservoir problems. This non-probabilistic method is especially

476

well-suited for dealing with uncertainty that are epistemic in nature. This type of reducible

477

uncertainty is very common in reservoir geomechanics because it is generally associated with the

478

lack of information. The main components of the adopted approach was presented in this work

479

and then it was applied to solve typical problems related to oil reservoirs, such as the estimation

480

of the fracture propagation pressure and reservoir compaction. Relatively simple analytical

481

solutions were adopted for the analyses, because the main effort was on exploring the capability

482

of the framework for dealing with uncertainties.

483

In both analyses (i.e. estimation of fracture pressure and reservoir compaction) it was observed

484

that the evidence theory facilitates the decision making process because it provides more complete

485

information than the probabilistic theory. For example, when defining the fracture propagation

486

pressure, the evidence theory gives based on the available information and for a given probability,

487

a range of variation of possible injection pressures that may induce fracture propagation. This

SUMMARY AND CONCLUSIONS

22

488

interval of variation for the injection pressures is a clear indicator of the uncertainties associated

489

with this problem. It was also shown that the evidence theory is a suitable framework for

490

incorporating consistently additional evidences and reducing consequently the uncertainty when

491

more information about the reservoir becomes available. This feature was evident when analyzing

492

the reservoir compaction problem, where the range of variation of the estimated settlements was

493

reduced in subsequent analyses that incorporated increasingly additional data about the

494

geomechanical properties of the reservoir.

495

The use of the evidence theory based on CBFs and CPFs, guarantee that the information is

496

consistent with all the known data associated with the input parameters of the model. In the limit,

497

these two curves (i.e CBF and CPF) converge to a single cumulative distribution function curve,

498

when the data that completely describes the model is available. It is worth mentioning that the use

499

of evidence theory in this case allowed the inference of new data without the requirement of the

500

definition of priors (as generally happens in probabilistic theories). It was observed that the curve

501

related to the probability theory considering uniform distribution kept constant during the analysis

502

because the range that defines the minimum and maximum values of the parameters were not

503

modified as new information became available (and thus the uniform distribution did not be

504

change). Furthermore, the evidence theory was able to properly incorporate in the analyses the

505

experts’ opinion and to weight the available information based on the reliability of the data source.

506

ACKNOWLEDGEMENTS

507

The authors acknowledge the financial support from PETROBRAS, CNPq (Brazilian National

508

Research Consul) and Foundation CMG (Chair on Reservoir Simulation at UFPE).

23

509

REFERENCES

510

[1]

Uusitalo L, Lehikoinen A, Hellec I, Myrberga K. An overview of methods to evaluate

511

uncertainty of deterministic models in decision support. Environmental Modelling &

512

Software. 2015;63:24–31.

513

[2]

514 515

& Francis. 2012. [3]

516 517

Fenton N, Neil M. Risk Assessment and Decision Analysis with Bayesian Networks. Taylor

Regan H.M, Colyvan M, and Burgman M.A. A Taxonomy and Treatment of Uncertainty for Ecology and Conservation Biology. Ecological Applications. 2002;12:618-628

[4]

Skinner D, Rocks S, Pollard S, Drew G. Identifying Uncertainty in Environmental Risk

518

Assessments: The Development of a Novel Typology and Its Implications for Risk

519

Characterization. Human and Ecological Risk Assessment: An International Journal.

520

2013;20:607-640.

521

[5]

Walker W, Hattemoes P, Rotmans J, Van Der Sluijs, J, Van Asselt, M, Janssen P, Krayer

522

Von Krauss M.. Defining Uncertainty: A Conceptual Basis for Uncertainty Management in

523

Model-Based Decision Support. Integrated Assessment. 2003;4:5-17.

524

[6]

Helton J, Johnson J, Oberkampf W. Storlie C. A sampling-based computational strategy for

525

the representation of epistemic uncertainty in model predictions with evidence theory.

526

Computer Methods in Applied Mechanics and Engineering. 2007;196:3980-3998.

527

[7]

Cabral L, Guimarães L, Horowitz, Sánchez M. Coupled Hydro-Mechanical Fault

528

Reactivation Analysis Incorporating Evidence Theory for Uncertainty Quantification.

529

Computers and Geotechnics. 2014;56: 202–215.

530 531

[8]

De Souza A, Fernandez P, Mendes R, Rosa A, Futado C. The Impact of Fracture Propagation on Sweep Efficiency During a Waterflooding Process. SPE Latin American

24

532

and Caribbean Petroleum Engineering Conference held in Rio de Janeiro, Brazil, 20 – 23

533

June 2005. Rio de Janeiro: Society of Petroleum Engineers.

534

[9]

535 536

Hydraulic Fractures. SPE Journal. 1980; 20(4), 293-303. [10]

537 538

Hagoort J, Weatherill B, Settari A. 1980. Modeling the Propagation of Waterflood-Induced

Perkins T, Gonzalez J. The Effect of Thermoelastic Stresses on Injection Well Fracturing. OLD SPE Journal. 1985

[11]

Saripalli K, Bryant S, Sharma M. Role of Fracture Face and Formation Plugging in

539

Injection Well Fracturing and Injectivity Decline. SPE/EPA Exploration and Production

540

Environmental Conference. Austin, Texas, 28 February–3 March 1999. Society of

541

Petroleum Engineers. 1999.

542

[12]

Suri A, and Sharma M. Fracture Growth in Horizontal Injectors. SPE Hydraulic Fracturing

543

Technology Conference. The Woodlands, Texas, USA, 19–21 January 2009. Society of

544

Petroleum Engineers. 2009

545

[13]

Rahman K, Khaksar A. Fracture Growth and Injectivity Issues for Produced Water

546

Reinjection Wells - Case Studies with Fields from offshore Australia and UK North Sea.

547

SPE Asia Pacific Oil and Gas Conference and Exhibition. Perth, Australia, 22–24 October

548

2012. Society of Petroleum Engineers. 2012.

549

[14]

Schmitt D, Zoback M. Poroelastic effects in the Determination of the Maximum Horizontal

550

principal stress in hydraulic fracturing tests - A proposed breakdown equation employing

551

a modified effective stress relation for tensile failure. Int. J. Rock Mech. Min. Sci. &

552

Geomech. 1989;26(6):499-506.

553

[15]

Yew C, Weng X.. Mechanics of Hydraulic Fracturing, Elsevier Science. 2014

25

554

[16]

Della Vecchia G, Pandolfi A, Musso G, Capasso G. An analytical expression for the

555

determination of in situ stress state from borehole data accounting for breakout size.

556

International Journal of Rock Mechanics & Mining Sciences. 2014;66: 64–68.

557

[17]

558 559

Geertsma, J. Problems of Rock Mechanics In Petroleum Production Engineering. 1st ISRM Congress. Lisbon, Portugal: International Society for Rock Mechanics. 1966

[18]

Zimmerman R. Implications of Static Poroelasticity for Reservoir Compaction. Pacific

560

Rocks 2000. Girard, Liebman, Breeds & Doe (eds) Rotterdam, Netherlands: American

561

Rock Mechanics Association. 2000

562

[19]

Bruno M. Geomechanical and Decision Analyses for Mitigating Compaction-Related

563

Casing Damage. SPE Annual Technical Conference and Exhibition. New Orleans, USA.

564

Society of Petroleum Engineers. 2002

565

[19]

Holt R, Flornes O, Fjaer, E.. Consequences Of Depletion-Induced Stress Changes On

566

Reservoir Compection And Recovery. 6th North America Rock Mechanics Symposium

567

(NARMS): Rock Mechanics Across Borders and Disciplines. Houston, Texas, June 5 - 9,

568

2004: American Rock Mechanics Association. 2004

569

[20]

Schutjens P, Hansen H, Hettema M, Merour J, De Bree P, Coremans J, Helliesen G.

570

Compaction-Induced Porosity/Permeability Reduction in Sandstone Reservoirs: Data and

571

Model for Elasticity-Dominated Deformation. SPE Annual Technical Conference and

572

Exhibition. New Orleans, USA: Society of Petroleum Engineers. 2004

573

[21]

574 575

Ferreira F. Subsidence due to Reservoir Compaction: From Theory to Well Integrity Problems. DSc, Federal University of Rio de Janeiro. 2014.

[22]

Savage L. The Foundations of Statistics. Dover Publications. 2012

26

576

[23]

Helton J, Johnson J, Oberkampf W. An exploration of alternative approaches to the

577

representation of uncertainty in model predictions. Reliability Engineering & System

578

Safety. 2004;85:39-71.

579

[24]

580

Helton J, Oberkampf W, Johnson J. Competing Failure Risk Analysis Using Evidence Theory. Risk Analysis: An International Journal. 2005;25:973-995.

581

[25]

Shafer G. A Mathematical Theory of Evidence. Princeton University Press. 1976

582

[26]

Dempster A. Upper and Lower Probabilities Induced by a Multivalued Mapping. The

583 584

Annals of Mathematical Statistics. 1967;38(2):325-339. [27]

Waltz R, Morales J, Nocedal J, Orban D. An interior algorithm for nonlinear optimization

585

that combines line search and trust region steps. Mathematical Programming.

586

2006;107:391-408.

587

[28]

588 589

nonlinear programming. Mathematical Programming. 2000;89:149-185. [29]

590 591 592

Byrd R, Gilbert J, Nocedal J. A trust region method based on interior point techniques for

Sentz K, Ferson S. Combination of Evidence in Dempster-Shafer Theory. Sandia National Laboratories. SAND 2002-0835. 2002

[30]

Ben-Haim Y. Info-Gap Decision Theory: Decisions Under Severe Uncertainty. Elsevier Science. 2006.

593

27

Figure

Sea or ground level

Figure 1: Geometry for determining the displacement field in the surroundings of a disc-shaped reservoir (adapted from Geertsma [17]).

Figure 2: Vertical displacement field along the depth. The reservoir is located at 2900 m depth.

.

A E

A

B

B

C

C

E

D

b)

a)

D

A

B

C

A C

E

E

B

D

D c)

d)

Figure 3: Types of evidence: a) inclusive; b) consistent; c) arbitrary and d) disjointed. The letters A, B, C, D and E represent diferent sources of evidence.

Figure 4: Fracture pressure calculation. Uniform probability density function (in dark blue), normal (in light blue) and a distribution that considers the BPA's as weights in the respective ranges (in green). Results obtained with evidence theory: CBF (in red) and CPF (in black).

Figure 5: Incorporation of new data over time. Uniform probability density function (green curves), and a distribution that considers the BPA's as weights (blue curves) in their respective ranges. Via theory of evidence it is estimated the CBF (in red) and the CPF (in black).

Table

Table 1: Input data used for the calculation of the reservoir fracture pressure

Geomechanical data Young modulus (E)

25 GPa

Biot’s coefficient (α)

0.8

Tensile strength (σT)

4.02 MPa

Poisson’s coefficient (υ)

0.25

Thrust coefficient (K0)

0.8

Pore pressure factor in tensile failure criterion (φ)

0.07 Length data

Water depth (LA)

2100 m

Overburden tickness (Sot)

2450 m

Reservoir thickness (hR)

100 m Temperature and pressure data

Reservoir temperature (Tr) Injected water temperature (Ti) Reservoir initial pressure (Pe)

60 oC 35 oC 48.05 MPa

Table 2: Required input data for the estimation of the vertical displacement field in a depleted reservoir

Geomechanical data Shear Modulus (G)

147.1 MPa

Biot’s Coefficient (α) Poisson’s Coefficient (υ)

1.0 0.286 Reservoir properties

Reservoir thickness (hR)

100 m

Top (c)

2900 m

Radius (R) Total depth (z)

1000 m 4000 m

Porepressure variation (depletion) (Δp)

13.1 MPa

Table 3: Ranges of possible values characterizing the uncertainty of the input variables. Each interval was defined by different experts’ opinions.

Parameter

Expert 1

Expert 2

Expert 3

E (GPa) υ

35 - 40 0.2 - 0.3

35-45 0.2 - 0.3

30 - 40 0.2 - 0.3

α

0.6 - 0.8

0.6 - 0,8

0.5 - 0.7

K0

0.7 - 0.9

0.75 - 0.8

0.7 - 0.8

BPAs

0.33

0.33

0.33

Table 4: Information (evidence) over time. The uncertainty quantification process was modified with each new piece of evidence.

T = 0 - Similar reservoirs Parameter Α

T0 0.6 - 1.0

Υ

0.2 - 0.4

G (MPa)

49 - 294

BPA0

1.0 T = 1 - Well drilled

Parameter Α

T0 0.6 - 1.0

T1 0.7 - 0.8

Υ

0.2 - 0.4

0.32 - 0.36

G (MPa)

49 - 294

245.2 - 313.8

BPA1

0.2

0.8 T = 2 - Lab Test

Parameter Α

T0 0.6 - 1.0

T1 0.7 - 0.8

T2 0.95 – 1.0

Υ

0.2 – 0.4

0.32 - 0.36

0.35 – 0.4

G (MPa)

49 - 294

245.2 - 313.8

196.1 - 214.6

BPA2

0.1

0.7

0.2

T = 3 - New Well Drilled Parameter

T0

T1

T2

T3

Α

0.6 - 1.0

0.7 - 0.8

0.95 – 1.0

0.75 – 0.9

Υ

0.2 – 0.4

0.32 - 0.36

0.35 – 0.4

0.3 – 0.35-

G (MPA)

49 - 294

245.2 - 313.8

196.1 - 214.6

294 – 295.2

BPA3

0.02

0.4

0.18

0.4