You are on page 1of 17

Journal of Energy Resources Technology.

Received December 29, 2018;


Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
Drilling rate of penetration prediction of high-angled wells using artificial neural
networks
Ahmed K. Abbas*
Iraqi Drilling Company, Basra 61004, Iraq.
Email: akayr4@mst.edu
ahmed.khudair.abbs@idc.gov.iq
Tel: + 1 5732010883
*Corresponding author
Salih Rushdi

d
College of Engineering, University of Al-Qadisiyah, Al-Qadisiyah 58002, Iraq.

te
Email: salih.rushdi@qu.edu.iq
Tel: + 964 7901230881

di
Mortadha Alsaba
Australian College of Kuwait, Safat 13015, Kuwait.

e
Email: m.alsaba@ack.edu.kw

py
Tel: + 965 98059212
Mohammed F. Al Dushaishi

Co
School of Engineering, Texas A&M International University, Laredo, Texas 78041, United States.
Email: mohammed.aldushaishi@tamiu.edu
Tel: +1 5733561764

ot
ABSTRACT
Predicting the rate of penetration (ROP) is a significant factor in drilling optimization and minimizing expensive drilling
tN
costs. However, due to the geological uncertainty and many uncontrolled operational parameters influencing the ROP, its
prediction is still a complex problem for the oil and gas industries. In the presented study, a reliable computational approach
for prediction of ROP is proposed. Firstly, fscaret package in R environment was implemented to find out the importance and
rip

ranking of the inputs parameters. According to the feature ranking process, out of the 25 variables studied, 19 variables had
the highest impact on ROP based on their ranges within this dataset. Secondly, a new model that is able to predict the ROP
using real field data, which is based on artificial neural networks (ANNs), was developed. In order to gain a deeper
sc

understanding of the relationships between input parameters and ROP, this model was used to check the effect of the weight
on bit (WOB), revolutions per minute (RPM), and flow rate (FR). Finally, the simulation results of three deviated wells
showed an acceptable representation of the physical process, with reasonable predicted ROP values. The main contribution of
nu

this research as compared to previous studies is that it investigates the influence of well trajectory’s (azimuth and inclination)
and mechanical earth modeling parameters on the ROP for high-angled wells. The major advantage of the present study is
Ma

optimizing the drilling parameters, predicting the proper penetration rate, estimating the drilling time of the deviated wells
and eventually reducing the drilling cost for future wells.

Keywords: drilling; rate of penetration; high-angled wells; feature ranking; artificial neural networks
ed

1. Introduction
pt

In the past 10 years, the petroleum industry has faced significantly higher operating costs associated with drilling. These
expensive operational costs prompted a renewed emphasis on optimizing drilling performance [1]. The rate of penetration
ce

(ROP) is one of the main factors that can be varied to improve drilling performance [2]. Penetration rate or drilling speed is
defined as the depth of penetration achieved per unit of time [3]. In most cases, considerable drilling cost reduction is
Ac

accomplished by maximizing the ROP. This is chiefly accomplished by finding the optimum drilling parameters [4].
However, applying very fast ROP results in improper hole cleaning. In fact, one of the predictable problems may occur if the
drilling cuttings are not removed from the bottom hole properly [5]. Therefore, it is essential to optimize a relationship
between the ROP and drilling parameters to obtain superior drilling performance with less cost, while maintaining safety.
The ROP depends on so many variables that one cannot be changed without affecting the others [6]. Generally, these
variables can be classified into three categories: rig- and bit-related parameters, one-dimension mechanical earth model (1D
MEM) parameters, and drilling fluid properties [7]. These categories are divided into two major subgroups: controllable
parameters and uncontrollable or environmental parameters. The controllable variables, known as the operational parameters,
refer to those parameters that can be modified quickly to adjust the ROP without affecting the operations economics, such as
the weight on bit (WOB), revolutions per minute (RPM), flow rate (FR), total flow area (TFA), and bit type (IADC).
However, it is difficult to change the uncontrollable parameters because of economical or geological reasons, such as mud

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
weight (MW), bit diameter, wellbore trajectory, pore pressure gradient, unconfined compressive strength (UCS), and three
principal stresses [8, 9].
In the literature, traditional mathematical models have been used with some success to predict the ROP in drilling [10,
11]. These models employ a multiple regression analysis of detailed drilling datasets taken over short intervals [12]. Data
included in the analysis were the effects of auxiliary data (i.e., formation strength, bit properties, mud properties, bit design,
bit diameter, WOB, RPM, bit wear, bit hydraulics, etc.) [13, 14]. Due to the high complexity and nonlinear behavior of
drilling parameters in relation to the ROP, these mathematical models cannot accurately and comprehensively predict the
ROP [15]. Furthermore, none of these models includes the influence of the wellbore trajectory and mechanical earth model
parameters on the ROP values. Thus, application of these models for ROP prediction cannot reveal the actual downhole
conditions.
Because of these difficulties and the restriction of the mathematical models that were developed earlier, machine learning

d
(ML) seems to be an attractive alternative to model this complicated physical process [16]. A key characteristic regarding

te
ML is its ability to recognize complex patterns with good predictive accuracy through a typical learning process [17]. Thus,
ML can provide intuitive solutions to complicated problems with no need for the formal description of the underlying physics

di
[18, 19]. In addition, the ROP predictions by these methods are not bound to bit specifications, borehole conditions, or
bottom hole assembly (BHA). Among the ML methods, artificial neural networks (ANNs) can most powerfully and

e
efficiently produce accurate ROP predictions [20-22]. Usually, ANNs are utilized to identify complex relationships when
sufficient data exist. The increased availability of drilling parameter measuring tools has resulted in datasets with many

py
recorded variables related to the drilling process. A number of studies have succeeded in modeling many of the oil and gas
industry processes using ANNs [23-26]. In general, their conclusions find ANNs to be a good modeling method due to their

Co
learning, generalization, and nonlinear behavior properties.
This research seeks to assess the potential of using ANNs to predict the drilling performance in terms of the ROP using
datasets from previously drilled wells.

ot
2. Methodology tN
2.1. Data acquisition
Database preparation is a very challenging and time-consuming process. The ANNs method can learn the trend or
behavior of previous experiences with certain experiments or data-set. Thus, the quality and authenticity of the datasets
rip

(input) play a significant role in the reliability of the prediction process and the decisions made by the ANNs technique. For
these reasons, the learning datasets should be large enough and also of high enough quality to cause the model to refine itself.
Otherwise, lots of noise might be created, which negatively affects the model [27].
sc

2.1.1. Rig- and bit-related parameters


nu

Due to drilling requirements and the similarity of wells located close to one another, collecting past data and utilizing the
data in a useful manner have an important impact on drilling cost reduction. As a matter of fact, there are various factors that
can affect ROP. Previous studies showed that the ROP is largely dependent on the rig- and bit-related parameters such as
Ma

measured depth (MD), rotation per minute (RPM), torque (TQ), weight on bit (WOB), flow rate (FR), circulating pressure
(CP), bit type (IADC), bit working-hours (BW), total flow area (TFA), and wellbore trajectory (azimuth [AZI] and
inclination [INC]).
In this paper, a total of 13,125 cases (datasets) were collected from MEM data, geophysical logs, daily drilling reports
ed

(DDRs), daily mud reports (DMRs), final well reports (FWRs), and mud logs (master logs) of 14 deviated wells drilled in the
Zubair Formation located in Southern Iraq. All of the studied wells were highly deviated (i.e., inclinations higher than 45°),
where water-based mud and conventional bottom hole assembly (BHA) were used to drill these wells. Moreover, the size of
pt

the hole section was 8 1/2”, which implies that this was a production section. Thus, the mud type and hole diameter were not
considered as input parameters. It is important to note that in this study, a drag bit with a polycrystalline diamond compact
ce

(PDC) was considered.


Ac

2.1.2. Drilling fluid properties


Drilling mud is a complex fluid designed to serve several purposes: (1) maintaining the integrity of the formation, (2)
cooling and lubricating the drill bit, and (3) removing the rock particles (cuttings) from the bit face [28]. Therefore, mud
weight (MW), marsh funnel viscosity (MFV), plastic viscosity (PV), yield point (YP), 10-second gel strength (Gel 10”), 10-
minute gel strength (Gel 10’), fluid loss, and solids content have been confirmed as significant factors in shaping the ROP
[29]. Basically, the ROP decreases with increasing MW, mud viscosity, and solids content [30]. Unfortunately, these
parameters cannot be instantaneously measured in real time, at least not with today’s technologies. These parameters are
measured routinely based on drilling operator policy. This can be a challenge as only some data relative to drilling fluid
properties can be obtained from the daily mud report. To overcome the limitations of the data quantity, the drilling fluid

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
properties were measured manually on the meter readings. Therefore, the number of data points will be the same as the
number of the rig- and bit-related parameters.

2.1.3. One-Dimension Mechanical earth model (1D MEM) parameters


Intrinsically, each formation has some specific characteristics and some properties that have an effect on the ROP. Thus,
the lithology type, rock mechanical properties, pore pressure, vertical stress (σv), horizontal minimum stress (σh), and
horizontal maximum stress (σH) were identified as formation characteristics [31].
An integrated workflow was applied to construct a 1D MEM. The first step in building the model was to collect a proper
set of data for the Zubair Formation. The needed data were gathered from various sources from the offset wells. The second
step was to assess the logging input data. A data audit was performed to make sure that all data, which were acquired from
the laboratory and field, were complete, accurate, and reasonable. In general, almost all of the data gathered for this study

d
were complete and of good quality. Moreover, the necessary data, including gamma-ray logs, density logs, porosity logs,

te
sonic logs (compression and shear wave velocities), resistivity logs, formation micro-imager (FMI) logs, caliper logs, mud
logs (master logs), mini-frac tests, and laboratory measurements, were available for most of the selected wells. Subsequently,

di
the data were used to build a basic 1D MEM [32].

e
2.1.3.1. Rock mechanical parameters

py
Rock mechanical properties consist mainly of elastic parameters and strength parameters. These properties are considered
primarily to determine rock strength and formation drillability. The continuous profile of rock mechanical properties provides
a good indication of the natural variation in the formation hardness in different layers within the interval of interest.

Co
Laboratory tests are considered the most direct and reliable way of determining rock mechanical properties. Therefore, both
triaxial and triaxial-multistage testing techniques were carried out on retrieved core samples from the Zubair Formation to
measure the rock mechanical properties. Valid local empirical correlations were established between laboratory-derived rock
mechanical properties and geophysical well logs, such as porosity (∅) and compressional acoustic wave velocity (vp). Further

ot
details of estimating the rock mechanical properties of Zubair sandstone and shale formations using wireline measurements
were comprehensively discussed in the studies conducted by Abbas et al. [33, 34].
tN
Rock strength parameters such as unconfined compressive strength (UCS), and internal friction angle (φ) indicate to the
ability of the rock formation to withstand the in-situ stress environment around the wellbore. The UCS and φ were
determined using consolidated drained (CD) multistage triaxial tests for sandstone and consolidated undrained (CU) triaxial
rip

tests for shale. In this study, the UCS and φ were determined using Eqs. 1, and 2 for the sandstone formation as a function of
porosity (∅), while Eqs. 3 and 4 were used to estimate these parameters for the shale formation as a function of the
compressional wave velocity (vp) [33, 34].
sc

UCS= sand 133.2 − 370.82φ (1)


ϕ sand 64.369 − 99.238φ
= (2)
nu

UCS shale = 2.6477e


0.6006 v p
(3)

ϕ shale = 17.134e
0.239 v p
Ma

(4)
The elastic parameters (such as Young's modulus (E) and Poisson's ratio (v)) demonstrate the deformation behavior for
isotropic elastic materials. The static values of these parameters were calculated using porosity (∅) for the sandstone
formation and compressional acoustic wave velocity (vp) for the shale formation, as follows [33, 34]:
ed

= Esand 40.476 − 136.79φ (5)


vsand 0.1203 − 0.766φ
= (6)
pt

Eshale = 0.2966e
0.6984 v p
(7)
ce

−0.353v p
vshale = 0.7621e (8)
The mechanical responses, as well as the properties of both the sandstone and shale formations, were significantly
Ac

different. Therefore, by classifying rocks according to their mechanical stratigraphy, it became possible to use different
correlations for different formations to best estimate their rock mechanical properties. The differentiation of non-shale from
shale was realized by applying a threshold of 75 gAPI to the gamma-ray logs in the studied wells [35].

2.1.3.2. Formation pore pressure


Formation pore pressure is a key constituent in reservoir geomechanical modeling. It is one of the critical 1D MEM
parameters widely used to estimate in situ horizontal principal stress magnitudes. Two methods (i.e., direct and indirect) are
often employed in the petroleum industry to determine pore pressure. Direct measurement methods use well test techniques,
such as the drill stem test (DST) and repeated formation test (RFT), to measure formation pore pressure for specific depths,
3

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
whereas indirect (i.e., empirical and theoretical) methods are based on petrophysical data that are developed to predict pore
pressure along the well length [36]. For more confidence, the estimated formation pore pressure from indirect methods is
usually validated with the available measured formation pressure points of the DST or RFT. Eaton’s (1969) equation is
conventionally applied to predict the formation pore pressure based on the sonic wireline measurements [37]. This equation is
formulated as
3
 NCT 
Ppg =OBG − (OBG − Phg )   (9)
 DT 
where Ppg is the pore pressure gradient, OBG is the overburden gradient, Phg is the hydrostatic pressure gradient (also known
as the normal pore pressure), NCT is the normal compacted trend line fitting compressional wave log measurements, and DT
is the P-wave transit time.

d
te
2.1.3.3. Vertical stress
The overburden weight per unit area is called overburden stress. The overburden or vertical stress is the most basic input

di
parameter in the 1D MEM analysis. In areas with low tectonic activity, vertical stress represents the weight of overlying
formations [38]. The vertical stress in the Zubair Formation was calculated by integrating the bulk density log over the

e
vertical depth, using Eq. 10.

py
z
σ v = ∫ ρ ( z ) gdz (10)
0

Co
where g represents the acceleration constant due to gravity (m/s2), z is the vertical depth (m), and ρ is the rock bulk density
(g/cm3).

2.1.3.4. Horizontal stresses (minimum and maximum)

ot
The determination of minimum and maximum horizontal stress magnitude involves technical challenges in
geomechanical modeling. Since 1930 s, many researchers, geologists, geophysicists and engineers attempted to determine the
tN
magnitude of horizontal in-situ stresses. Hence, various laboratory and field methods are introduced for the estimation of the
horizontal in-situ stress magnitude as hydraulic, relief, jacking, strain recovery, and focal mechanism [39]. The poro-elastic
relations are the most commonly used method for estimation of horizontal in-situ stresses along the well length [40].
rip

Assuming flat-layered, poro-elastici deformation in the formation rock, a pair of particular constant strains, ɛy and ɛx, is
applied to the formation in the direction of the maximum and minimum stress, respectively. For a fluid-saturated porous
material that is assumed to be linear, elastic, and isotropic, considering anisotropic tectonic strain, the horizontal stresses
sc

(minimum and maximum) are expressed in Eqs. 11 and 12, respectively [41]. This approach estimates the magnitudes of the
horizontal principal stresses along the well length using the static Young's modulus, static Poisson's ratio, regional pore
nu

pressure, rock deformation, and overburden stress.


ν 1 − 2ν E νE
σh = σν + α Pp + ε + εy (11)
1 −ν 1 −ν 1 −ν 2 x
1 −ν 2
Ma

ν 1 − 2ν E νE
σH = σν + α Pp + ε + εx (12)
1 −ν 1 −ν 1 −ν 2 y
1 −ν 2
ed

In the above equations, α is Biot’s coefficient, which is maintained at unity to account for the brittle failure of rocks
(conventionally α = 1), E is the static Young’s modulus (GPa), ɛx is the strain in the minimum horizontal stress direction, and
pt

ɛy is the strain in the maximum horizontal stress direction. The two horizontal strains (ɛy and ɛx) can be measured by Eqs. 13
and 14, respectively [42].
ce

σ vv 
v2 
εy
=  1 −  (13)
E  1− v 
Ac

σ vv  1 
=εx  − 1 (14)
E  1− v 
The magnitude of the minimum horizontal stress obtained from the above formulae can be calibrated against the direct
measurement procedures based on the field-tests (i.e. extended leak off test (ELOT), a standard leak off test (LOT), or a mini-
frac test) [43].

2.2. Input data selection


4

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
Understanding the influence of the input parameters is considered the primary concern when developing ANN models.
Introducing more input parameters than required will result in large network size and consequently decrease learning speed
and efficiency. Since the drilling process has many effective parameters, it is essential to find the best set of variables that are
related to ROP.
Feature ranking is a technique which finds the importance of features and can be used to reduce the dimensions of the
dataset. Using this technique also has the benefits of shorter training time, ease of interpretation of models, overfitting
reduction and lower cost in data collection [44]. Based on the choice of predictive models for feature ranking, different
variables will be viewed as essential [45]. Therefore, researchers may disagree on which is the appropriate model for feature
ranking.
In this work, fscaret package as a in R environment is used to reduce the input vector and give stable variable importance
[46]. This package provides fast and automated feature selection based on caret package modeling methods. The aim of these

d
method are to simplifying training and perform variety of predictive modeling technics. The main advantage of this extension

te
is that it requires minimum user involvement. Also, the variety of the methods used in combination with the scaling
according to RMSE or MSE obtained from models profit the user. The idea is based on the assumption that the variety of

di
models will balance the roughness of calculations (default model settings were applied). This method is a simple package
providing fast and automated feature ranking based on the caret package [47]. It uses a large number of 103 various models

e
in order to determine the importance of features and rank them. In the end, the importance of variables is averaged through all
models to give reliable results and is scaled from 0 to 100.

py
In this work, the following parameters were collected as input: drilling operation parameters, formation type and lithology
of the rock, and fluid mud characteristics. All these input parameters are considered the independent variables to estimate the

Co
output i.e. ROP. Out of the 25 variables studied, six constant or nearly constant variables had to be excluded, and the
remaining 19 variables were considered to be the input parameters to predict the ROP, as shown in Table 1. The ranges for
these variables are very significant factors, and if new input data that fall outside of these ranges were applied to machine
learning models, the results of the ROP predictions would be not reliable [48].

ot
Table 1—Ranges of input parameters
tN
Parameter Minimum Maximum
Measure drilling depth (MD) (m) 2213 3834
Weight on bit (ton) 3.5 20
rip

Bit rotational speed (rev/min) 40 200


Bit type (IADC) 211 323
Bit working hours (hr) 0.01 155
Torque (klb.ft) 0.5 23
sc

Flow rate (l/min) 500 2500


Circulating pressure (psi) 366 2766
Total flow area (in2) 0.6 1.19
nu

Azimuth (degree) 22 338


Inclination (degree) 46 78
Mud weight (g/cc) 1.04 1.3
Marsh funnel viscosity (sec.) 40 77
Ma

Plastic viscosity (cp) 9 35


Yield point (g/100 cm2) 10 51
Unconfined compressive strength (MPa) 3 149
Vertical stress (MPa) 30 78
Maximum horizontal stress (MPa) 24 94
ed

Minimum horizontal stress (MPa) 22 78

2.3. Artificial neural networks (ANNs)


pt

ANNs are information processing systems that are a rough approximation and simplified simulation of a biological
ce

learning process and have performance characteristics similar to those of biological neural networks. These are adaptive,
parallel information processing systems, which are able to develop associations, transformations or mappings between
objects or data and have proven to have potential in solving problems that require pattern recognition [49]. The processing
Ac

elements of ANNs are artificial neurons. These neurons consist of four basic components that include input data, connection
strengths (weights), a transfer function (activation function), and output values. The inputs to each neuron are multiplied by
the adjusted weight elements (w1j ... wnj), and these modified inputs are summed up with a value called “bias.” Then, the
weighted and summed output is fed through a nonlinear function (i.e., a transfer function) to produce the results. Usually, the
neurons are interconnected in a multilayer network topology that consists of three layers: (1) an input layer, (2) one or more
hidden layers, and (3) an output layer (Fig. 1). The hidden layer(s) are the coefficients that provide the relationship between
the input and output layers.

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME

Fig. 1– An ANNs structure with one hidden layer.

d
te
The most common types of ANNs are feed-forward networks, which are the most efficient ones [50]. In such ANNs
architecture, the information will propagate in one direction from input neurons through the transfer function of the hidden

di
neurons to the outputs. For supervised learning, a back-propagation algorithm with a Levenberg-Marquardt training function
is probably the best-known learning algorithm for neural networks [51]. Back-propagation refers to the mechanism adjusting
and modifying the connection weights and biases of the network. In the beginning, the training algorithm is created by

e
assigning random weights and biases. In the first stage of this cycle, the input data are transformed through each layer until

py
an output is calculated. An error signal is computed for each output unit by comparing the generated output with the desired
output (target). These error signals are propagated back from the output layer to each neuron in the hidden layer, which, in
general, contributes directly to the output. Based on the relative contribution of each unit in the hidden layer to create the

Co
original output, these units will receive only a portion of the total error signal. The neurons use transfer functions to generate
their output from the net input. The most commonly used transfer functions for back-propagation are PURELIN, TANSIG,
and LOGSIG. The number of neurons in the input layer corresponds to the number of parameters being fed to the network as

ot
input. The same is true for the target layer. Therefore, the model was built using ANNs with 19 neurons in the input layer and
only one neuron in the output layer. The optimum number of neurons and layers was selected based on trial and error. On the
tN
other hand, many hidden layers of various numbers of neurons have been tested with the TANSIG and LOGSIG transfer
functions, while the output layer was one neuron, using the PURELIN transfer function.
In the modeling process, the database is randomly divided into two parts: a training dataset is applied to develop and
adjust the weights in a network, and a testing dataset is applied to examine the final performance of the ANNs. Among
rip

13,125 datasets of acquired data, which contained 19 input parameters and one output parameter (ROP), the ratio of 4:1was
applied for training and testing the developed ANNs model, respectively.
sc

2.3.1. Data preprocessing


Before supplying the available data to the neural network, it is crucial to pre-process the data. Data pre-processing helps
nu

to speed up the learning process and ensures that every parameter receives equal attention by the network and improving the
overall network performance. In addition, the ANNs model cannot be trained with textual or symbolic data. In this case, the
bit type must be translated into a numeric form. There are many techniques to translate textual or nonnumeric information
Ma

into numeric data. Some of the common textual translation methods consist of numbering classes, unary encoding, and binary
encoding. Numbering classes are used in this study to translate the bit type into a numeric form. The three-digit numbers of
the International Association of Drilling Contractors (IADC) were set as an equivalent for each drilling bit type. Furthermore,
these input parameters fall within various ranges; for example, the measured depths are described in thousands of meters,
ed

while borehole inclinations are described in single-digit values or decimals. Therefore, the next step is to preprocess the
inputs and target data, such as normalization, into a form that can be used with the ANNs.
pt

The data are normalized depending on either the applied-transfer function in designing the ANNs [52]. A transfer
function used in training can be either unipolar, with values between 0 and 1, or bipolar, with values between −1 and +1.
Regarding the FFBP with TANSIG hidden layers, the normalization can be applied to fall input and target data in the range
ce

of −1 to +1 (Eq. 15). On the other hand, when LOGSIG transfer function is used, the input and target data are scaled in the
range of 0 to 1 (Eq. 16).
Ac

2( X − X Min )
=Xi −1 (15)
( X Max − X min )
( X i − X min )
Xi = (16)
( X max − X min )

where Xi is the normalized value, Xmin is the minimum of original values, Xmax is the maximum of original values, and X is the
original value.
6

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
2.3.2. Model performance assessment criteria
The criteria implemented for assessing performance are the three most commonly used in petroleum engineering journals
[53]. This is done in order to align with the best practice and to carry out a fair assessment of the developed model. The
performances of neural network models are assessed by the root mean square error (RMSE), Average Absolute Percentage
Error (AAPE), and correction coefficient (𝑅𝑅2).
The RMSE is conventionally used as an error function for quality monitoring of the model. Model performance increases
as RMSE decreases. RMSE can be calculated by the following equation:

1 n
=
RMSE ∑
n i =1
( f ( xi ) − yi ) 2 (17)

d
te
Absolute error expresses the difference (error) between each of the predicted output and its corresponding actual value.
Relative error is calculated as the ratio of the mean of the absolute error and mean value of the predicted output. Percent

di
relative error is the relative error measured in percentage. Hence, the formula is as expressed in Eq. 18.

e
1 n xi − yi
= ∑ ] ∗100

py
AAPE [ (18)
n i =1 xi

Co
𝑅𝑅2 measures the statistical correlation between the predicted and actual values. The value ranges between 0 and 1. The value
of 0 indicates no similarity and a bad performance while 1 represents a perfect correlation between the model output and
actual value to be predicted indicates perfect generalization accuracy. The formula for correlation coefficient depicted in (19).

ot
n

∑ ( f (x ) − y )
i i
2

R2 = 1 − i =1
tN
n n
(19)
i
=i 1 =i 1
∑ f (x ) − ∑ ( y )
2
i
2
/n
rip

In the above equations, yi is the actual data; f(xi) represents the predicted data; xi and xn are the input parameters; and n is
the total number of records. The best-developed model will have a lower RMSE and AAPE but a higher R2.
sc

3. Results and Discussion


3.1. Construction of the 1D MEM parameters
nu

The aforementioned empirical correlations were used to derive the rock mechanical properties from the neutron and sonic
logs. The suggested mechanical stratigraphy was used to apply correlations of shale to the shale formation and relationships
Ma

of sandstone to the sandstone formation. Then, each derived mechanical properties was combined to build a single log from
the top to the bottom of the Zubair Formation. The laboratory-measured rock mechanical properties data were added to the
plot to assess the level of agreement between the experimentally measured mechanical properties and the derived mechanical
properties logs [54, 55]. Figure 2 presents examples of the rock mechanical property logs for one of the cored wells in the
ed

Zubair Formation. As can be seen, there is a high degree of positive correlation between the calculated rock mechanical
property logs and those derived from the laboratory.
pt
ce
Ac

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME

d
te
e di
py
Co
Fig. 2– Predicted rock mechanical properties logs and laboratory measurements.

In the current study, the pore pressure of the depleted sandstone formations was computed using constant pressure
gradients that were determined from repeated formation test (RFT) measurements. Due to the discontinuity of the RFT data

ot
for the shale formations, the pore pressure was estimated using the modified Eaton (1969) technique of Eq. 3. The resultant
formation pore pressure profile was calibrated against the actual pore pressure measurements.
tN
The vertical stress magnitude on the Zubair Formation was calculated using the density log, according to Eq. 10. For
shallow zones, where the density log was not recorded because it was not of geologic interest, a compaction relationship
based on the regional density data was used to create a synthetic density curve to fill the gaps where the direct measurement
rip

was unavailable. Moreover, the continuous estimation of the minimum and maximum horizontal stress magnitudes along the
well length was obtained from poro-elastic formulations using Eqs. 11 and 12, respectively. The minimum horizontal stress
was calibrated against the direct measurements of closure pressures that were available from the mini-frac test. The predicted
sc

minimum horizontal stress from the poro-elastic formulation showed reasonably good agreement with the closure pressures.
Fig. 3 illustrates the estimate of the vertical stress, maximum horizontal stress, minimum horizontal stress, and pore pressure
magnitudes.
nu
Ma
ed
pt
ce
Ac

Fig. 3– Estimation of the pore pressure and in situ principal stress magnitudes at a single well location.
8

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME

3.2. Feature ranking


The initial collected parameters contained 25 independent variables. To reduce the dimensions of the dataset and to find
the ranking of each variable, the fscaret package in R environment was performed. Figure 4 shows the results of variable
ranking and importance achieved by using fscaret. The final selected input parameters consisted of 19 variables that are most
consistent with the specifications given in the literature.
The results demonstrate that the WOB, RPM, and FR were the first, second, and third major factors affecting the ROP for
this dataset, respectively. These parameters refer to the drilling mechanical parameters that have been investigated in some of
the literature as the most effective parameters influencing the ROP [56]. The UCS was the next most significant parameter,
which is related to drilling resistance and rock hardness. It should be mentioned that the result for variable importance is case
specific, and for other reservoirs and formations, it may yield different results. For example, the UCS, INC, AZI, σv, σH, and

d
σh may have a negligible impact on the ROP for other formation types.

te
e di
py
Co
ot
tN
Fig. 4– Ranking of variables.
3.3. Network architecture design
rip

The TANSIG and LOGSIG transfer functions were examined for one, two, and three hidden layers. In addition, the
optimal number of neurons in hidden layers was determined by increasing the numbers incrementally. For the TANSIG
transfer function, the RMSE reached its lowest value at 1.37 with 30 nodes and three hidden layers, while the RMSE attained
sc

its lowest point at 2.83 with 32 nodes and two hidden layers for the LOGSIG transfer function. Comparing the RMSE
performance indicated that the architecture of 30 neurons in three hidden layers with the TANSIG transfer function yielded
nu

the best efficiency in predicting the ROP.


After selecting the optimum structure of the FFBP, a comprehensive statistical and graphical error analysis were
conducted for assessment of the ANNs performance. Figs. 5a and 5b present cross-plots of the predicted versus the actual
Ma

ROP for the developed model for the training and testing datasets, respectively. The ANNs model achieved a high R2 of 0.92
for the training dataset, while R2 was found to be 0.97 for the testing dataset. In addition, error distribution plots were also
used to allow more statistical analysis of ANNs performance (Figs. 6a and 6b). Mean and standard deviation (SD) were in
turn equal to 0.26 and 1.03 for training dataset, while the mean and standard deviation (SD) were equal of -0.16 and 0.72 for
ed

the testing dataset. Small values of mean and standard deviation (SD) verify the robustness of ANNs model. The error
distribution indicates 93% of predicated ROP values have errors in the range of ±2 at the training phase, while 99% of
predicated ROP values have errors in the range of ±2 m/hr at the testing phase which is an acceptable error for ROP. An
pt

inspection of Figs. 5 and 6 clearly show that the performances of the ANNs model for the training and testing datasets varied
slightly. Furthermore, the developed ANNs model provided very good results for the two datasets.
ce
Ac

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
(a) 20 (b) 20

R 2 =0.92 R 2 =0.97
15 15

Real ROP (m/hr)


Real ROP (m/hr)

10 10

5 5

0 0

d
0 5 10 15 20 0 5 10 15 20
Predicted ROP (m/hr) Predicted ROP (m/hr)

te
Fig. 5– Model outputs vs. real data: (a) training dataset, (b) testing dataset.

di
(a) (b)

e
py
Co
ot
tN
rip

Fig. 6– Error distribution statistics for the developed ANNs model: (a) training dataset, (b) testing dataset.
sc

To demonstrate the robustness of the proposed ANNs model, the RMSE, AAPE, and R2 were calculated for the training
and testing datasets. As seen in Table 2, the predicted performance for the testing dataset is better than for the training
nu

dataset. The high performance of the ANNs model in the testing phase can be considered a good indication of the model’s
capabilities.
Ma

Table 2—Prediction performance of the proposed ANNs models


Dataset Performance Indicator ANNs
RMSE 1.37
Training AAPE 9.1%
ed

R2 0.92
RMSE 1.23
Testing AAPE 8.8%
R2 0.97
pt

3.4. Effect of operating parameters on the ROP


ce

The impact of the most effective controllable parameters (i.e., WOB, RPM, and FR) was investigated to determine their
effect on the ROP. Figs. 8a, 8b, and 8c show the effect of these variables on the ROP; for each of these variables, the other
variables were kept constant at their midrange or recommended values (Table 3), while the parameters under study (i.e.,
Ac

WOB, RPM, and FR) varied within the range of the input parameters used in the ANNs training process.

Table 3—Constant values of input parameters

Parameter Fixed Value


Measure drilling depth (MD) (m) 3350
Weight on bit (ton)* 12
Bit rotational speed (rev/min)* 130
Flow rate (l/min)* 2500
Bit type (IADC) M322
Bit working hours (hr) 10

10

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
Torque (klb.ft) 8
Circulating pressure (psi) 2000
Total flow area (in2) 0.663
Azimuth (degree) 140
Inclination (degree) 60
Mud weight (g/cc) 1.2
Marsh funnel viscosity (sec.) 54
Plastic viscosity (cp) 16
Yield point (g/100 cm2) 20
Unconfined compressive strength (MPa) 70
Vertical stress (MPa) 60
Maximum horizontal stress (MPa) 50
Minimum horizontal stress (MPa) 45

d
*Parameter under study is varied.

te
Figure 7a demonstrates a semi-straight line response of the ROP to the WOB until a critical point (i.e., founder point).
The founder point is the point at which the ROP is no longer a linear response to the applied WOB. This point depends on the

di
bit design and the structure of the bit teeth or cutters [57]. Increasing the WOB will push the bit cutters further down into a
formation and disintegrate more rocks. However, if too much WOB is applied (i.e., the founder point is reached), rapid bit-

e
wearing will occur, resulting in a lower ROP. In Fig. 7a, the WOB was increased in 0.5-ton steps, and the ROP responded by

py
increasing 0.47 m/hr with each step, up to 13.5 tons. The bit founder point was between 12 and 13.5 tons, and the next
increase in the ROP was at less than 0.45 m/hr (0.12 m/hr). To achieve a high ROP with a long bit life, the drilling team must
limit the WOB so that it remains at or below the founder point (13.5 tons).

Co
The relationship between the ROP and RPM is presented in Fig. 7b. An increase in the RPM generally enhanced the
ROP, but slight growth in the ROP was seen at a high RPM of more than 130 rev/min. When starting at a low RPM, an
increase in the RPM could raise the shearing force of the bit cutters to disintegrate more rocks and therefore increase the
ROP. This improvement continued until hole-cleaning problems occurred and reversed the result. When the RPM range was

ot
limited, a later reverse of the performance was not observed (Fig. 7b).
Figure 7c shows that the ROP is directly proportional to the FR, where increasing the FR improves the lifting capacity of
tN
the drilling fluids to transport the cuttings, which consequently raises the ROP. However, an excess amount of FR may cause
a positive effect on the hydraulic impact force by increasing the equivalent circulating density (ECD), which decreases the
ROP. This behavior was not observed because the FR range was limited to 2200 l/min (Fig. 7c).
rip

(a) (b)
sc
nu
Ma
ed

(c)
pt
ce
Ac

Fig. 7– Effect of operating parameters on the ROP: (a) effect of the WOB on the ROP, (b) effect of the RPM on the ROP, (c) effect
of the FR on the ROP.

11

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME

3.5. Predicting the ROP for a number of high-angled wells


A simulation was conducted using the collected datasets (i.e., MD, RPM, TQ, WOB, FR, CP, IADC, BW, TFA, AZI,
INC, MW, MFV, PV, YP, UCS, σv, σH, and σh) to predict the ROP for high-angled wells (i.e., wells numbered 1, 2, and 3)
(Figs. 8, 9, and 10, respectively). The validity of these parameters was first examined to remove outlier data. Then, the
compatibility of these parameters to the developed ANNs was checked by matching them with the range of the dataset used
for the training network. If the new dataset fell outside the range, the model output would not be reliable.
The predicted ROP was compared to the actual ROP and introduced as a function of the measured depth (MD). Residual
error analyses were conducted to visualize the quality of the predicted values compared to the actual values. The results show
a reasonable match between the ROPs, which indicates the success of modeling with ANNs and of the entire procedure
undertaken by this study. Furthermore, the datasets that were used to develop the ANNs were nosier than the wells. This is

d
advantageous for prediction because the ANNs model underwent many scenarios to better learn about the ROP behavior in
the Zubair Formation.

te
Thus, the developed model was successful at indicating changes in the ROP due to a change in input parameters. For
example, after a sudden decrease in the WOB and RPM occurred, the actual ROP experienced a sudden significant decrease,

di
as shown in well 1 in the short interval between 3012 m and 3026 m (Fig. 8). This was detected and reflected on the value of
the predicted ROP at the mentioned depth. In addition, the sudden increase in the ROP that was associated with an increase in

e
the WOB and RPM was identified by the predicted ROP over several short intervals in well 2 from 3460 m to 3466 m and

py
also from 3585 m to 3590 m and for well 3, between 3224 m and 3229 m (Fig. 9 and Fig. 10, respectively). However, no
unjustified changes in the actual ROP without any clear known changes in the input parameters was detected by the
developed model; therefore, none was reflected in the predicted ROP. Changes in the predicted ROP were smooth, an

Co
illustrative example of which can be found in well 3, at a depth 3278 m to 3300 m and 3453 m to 3500 m, where a sudden
change in the actual ROP was indicated, although there were no changes in the input parameters (Fig. 10).
Measured ROP

ot
(a) Predicted ROP
(b) Residual Errors
3000 3000
tN
3050 3050
rip
sc

3100 3100
nu
Depth(M)
Depth(M)

3150 3150
Ma

3200 3200
ed

3250 3250
pt
ce

3300 3300
0 5 10 15 20 -10 -5 0 5 10
ROP (m/hr) Residual Error of ROP
Ac

Fig. 8– ROP prediction for well 1: (a) predicted and measured ROP along the depth, (b) residual errors of the predicted ROP.

12

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
Measured ROP
(a) Predicted ROP
(b) Residual Errors
3350 3350

3400 3400

3450 3450

d
Depth(M)
Depth(M)

te
3500 3500

di
3550 3550

e
py
3600 3600

Co
3650 3650
0 5 10 15 20 -10 -5 0 5 10

ot
ROP (m/hr) Residual Error of ROP
tN
Fig. 9– ROP prediction for well 2: (a) the predicted and measured ROP along the depth, (b) residual errors of the predicted ROP.

Measured ROP
(a) Predicted ROP
(b) Residual Errors
rip

3200 3200
sc

3250 3250
nu

3300 3300
Ma
Depth(M)

Depth(M)

3350 3350
ed

3400 3400
pt
ce

3450 3450
Ac

3500 3500
0 5 10 15 20 -10 -5 0 5 10
ROP (m/hr) Residual Error of ROP

Fig. 10– ROP prediction for well 3: (a) the predicted and measured ROP along the depth, (b) residual errors of the predicted
ROP.

13

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME

4. Conclusions
Introducing an artificial neural network to petroleum industry eases the process of recognizing complex patterns and
relations between different parameters without any assumptions about the nature and distribution of the data. Presented work
is an example of the implementation of the artificial neural network to model the rate of penetration using different
parameters such as rig- and bit-related parameters, 1D MEM parameters, and drilling fluid properties. Moreover, following
this technique can help to unfold knowledge from the dataset.
Applying fscaret package in R environment enabled to find the importance and ranking of all 19 variables. The results of
feature ranking demonstrate that the WOB, RPM, and FR had the most impact on the ROP. According to the performance
indicators, the structure of the FFBP, with a five-layer network, with an architecture of {19 30 30 30 1} neurons provided the
best performance for ROP prediction. Among several tested transfer functions, TANSIG for hidden layers and PURELIN for

d
the output layer yielded better efficiency in predicting the ROP.

te
Finally, the newly developed model can only be used within the range of applied datasets, and caution should be taken
when selecting the input parameters. However, for a new system with variables beyond the range of the input parameters

di
used in the training operations of this study or for a new bit type/size (e.g., roller cone bits), it is necessary to develop new
ANNs models. Also, this model is valid for other datasets that fall within the range of the dataset employed in the training

e
process. Therefore, the use of a wider range of input parameters for the training process can enable the ANNs to have wider
applications, in select cases.

py
Acknowledgments

Co
The authors would like to gratefully acknowledge Basrah Oil Company (BOC) and Iraqi Drilling Company (IDC) in Iraq
for providing technical data and their permission to publish the results.

ot
Nomenclature
fscaret Automated Feature Selection from 'caret'
tN
ROP Rate of Penetration
1D MEM One-Dimension Mechanical Earth Model
MD Measured Depth
rip

WOB Weight on Bit


RPM Rotation per Minute
TQ Torque
sc

FR Flow Rate
CP Circulation Pressure
nu

BW Bit Working-hours
TFA Total Flow Area
AZI Azimuth
Ma

INC Inclination
UCS Unconfined Compressive Strength
σv Vertical Stress
σh Minimum Horizontal Stress
ed

σH Maximum Horizontal Stress


MW Mud Weight
MFV Marsh Funnel Viscosity
pt

PV Plastic Viscosity
YP Yield Point
ce

References
Ac

[1] Hegde, C., and Gray, K., 2018, “Evaluation of coupled machine learning models for drilling optimization,” J. Nat. Gas
Sci. Eng. 56, pp.397–407. http://dx.doi.org/10.1016/j.jngse.2018.06.006.
[2] Yi, P., Kumar, A., and Samuel, R., 2014, “Realtime Rate of Penetration Optimization Using the Shuffled Frog Leaping
Algorithm,” ASME J. Energy Resour. Technol., 137(3), p. 032902. https://doi.org/10.1115/1.4028696.
[3] Al-AbdulJabbar, A., Elkatatny, S., Mahmoud, M., Abdelgawad, K., and Al-Majed, A., 2018, “A Robust Rate of
Penetration Model for Carbonate Formation,” ASME J. Energy Resour. Technol., 141(4), p. 042903.
https://doi.org/10.1115/1.4041840.
[4] Gidh, Y. K., Ibrahim, H., and Purwanto, A.,2011, “Real-time drilling parameter optimization system increases ROP by
predicting/managing bit wear,” SPE Digital Energy Conference and Exhibition, Woodlands, Texas, April 19–21, Paper
14

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
No. SPE-142880-MS. https://doi.org/10.2118/142880-MS.
[5] Callejo, A., Arbatani, S., Kövecses, J., Kalantari, M., and Marchand, N. R., 2017, “Drill Bit Contact Dynamics Including
Side Cutting: Simulation and Validation,” ASME J. Energy Resour. Technol., 139(2), p. 022910.
https://doi.org/10.1115/1.4035514.
[6] Wang, P, Ni, H., and Wang, R., 2018, “A novel vibration drilling tool used for reducing friction and improve the
penetration rate of petroleum drilling,” J. Pet. Sci. Eng. 165, pp.436-443. http://dx.doi.org/10.1016/j.petrol.2018.02.053.
[7] Keshavarz Moraveji, M., and Naderi, M., 2016, “Drilling rate of penetration prediction and optimization using response
surface methodology and bat algorithm,” J. Nat. Gas Sci. Eng. 31, pp.829–841.
https://doi.org/10.1016/j.jngse.2016.03.057.
[8] Kahraman, S., Bilgin, N., and Feridunoglu, C., 2003, “Dominant rock properties affecting the penetration rate of
percussive drills,” Int. J. Rock Mech. Min. Sci. 40(5), pp.711–723. https://doi.org/10.1016/s1365-1609(03)00063-7.

d
[9] Ataei, M., KaKaie, R., Ghavidel, M., and Saeidi, O., 2015, “Drilling rate prediction of an open pit mine using the rock

te
mass drillability index,” Int. J. Rock Mech. Min. Sc. 73, pp.130-138. http://dx.doi.org/10.1016/j.ijrmms.2014.08.006.
[10] Deng, Y., Chen, M., Jin, Y., Zhang, Y., Zou, D., and Lu, Y., 2016, “Theoretical and experimental study on the

di
penetration rate for roller cone bits based on the rock dynamic strength and drilling parameters,” J. Nat. Gas Sci. Eng. 36,
pp.117–123. http://dx.doi.org/10.1016/j.jngse.2016.10.019.

e
[11] Hegde, C., Daigle, H., Millwater, H., and Gray, K., 2017, “Analysis of rate of penetration (ROP) prediction in drilling
using physics-based and data-driven models,” J. Pet. Sci. Eng. 159, pp.295–306.

py
https://doi.org/10.1016/j.petrol.2017.09.020.
[12] Rashidi, B., Hareland, G., and Wu, Z., 2015, “Performance, simulation and field application modeling of rollercone

Co
bits,” J. Pet. Sci. Eng.133, pp.507–517. http://dx.doi.org/10.1016/j.petrol.2015.06.003.
[13] Kahraman, S., Balcı, C., Yazıcı, S., and Bilgin, N., 2000, “Prediction of the penetration rate of rotary blast hole drills
using a new drillability index” Int. J. Rock Mech. Min. Sci. 37(5), pp.729–743. http://dx.doi.org/10.1016/s1365-
1609(00)00007-1.

ot
[14] Eren, T., and Kok, M. V., 2018, “A new drilling performance benchmarking: ROP indexing methodology,” J. Pet. Sci.
Eng. 163, pp.387–398. http://dx.doi.org/10.1016/j.petrol.2018.01.002.
tN
[15] Soares, C., Daigle, H., and Gray, K., 2016, “Evaluation of PDC bit ROP models and the effect of rock strength on model
coefficients,” J. Nat. Gas Sci. Eng. 34: pp.1225–1236. https://doi.org/10.1016/j.jngse.2016.08.012.
[16] Elkatatny, S., 2018, “Application of Artificial Intelligence Techniques to Estimate the Static Poisson's Ratio Based on
rip

Wireline Log Data,” ASME J. Energy Resour. Technol., 140(7), p. 072905. https://doi.org/10.1115/1.4039613.
[17] Paul, A., Bhowmik, S., Panua, R., and Debroy, D., 2018, “Artificial Neural Network-Based Prediction of Performances-
Exhaust Emissions of Diesohol Piloted Dual Fuel Diesel Engine Under Varying Compressed Natural Gas Flowrates,”
sc

ASME J. Energy Resour. Technol., 140(11), p. 112201. https://doi.org/10.1115/1.4040380.


[18] Alkamil, E. H., Abbas, A. K., Flori, R., Silva, L. E., Wunsch, D. C., and Chumkratoke, C.,2018, “Learning from
Experience: Real-Time H2S Monitoring System Using Fuzzy ART Unsupervised Learning,” IADC/SPE Asia Pacific
nu

Drilling Technology Conference and Exhibition, Bangkok, Thailand, August 27-29, Paper No. SPE-191097-MS.
http://dx.doi.org/10.2118/191097-ms.
[19] Alkamil, E. H., Al-Dabooni, S., Abbas, A. K., Flori, R., and Wunsch, D. C., 2018, “Learning From Experience: An
Ma

Automatic pH Neutralization System Using Hybrid Fuzzy System and Neural Network,” Procedia Computer Science,
140, pp.206-215. http://dx.doi.org/10.1016/j.procs.2018.10.330.
[20] Mantha, B., and Samuel, R., 2016, “ROP optimization using artificial intelligence techniques with statistical regression
ed

coupling,” SPE Annual Technical Conference and Exhibition, Dubai, UAE, September 26–28, Paper No. SPE-181382-
MS. http://dx.doi.org/10.2118/181382-ms.
[21] Abbas, A. K., Alameedy, U., Alsaba, M., and Rushdi, S., 2018, “Wellbore Trajectory Optimization Using Rate of
pt

Penetration and Wellbore Stability Analysis,” SPE International Heavy Oil Conference and Exhibition, Kuwait City,
Kuwait, December 10-12. SPE-193755-MS. http://dx.doi.org/10.2118/193755-ms.
ce

[22] Bhowmik, S., Panua, R., Debroy, D., and Paul, A., 2017, “Artificial Neural Network Prediction of Diesel Engine
Performance and Emission Fueled With Diesel–Kerosene–Ethanol Blends: A Fuzzy-Based Optimization,” ASME J.
Energy Resour. Technol., 139(4), p. 042201. http://dx.doi.org/10.1115/1.4035886.
Ac

[23] Rostami, H., and Khaksar Manshad, A., 2014, “A New Support Vector Machine and Artificial Neural Networks for
Prediction of Stuck Pipe in Drilling of Oil Fields,” ASME J. Energy Resour. Technol., 136(2), p. 024502.
http://dx.doi.org/10.1115/1.4026917.
[24] Wang, Y., and Salehi, S., 2015, “Application of Real-Time Field Data to Optimize Drilling Hydraulics Using Neural
Network Approach,” ASME J. Energy Resour. Technol., 137(6), p. 062903. http://dx.doi.org/10.1115/1.4030847.
[25] Manshad, A. K., Rostami, H., Rezaei, H., and Hosseini, S. M., 2015, “Application of Artificial Neural Network-Particle
Swarm Optimization Algorithm for Prediction of Asphaltene Precipitation during Gas Injection Process and Comparison
with Gaussian Process Algorithm,” ASME J. Energy Resour. Technol., 137(6), p. 062904.
http://dx.doi.org/10.1115/1.4031042.

15

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
[26] Ma, Z., Leung, J. Y., and Zanon, S., 2017, “Practical Data Mining and Artificial Neural Network Modeling for Steam-
Assisted Gravity Drainage Production Analysis,” ASME J. Energy Resour. Technol., 139(3), p. 032909.
http://dx.doi.org/10.1115/1.4035751.
[27] Bhowmik, S., Panua, R., Debroy, D., and Paul, A., 2017, “Artificial Neural Network Prediction of Diesel Engine
Performance and Emission Fueled With Diesel–Kerosene–Ethanol Blends: A Fuzzy-Based Optimization,” ASME J.
Energy Resour. Technol., 139(4), p. 042201. http://dx.doi.org/10.1115/1.4035886.
[28] Li, W., Zhao, X., Li, Y., Ji, Y., Peng, H., Liu, L., and Yang, Q., 2015, “Laboratory investigations on the effects of
surfactants on rate of penetration in rotary diamond drilling,” J. Pet. Sci. Eng. 134, pp.114–122.
http://dx.doi.org/10.1016/j.petrol.2015.07.027.
[29] Zhao, X. H., and Gao, F., 2013, “Initial research on the effect of inorganic salt on rate of penetration (ROP) in drilling
fluids,” Petrol. Sci. Technol. 31(4), pp.380–387. http://dx.doi.org/10.1080/10916466.2010.529554.

d
[30] Shi, X., Liu, G., Gong, X., Zhang, J., Wang, J., and Zhang, H., 2016, “An efficient approach for real-time prediction of

te
rate of penetration in offshore drilling,” Math. Probl. Eng. Article ID 3575380. http://dx.doi.org/10.1155/2016/3575380.
[31] Mohammed, H. Q., Abbas, A. K., and Dahm, H. H., 2018, “Wellbore instability analysis for Nahr Umr Formation in

di
southern Iraq,” 52nd US Rock Mechanics/Geomechanics Symposium (ARMA), Seattle, Washington, June 17–20, Paper
No. ARMA 18–916.

e
[32] Gholami, R., Aadnoy, B., Foon, L. Y., and Elochukwu, H., 2017, “A methodology for wellbore stability analysis in
anisotropic formations: A case study from the Canning Basin, Western Australia,” J. Nal. Gas Sci. Eng. 37, pp.341–360.

py
http://dx.doi.org/10.1016/j.jngse.2016.11.055.
[33] Abbas, A. K., Flori, R. E., and Alsaba, M., 2018, “Estimating rock mechanical properties of the Zubair shale formation

Co
using a sonic wireline log and core analysis,” J. Nal. Gas Sci. Eng. 53, pp.359–369.
https://doi.org/10.1016/j.jngse.2018.03.018.
[34] Abbas, A. K., Flori, R. E., Alsaba, M., Dahm, H., and Alkamil, E. H., 2018, “Integrated approach using core analysis
and wireline measurement to estimate rock mechanical properties of the Zubair reservoir, Southern Iraq,” J. Pet. Sci. Eng.

ot
166, pp.406–419. https://doi.org/10.1016/j.petrol.2018.03.057.
[35] Ahmed, M., Al-Shehri, H. A., Haidary, S. A., and Povstyanova, M., 2016, “A comprehensive geomechanical study to
tN
understand drilling challenges in the Manifa field offshore, Saudi Arabia,” Kingdom of Saudi Arabia Annual Technical
Symposium and Exhibition, Dammam, Saudi Arabia, April 25–28, Paper No. SPE-182833-MS.
https://doi.org/10.2118/182833-ms.
rip

[36] Zhang, J., 2011, “Pore pressure prediction from well logs: Methods, modifications, and new approaches,” Earth Sci.
Rev. 108(1–2), pp.50–63. http://dx.doi.org/10.1016/j.earscirev.2011.06.001.
[37] Eaton, B.A., 1969, “Fracture gradient prediction and its application in oilfield operations,” J. Petrol. Tech. 21 (10),
sc

pp.1353–1360. http://dx.doi.org/10.2118/2163-pa.
[38] Jaeger, J. C, Cook, N. G, and Zimmerman, R. W., 2007, “Fundamentals of Rock Mechanics,” fourth ed. Wiley-
Blackwell.
nu

[39] Najibi, A. R., Ghafoori, M., Lashkaripour, G. R., and Asef, M. R., 2017, “Reservoir geomechanical modeling: In-situ
stress, pore pressure, and mud design,” J. Pet. Sci. Eng. 151 pp.31–39. http://dx.doi.org/10.1016/j.petrol.2017.01.045.
[40] Gholami, R., Rabiei, M., Aadnoy, B., and Rasouli, V., 2017, “A methodology for wellbore stability analysis of drilling
Ma

into presalt formations: A case study from Southern Iran,” J. Pet. Sci. Eng. 167, pp.249–261.
http://dx.doi.org/10.1016/j.petrol.2017.11.023.
[41] Thiercelin, M., and Plumb, R., 1994, 1994, “A core-based prediction of lithologic stress contrasts in East Texas
ed

formations,” SPE Form. Eval. 9(04), pp.251–258. http://dx.doi.org/10.2118/21847-pa.


[42] Kidambi, T., and Kumar, G. S., 2016, “Mechanical earth modeling for a vertical well drilled in a naturally fractured tight
carbonate gas reservoir in the Persian Gulf,” J. Petrol. Sci. Eng. 141 pp.38–51.
pt

http://dx.doi.org/10.1016/j.petrol.2016.01.003.
[43] Abbas, A. K., Flori, R. E., and Alsaba, M., 2019. “Stability analysis of highly deviated boreholes to minimize drilling
ce

risks and nonproductive time,” ASME J. Energy Resour. Technol., 141(8), p. 082905.
http://dx.doi.org/10.1115/1.4042733.
[44] Chandrashekar, G., and Sahin, F., 2014, “A survey on feature selection methods,” Comput. Electr. Eng.; 40 (1), pp.16–
Ac

28. https://doi.org/10.1016/j.compeleceng.2013.11.024.
[45] Eskandarian, S., Bahrami, P., and Kazemi, P., 2017, “A comprehensive data mining approach to estimate the rate of
penetration: Application of neural network, rule based models and feature ranking,” J. Pet. Sci. Eng. 156, pp.605–615.
https://doi.org/10.1016/j.petrol.2017.06.039.
[46] Szlek, J., and Mendyk, A., 2018, “Fscaret: Automated feature selection from caret,” https://cran.r
project.org/web/packages/fscaret/index.html (accessed 3 November 2018).
[47] Kazemi, P., Khalid, M. H., Szlek, J., Mirtič, A., Reynolds, G. K., Jachowicz, R., and Mendyk, A., 2016, “Computational
intelligence modeling of granule size distribution for oscillating milling,” Powder Technol. 301, pp.1252–1258.
https://doi.org/10.1016/j.powtec.2016.07.046.

16

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Journal of Energy Resources Technology. Received December 29, 2018;
Accepted manuscript posted April 29, 2019. doi:10.1115/1.4043699
Copyright © 2019 by ASME
[48] Arabjamaloei, R., and Shadizadeh, S., 2011, “Modeling and optimizing rate of penetration using intelligent systems in
an Iranian southern oil field (Ahwaz Oil Field),” Petrol. Sci. Technol. 29(16), pp.1637–1648.
https://doi.org/10.1080/10916460902882818.
[49] Shadizadeh, S. R., Karimi, F. and Zoveidavianpoor, M., 2010, “Drilling stuck pipe prediction in Iranian oil fields: An
artificial neural network approach. Iran J. Chem. Eng. 7 (4): 29–41. http://www.ijche.com/article_10326_1700.html.
[50] Abbas, A. K., Rushdi, S., and Alsaba, M., 2018, “Modeling rate of penetration for deviated wells using artificial neural
network,” Abu Dhabi International Petroleum Exhibition and Conference (ADIPEC), Abu Dhabi, UAE, November 12–
15, Paper No. SPE-192875-MS. https://doi.org/10.2118/192875-ms.
[51] Aalizad, S. A., Rashidinejad, F., 2012, “Prediction of penetration rate of rotary-percussive drilling using artificial neural
networks – A case study,” Arch. Min. Sci. 57(3), pp.715–728. https://doi.org/10.2478/v10267-012-0046-x.
[52] Jahanbakhshi, R., Keshavarzi, R., Aliyari Shoorehdeli, M., Emamzadeh, A., 2012, “Intelligent prediction of differential

d
pipe sticking by support vector machine compared with conventional artificial neural networks: An example of Iranian

te
offshore oil fields,” SPE Drill. & Compl. 27 (04), pp.586–595. http://dx.doi.org/10.2118/163062-pa.
[53] Akande, K. O., Owolabi, T. O., and Olatunji, S. O., 2015, “Investigating the effect of correlation-based feature selection

di
on the performance of support vector machines in reservoir characterization,” J. Nal. Gas Sci. Eng. 22, 515-522.
http://dx.doi.org/10.1016/j.jngse.2015.01.007.

e
[54] Abbas, A. K., Flori, R. E., and Alsaba, M., 2018, “Laboratory Geomechanical Characterization of the Zubair Shale
Formation,” 52nd US Rock Mechanics/Geomechanics Symposium (ARMA), Seattle, Washington, USA, June 17–20,

py
Paper No. ARMA 18– 78.
[55] Abbas, A. K., Dahm, H. H., Flori, R. E., and Alsaba, M., 2018, “Laboratory Measurements of Petrophysical and

Co
Geomechanical Properties for Zubair Sandstone Formation in Southern Iraq,” 52nd US Rock Mechanics/Geomechanics
Symposium (ARMA), Seattle, Washington, USA, June 17–20, Paper No. ARMA 18–243.
[56] Soares, C., and Gray, K., 2019, “Real-time predictive capabilities of analytical and machine learning rate of penetration
(ROP) models,” J. Pet. Sci. Eng. 172, pp.934–959. http://dx.doi.org/10.1016/j.petrol.2018.08.083.

ot
[57] Chen, X., Gao, D., Guo, B., Feng, Y., 2016, “Real-time optimization of drilling parameters based on mechanical specific
energy for rotating drilling with positive displacement motor in the hard formation,” J. Nat. Gas Sci. Eng. 35, pp.686–694.
tN
http://dx.doi.org/10.1016/j.jngse.2016.09.019.
rip
sc
nu
Ma
ed
pt
ce
Ac

17

Downloaded From: https://energyresources.asmedigitalcollection.asme.org on 05/14/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like