You are on page 1of 48

Author’s Accepted Manuscript

A Computational Fluid Dynamics Based Artificial


Neural Network Model to Predict Solid Particle
Erosion

D.A. Pandya, B.H. Dennis, R.D. Russell

www.elsevier.com/locate/wear

PII: S0043-1648(16)30370-2
DOI: http://dx.doi.org/10.1016/j.wear.2017.02.028
Reference: WEA102082
To appear in: Wear
Received date: 4 October 2016
Accepted date: 10 February 2017
Cite this article as: D.A. Pandya, B.H. Dennis and R.D. Russell, A
Computational Fluid Dynamics Based Artificial Neural Network Model to
Predict Solid Particle Erosion, Wear,
http://dx.doi.org/10.1016/j.wear.2017.02.028
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 galley proof before it is published in its final citable 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.
A Computational Fluid Dynamics Based

Artificial Neural Network Model to Predict

Solid Particle Erosion

D. A. Pandyaa1, B.H. Dennisa, R.D. Russellb

a
Dept. of Mechanical and Aerospace Engineering, The University of Texas at Arlington;

b
Baker Hughes

Abstract

Solid particle erosion plays a critical role in the design and reliability of equipment employed in the oil

and gas industry. Significant erosion occurs due to solid particle loading, especially in applications

involving sand production. Low particle loading in drilling fluids (<10%) is also a source of erosion

inside downhole tools and in rig equipment at the surface. Accurate prediction of erosion rates can save

money and lives by predicting failure accurately and helping to maintain the safety of the equipment.

Empirical and mechanistic models to predict erosion were primarily developed based on observations of

extensive experiments and field studies. Computational fluid dynamics (CFD) has emerged as an

alternative tool to predict erosion in recent years. The ability to simulate multiphase flows in complex

geometries using CFD makes it a valuable and less-expensive method to predict erosion flow loop

experimentations and field trials. Various empirical relations have been established to predict erosion

1
Now with Shell Global Solutions US Inc.
using CFD. These methods often predict erosion regions accurately, but typically are highly inaccurate in

predicting an erosion rate. An order-of-magnitude error is observed in many cases.

This study employs machine learning approach along with CFD-based methodology to develop robust

erosion models. A generalized model is developed based on experiments conducted on 90-degree elbows

of 1-inch diameter and made from Inconel 718, Nickel Alloy 825, 25% Cr, Nickel Alloy 925, and 13% Cr

L-80 materials. The Baker Hughes erosion model developed in 2008 is studied as a baseline. Statistical

analysis was performed on CFD output parameters to identify those that most affect erosion rates. A

correlation analysis and non-parametric statistical analysis is performed resulting in the development of

two new regression models based on turbulent kinetic energy, and surface shear stress was developed. A

25-percent improvement is observed in the predictions of cumulative erosion rate error compared to

baseline. An artificial neural network with multilayer feed-forward model with the back-propagation

algorithm and Levenberg-Marquardt training was developed. This model, along with Bayesian

regularization, reduced cumulative error to less than 10%, compared to more than 40% in the baseline

Baker Hughes model.

Introduction

Solid particle erosion can be defined as material damage caused by solid particles that impinge a surface.

Erosion damage is seen in almost any industrial application where a solid-fluid multiphase flow occurs.

Particles carried across the streamline due to momentum, strike the walls and cause material damage.

Erosion can be the limiting factor in equipment design and might cause failure. Consequently, the erosion

process has been studied extensively for decades, although CFD-based erosion modeling is fairly recent

when compared to experimental investigations. Erosion is a complex phenomenon and is of great interest

in the oil and gas community. Accurate prediction of erosion rates can save money and lives by predicting

failure accurately and helping to maintain the safety of the equipment. Drill bits and other downhole tools
are affected by erosion from sand production at various high-rate wells. Low sand concentration flows

can result in significant erosion where production velocities are high.

Artificial neural networks (ANN) are systems inspired from human brain functions. ANN are based on

simulated nerve cells or neurons that are connected in various ways to form a network. Much like the

brain, ANN can learn, memorize and create relationships among what may seem like random data.

Advances in computing power have enabled the application of ANN to real-world problems. McCulloch

and Pitts (1943) developed an ANN model based on their knowledge of neurology. Frank Rosenblatt can

be credited for the development of the first simple learning algorithm called Perceptron (1958). In the

1970s and 1980s, significant innovations occurred in the field of ANNs. The adaptive resonance theory

(ART) network was developed, based on Grossberg and Carpenter’s school of thought, which explored

resonating algorithms (1985). Klopf (1972, 1975), developed a basis for learning in artificial neurons

inspired from the biological principle of neuronal learning called heterostasis. However, one of the most

important contributions to learning was made by Paul Werbos (1974, 1988), the so-called back-

propagation learning method. It is one of the most extensively used methods today in ANN. Rezaul Begg

(2006) calls back propagation “a Perceptron with multiple layers, a different threshold function in

neurons, and a more robust and capable learning rule.” Today, ANN are used for classification,

forecasting, modeling and pattern recognition in fields like business, engineering, science, and medicine.

Artificial neural networks have been successfully developed for reservoir modeling and estimation in the

oil and gas industry (Mohaghegh et al., 1995, Shahab et al., 1997). They have also been implemented in

log data analysis (H Eskandari et al., 2004) and for continuous oil field optimization Saputelli, (2002).

Velten et al. (2000) and Zhang et al. (2003) were among the first to implement ANN in analyzing the

wear of polymer composites. More recently, work by Suresh et al. (2009) reports successful

implementation of ANN in predicting solid particle erosion in composites. Shamshirband S. et al. (2015)

have developed adaptive neuro-fuzzy inference system (ANFIS) to predict total and maximum erosion

rates in a 900 elbow. The results are promising when compared to CFD. This work majorly differs in the
methodology of using a hybrid approach compared to stand alone soft computing technique developed in

Shamshirband S. et al.’s work. A multilayer feed-forward network with back-propagation training

algorithm has been used widely in wear prediction (Zhang et al. 2003), and the Bayesian regularization

training algorithm was found to be more accurate for erosion modeling by Zhang et al. (2003), Danaher et

al. (2004), as well as Suresh et al. (2009).

Factors Affecting Erosion

Erosion is a complex phenomenon that varies primarily as a result of material strength. In brittle

materials, erosion occurs by crack formation. Solid particles are hitting the surface form cracks, which

then propagate by subsequent impacts. Depending on the material hardness, when the impact load

exceeds a defined level, plastic defragmentation occurs along the fracture (Levy, 1995). Evans, Gulden,

and Rosenblatt (1978) explained how planar cracks form along the interface when the area is unloaded

after particles rebound from the surface and pieces are then removed by subsequent particle impacts.

Finnie (1960) presented one of the earliest analytical models to predict erosion in ductile materials. The

model assumes that erosion in ductile materials is a result of a micro-cutting process. In this process, the

surface material is removed in the form of cuttings. Particles that strike the surface at low impact angles

form a crater, and material is continuously removed by further impacts. This theory suffers from a

drawback in that it neglects the effect of particles hitting at a right angle. Finnie (1978) later corrected this

by modifying the model and considering the effect of surface material piling when particles impact the

surface at larger angles. A more detailed two-step model was presented by Tilly (1973), which stated that

first the impacting particles produce an indentation and may remove a chip. Subsequent impacts break up

the hole and project fragments radially from the primary site.

Particle impact velocity can be considered the single-most important factor causing erosion. Many

researchers conclude that the erosion rate (ER) is proportional to particle velocity (Vp) raised to the power

n.
Various values of n have been proposed. Finnie (1978) proposed two as the value of n. Smeltzer et al.

(1970) and Grewal et al. (2013) also found negative values for n by curve-fitting the experimental data in

some cases. More recently, Oka et al. (2005a, 2005b) proposed that the value of n was not a constant, but

a function of hardness. The impingement angle is also an important factor in determining erosion. At low

impact angles in ductile materials, cutting action and platelet formation is more prevalent, whereas

repeated plastic deformation is dominant in brittle materials (Sheldon, 1970).

Erosion occurs in various applications. The particles in each of these applications vary in hardness, size,

shape, etc. This variation in particle properties also affects erosion rate and nature. Particle properties like

size, shape, and hardness have been observed to affect erosion rate. Oka et al. (2005a, 2005b)

incorporated constants in the model to account for the effect of particle shape and size. The University of

Tulsa’s Mclaury (1997, 1998, and 1999) and Ahlert (1994) developed models that incorporated shape

factors ranging from 0.2 for round particles to 1.0 for sharp particles. Levy (1983) concluded that angular

particles could cause up to four times more erosion than round particles. Tilly (1973) proposed a very

well-fitted empirical model that considers erosion rate as a function of particle size. Particle size, in the

form of a power law, was also included in the Oka et al. (2005a, 2005b) model. The effect of particle

concentration was explained regarding a shielding mechanism by Andrews and Horsfield (1983). The

sand volume concentration ranged from 0.38% to 8.61% in tests by Turenne et al. (1990). They

discovered that the erosion ratio decrease followed a power law of sand volume concentration.

Fluid properties, such as the effect of turbulence were studied by Pourahmadi and Humphry (1983, 1990).

Impingements occurred in regions of high turbulence due to higher momentum transfer to the particles.

As fluid density and viscosity increase, the drag force acting on the particle also increases. This

phenomenon can be directly related to the conclusion of research conducted by Smeltzer et al. (1970) that

indicated the erosion rate decreases at higher test temperatures. Clark and Burmeister (1992) proposed a
squeeze film model to account for the cushioning effect of a fluid boundary layer. Clark (1992) observed

a phenomenon of boundary layer filtration in which particles were deflected due to high normal gradients

in a flow.

The novelty of this work lies in the hybrid approach of combining CFD and machine learning models like

regression and ANN to build a predictive modeling pipeline for erosion rates. General purpose CFD

code, ANSYS Fluent is used to simulate fluid flow and particle transport for corresponding experiments

on a 900 elbow. Exploratory data analysis and correlation analysis was performed on eight CFD output

parameters to identify their significance in erosion rate predictions. Two new models based on shear

stress and turbulent kinetic energy is developed by curve fitting. Experimental data of erosion rates is

used as a target variable, and CFD output is used as input to develop ANN model. A detail of the

workflow and data flow is shown in the figure below:

Figure 1 Modeling Workflow


CFD-Based Erosion Modeling

Computational fluid dynamics (CFD)-based erosion modeling has been employed and explored for many

years. CFD-based erosion models offer a low-cost, fast solution to the erosion modeling problem.

However, erosion modeling using CFD is a complex task because it involves various phenomena that

must be considered. Flow modeling, particle interaction with fluid, walls and other particles, fluid

properties, and erosion modeling are some of these phenomena. Also, geometric modeling and meshing

(discretization of the domain) are critical in the overall CFD-based erosion modeling procedure. The

biggest advantage of CFD-based erosion modeling over other simplistic models is that complex

geometries and flow fields can be easily modeled using CFD. It is a powerful tool that can predict erosion

regions accurately. Consequently, it presents a low-cost, qualitatively accurate and timely method to

predict erosion in various tools and geometries. The Erosion Corrosion Research Center (ECRC) at The

University of Tulsa has been one of the pioneers in CFD-based erosion prediction research. McLaury

(1993, 1996) proposed a widely followed CFD-based erosion prediction procedure. More recently, Zhang

et al. (2009) presented a procedure and improved guidelines for using commercial CFD code for erosion

prediction. The authors have also recently published work (Pandya, Dennis, and Russell, 2014) presenting

the effect of CFD modeling parameters on erosion prediction and a new improved CFD-based erosion

model. The work in this paper can be considered an extension of this previous work.

CFD-based erosion modeling consists of fluid flow modeling, discrete particle modeling, and erosion

modeling. Detailed steps are presented in Figure 1.


Figure 2 Steps in CFD-based erosion modeling

ANSYS Fluent is a finite-volume CFD code that solves the Navier-Stokes equation to model fluid flow.

ANSYS automated meshing was used to discretize the domain. An unstructured mesh for flow region

with structured refinement on the walls is used. The wall y + values of around 30 are observed, and

scalable wall functions are used to simulate flow in near-wall regions. The realizable k-epsilon model was

used to model turbulence parameters. It has proved to be a better at capturing physics of localized

circulation in the elbow when compared with Standard k - epsilon. (Pandya, Dennis, and Russell, 2014).

Simulation is initiated with a first order discretization scheme and then switched to second order scheme

for pressure, Momentum, TKE, and dissipation rate. A SIMPLE scheme is used for pressure – velocity

coupling. The discrete phase (sand) was solved by tracking a large number of particles Lagrangian

approach after fluid field simulations were converged. In the Lagrangian approach, the fluid (water in this

case) is treated as the continuum, but the discrete phase (sand) is treated as single particles, where particle

trajectories, representing a stream of particles, are calculated as a result of forces acting on them. Water

is the primary phase, and the Sand particle is secondary phase in this multiphase simulations. The

secondary phase (sand) in our case had a very low concentration, so this approach was justified. The
effect of sand particles on fluid was considered negligible for low concentrations. This model gave

various output parameters such as particle velocity, particle impingement angle, wall shear stress, etc.

Depending upon which parameters were required in the empirical erosion model, erosion rates were

calculated as a post-processing step and displayed using custom field functions or user-defined functions

within ANSYS Fluent. A detailed modeling approach presented by authors (Pandya, Dennis, and Russell,

2014) was followed, the details of which are out of scope for this paper.

Baker Hughes Erosion Tests and Model

All analysis performed in this research was based on experimental results provided by Baker Hughes.

Tests were conducted at Baker Hughes for four different flow conditions considering two different water

flow velocities (i.e., 50 ft/sec and 85.8 ft/sec) and for two different particle sizes (i.e., 256 microns and 25

microns) (Russell,2004). Flow velocities of 50 ft/sec and 85.8 ft/sec correspond to Reynolds number of

3.83 x 105 and 6.5 x 105 respectively. The test was performed on a 90-degree elbow with a 1- inch inner

diameter. Inconel 718, Nickel Alloy 825, 25% Cr, Nickel Alloy 925, and 13% Cr L-80 material elbows

were placed in series. The sand concentration was approximately 1% by mass and 0.38% by volume

(Russell, 2004). The erosion rate was measured at 230 locations in each elbow. The experimental setup is

shown in a schematic in Figure 2. Detailed explanations of experimentation procedure are presented by

Mark McCasland et.al. (2004) and Ronnie Russell et. al. (2004)
Figure 3 Experimental setup for erosion experiments (Russell, 2004)

The ERC-2003 erosion model was developed by Baker Hughes in 2003. Ronnie Russell et al. (2004)

presented this model that was based on extensive experiments on a 90-degree elbow. A series of erosion

tests were performed with several materials and under various flow conditions and sand concentrations.

More than 95% of the impact angles were between 0-30 –degree with few around 90-degree in the

stagnation zone. A predictive model was then created based on the obtained results. The model is

primarily used for low-angle flows.

The Baker Hughes ERC-2008 model offers some accuracy improvements over the Baker Hughes ERC-

2003 model for flow velocities less than 30 ft/sec (Russell and Marsis, 2013). Baker Hughes ERC -2008

is used as the baseline in the study.


Correlation Analysis

Different data analysis techniques such as regression, exploratory data analysis, and artificial neural

networks were used in this work. MATLAB was used for all data analysis. A correlation analysis was

performed on outputs of CFD erosion models and experimental results. The parameters with high

correlation were identified for model generation. Regression for curve fitting and artificial neural network

models were developed by considering the erosion rate as a function of these identified parameters.

Two variables are said to be associated if the behavior of one affects the behavior of the other.

Correlation coefficients are measures of association. They assign a numerical value to the degree of

association or strength of the relationship between two variables (Gibbons, 1993). Several correlation

coefficients are proposed to measure a degree of correlation. Pearson correlation coefficient (Pearson,

1901), Spearman’s rank correlation coefficient and Kendall’s rank correlation coefficient (Kendall, 1948)

were obtained to establish the association between various independent parameters to erosion, the

dependent parameter in each case. A general interpretation based on the values of these coefficients

(ranging from -1 to 1) is as follows.

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1

Weak Strong positive


Strong negative Weak Little positive correlation
correlation negative correlation correlation
correlation

Pearson’s correlation is a measure of the linear correlation between two variables. It is also called

Pearson-product-moment correlations and is represented by ‘r’. Linear correlation may exist, even if

variables have a nonlinear relation to one another. Correlation varies from -1 to 1, where 0 means no
correlation, 1 means complete positive linear correlation, and -1 means complete negative linear

correlation.

Spearman’s rank coefficient is a nonparametric measure of dependence. As most of the variables

encountered in this work do not have a strong linear or parametric relation to erosion, a nonparametric

approach is important to consider while performing correlation analysis. Spearman correlation can give a

perfect value (1 or -1) when the two variables are related by any monotonic function, in contrast to only a

linear function for the Pearson correlation (Gibbons, 1993).

Like Spearman’s rho, Kendall’s tau is also a nonparametric measure of correlation. It was first introduced

by Maurice Kendall (1948). Although in many cases, as in ours, the interpretation of Spearman’s rho and

Kendall’s tau is very similar, some people have argued that Kendall’s tau has an advantage of very direct

observation and interpretation, considering agreeable (concordant) and non-agreeable (discordant) pairs

(Bolboaca and Jantschi, 2006). Details about the correlation coefficients and their formula can be found in

Appendix A

The data are resampled by bootstrapping to reaffirm the inference made based on parameter values.

Bootstrap resampling has a great advantage of being a simple and straightforward way to derive

properties of estimators. It measures the properties from some resamples created by random data

sampling with replacement from the data set. The number of elements in each bootstrap sample is equal to

that in the original data set. It helps to determine how certain are the conclusions made on a parameter

value.

Curve Fitting and Regression

A variety of fits, algorithms, and methods are explored to find the best fit for the identified CFD

parameters. MATLAB curve fitting toolbox is used for this purpose (2013). For all the methods

implemented, input data is first centered and scaled to normalize the predictor (input) data. The input
variables have wide differences in scale in our data set. Consequently, normalizing the data improved the

fit. A weighted least-squares method is helpful when the assumption that the response data is not of equal

quality. The method improves fit by adding weight (scale factor) in the fitting process. The weight

determines how much the response value influences the parameter estimates (i.e., a high-quality data

point influences the fit more than a low-quality data point). The disadvantage of least squares as being

highly sensitive to outliers is rectified in the robust scheme. Two methods are available in MATLAB

(2013) for robust least-squares fitting:

Least-absolute residuals (LAR): To reduce the effect caused by squaring the absolute residuals of outliers,

LAR calculates the curve that minimizes the absolute difference. This reduces the influence of outliers on

the fit.

Bisquare weights: Weights are assigned to each data point based on its distance from the fitted line. Points

that are much farther from the line get a zero weight. This minimizes the weighted sum of squares. The

Bisquare method is preferred most of the time because it finds the curve that fits the majority of the data

utilizing least-squares approach and neglects the outliers.

A robust iteratively re-weighted least-square algorithm is proposed by Holland and Welsh (1977).

A detailed discussion about development and implementation of each of the methods and algorithms is

out of the scope of this work, MATLAB implementation of Robust fitting with Bisquare weights (2013)

is presented in Appendix B, and it can be obtained from respective reference. Its application in MATLAB

is presented in Matlab Documentation (2013).

Artificial Neural Networks

The basic building blocks of an ANN are network architecture, transfer functions, and training methods.

A multilayer network is a network with multiple layers with different weight and biases for each layer.

Each layer may have different network architecture. The layer that produces the final output of the
network is called output layer. The remaining layers are usually called hidden layers. A simple three-layer

network with an input layer, one hidden layer, and one output layer is depicted in Figure 3.

Figure 4 Schematic of a typical multilayer neural network model (Hagan et al., 1996)

Linear and hyperbolic tangent sigmoid transfer functions were used in this study. As the name suggests, it

is an identity function given by:

Advantages of implementing sigmoidal functions to accelerate convergence in back-propagation

algorithms (discussed later in detail) have been discussed by Vogl et al. (1988) and Harrington (1993).

The functional representation is

The method of setting weights in an ANN enables the process of learning. Training is generally referred

to as the process of modifying the weights in the connection between network layers to achieve expected

output (Shivanandam at el. 2006). The process that takes place while the network is trained is called

learning. There are different types of training including supervised, unsupervised and reinforcement. The
back-propagation algorithm used in this study is a type of supervised training, where the network is

provided with sample inputs and the outputs are compared to expected responses. There are target output

vectors for comparing the sequence of training input vector. The weights are adjusted according to

different algorithms.

Multilayer Feed-forward Network with Back-propagation Algorithms

A feed-forward network with the back-propagation algorithm (Rumelhart and McClelland, 1986) is one

of the most commonly used networks. In this network, neurons are organized in layers and send their

signals "forward." The errors are then propagated backward. The back-propagation algorithm falls under

the category of supervised learning. The central idea for a back-propagation algorithm is that of

minimizing the error until the network is trained. As in many other networks, random weights are

assigned at the beginning of training, and they are adjusted until the minimization criteria for error is

reached. Hagan and Menhaj (1994) first demonstrated the development of the Levenberg-Marquardt

algorithm for neural networks. It has been reported to train the neural network 10 to 100 times faster than

the popular gradient descent back-propagation method. A Bayesian regularization algorithm to avoid

over-fitting of data is introduced to the Levenberg-Marquardt algorithm. A detailed explanation of the

methods is available in MacKay (1992) as well as Foresee and Hagan (1997).

A detailed derivation of any of the methods is available in the literature mentioned and will not be

discussed here. However, a top-level overview of the steps involved is presented.

Bayesian regularization is implemented in the Levenberg-Marquardt algorithm to minimize a liner

combination of squared errors and weights. This implementation is one of the approaches to stop over-

fitting a problem. It also reduces the need to test a different number of hidden neurons for a problem.

The typical performance function for a feed-forward network is a mean sum of squares of the errors.

A term consisting of a mean sum of squares of the network weights is added to performance to improve

generalization.

Where α and β are parameters optimized in the Bayesian framework (MacKay, 1992). This results in

smaller weight and forces a smooth response. In this framework, the weights are assumed to be random

variables with a specified distribution. The regularization parameters are associated with the variance of

these distributions. The parameters are then estimated by statistical techniques.

More details, proof, and derivation of the algorithms can be obtained in MacKay (1992) and Foresee and

Hagan (1997).

Results and Discussion

Before we discuss the results obtained by implementing various erosion-prediction models, it is important

to discuss the parameters used throughout results to measure efficiency and effectiveness of the models.

Following are quantities measured and their implications are related to results in this work.

The most frequently used quantity to measure the quantitative performance of the model throughout this

work is percentage error defined by the equation below:

The lower the percentage error, the better the model.


The ratio of the predicted value to the experimental value is a good measure to visualize the results when

plotted. An ideal situation would be when this value is 1, which would mean that predicted value and

experimental values are the same.

The sum of squares due to error, denoted by SSE, measures the total deviation of the predicted value form

the experimental value. It is also sometimes called a sum squared of residues.

Values closer to 0 mean the model has a smaller random error and is a better fit.

R-squared is the most widely used measure to determine a quality of fit for statistical models like

regression or neural network. It measures how well the model explains the variation in data. It is called

the coefficient of determination.

R-square is defined as the ratio of the sum of squares of regression (SSR) and the total sum of squares

(SST) in the following way:


Also, SST= SSR+SSE

Hence,

Values closer to 1 indicate a better fit.

The root mean squared error, is commonly called the standard error of fit or standard error of regression.

It is a measure of estimation of the standard deviation of the random component in the data:

√ ∑

Lower values characterize a better model.

Correlation analysis was performed to identify parameters that most affect erosion rate. This analysis was

performed on 920 data points obtained from CFD analysis of all four test cases. The aim was to develop a

more robust erosion rate prediction model. Three different correlation coefficients were calculated.

Pearson coefficient ‘R' is most commonly used correlation coefficient. However, it evaluates the linear

correlation between parameters (Note: a linear correlation between parameters with the nonlinear

relationship is possible). Consequently, Kendal's ‘tau' and spearman's ‘rho,' both of which rank
correlations, were also calculated. Values closer to one depict strong correlations. Figure 4 shows

correlation coefficient for various parameters.

Correlation Coefficients
0.0229
Impact Angle 0.0189
0.025
Turbulent Kinetic 0.9
0.79
Energy 0.76
0.78
Surface Shear stress 0.73
0.72
0.49
Concentration 0.33
0.29
0.69
Velocity 0.63
0.62

0 0.2 0.4 0.6 0.8 1


Spearman's 'rho' Kendal's 'Tau' Pearson Coefficient 'R'

Figure 5 Correlation of CFD parameters VS erosion rate

A bootstrap resampling method is applied to calculate the correlation coefficient for 100 resamples. The

results further support our statistical inference of the correlation analysis mentioned above. Pearson’s

coefficient for each of these resampled data is calculated, and distribution is presented in Figure 6.
Figure 6 Bootstrapping: Resampling distributions of the Pearson’s correlation coefficients

The results agree well with the correlation analysis performed. It can be observed that turbulent kinetic

energy is identified as most influential parameter followed by surface shear stress, velocity, and

concentration. The impact angle does not have any conclusive correlation with erosion rate prediction. It

is also verified when overall performance of ANN was degraded by including impact angle as an input

variable in the model.

Concentration has some positive correlation with the erosion rate. An analysis is performed to check if the

Discrete particle (DPM) –Sand concentration can be used as a conditioning parameter in velocity-based

erosion models. A co-plot for velocity, concentration and erosion rate presented in in Figure 7 shows that

for every range of concentration, a similar trend line is observed. Consequently, use of DPM

concentration as a conditioning parameter in the present model is validated.


Figure 7 Velocity, concentration, erosion rate co-plot

Surface shear stress has a strong positive correlation with the erosion rate, so curve fitting through

regression was performed to develop an erosion model as a function of surface shear stress. Two different

models were developed, a polynomial model and exponential model. Results are presented in Table 1.
Table 1 Shear stress model - goodness-of-fit parameters

SSE Adjusted RSME Percentage

R -square error

Polynomial fit 5.94E-05 0.9919 2.55E-04 47

Exponential fit 2.43E-05 0.9967 1.63E-04 42

An exponential fit has a lesser percentage error of 42-percent mean error. A regression plot for an

exponential plot is shown in Figure 8.

Figure 8 Exponential curve fit for surface shear stress model


Pearson's rank correlation coefficient showed an unyielding positive correlation between TKE and

erosion rate. Thus, as expected, the TKE-based model proved to be very efficient. Table 2 presents

goodness-of-fit data for TKE-based models

Table 2 TKE models - goodness-of-fit parameters

SSE Adjusted RSME Percentage

R -square error

5th-degree polynomial 2.10E-07 0.9801 1.52E-05 31

Exponential 4.60E-08 0.9935 7.10E-06 29

Three different models were developed including linear fit, fifth-degree polynomial fit, and an

exponential fit. Only exponential fit, which showed the best results, is presented in Figure 9.

Figure 9 TKE exponential fit model


The exponential fit for TKE vs. erosion rate gives the best predictions. It also shows excellent goodness-

of-fit parameter values. The R-squared value is 0.9972, which is one of the best seen so far. Also, the

mean percentage error is 29% which is excellent. It should also be observed that TKE model provides a

better fit for large range as seen from Figure 9 above.

An ANN model was developed as a black box to predict the erosion rate. The four most influential

parameters, velocity, concentration, TKE and surface shear stress, were taken as an input parameter to the

ANN and the erosion rate was considered target variable. A multilayer feed-forward/back-propagation

algorithm with Lavernberg-Marquardt training was implemented. A Bayesian regularization on the

training algorithm gave the best results. More than 45 different networks were modeled and analyzed. The

description and goodness-of-fit parameters for 11 different ANN models are presented in Table 3.
Table 3 ANN model description and goodness-of-fit parameters

Model Number Model R-Square value RSME

1 4-[20]-1 0.8757 7.2e-6

2 4-[35]-1 0.8987 6.3e-6

3 4-[50]-1 0.9224 2.4e-6

4 4-[10-10]-1 0.9654 1.1e-6

5 4-[20-20]-1 0.9845 1.8e-6

6 4-[10-20]-1 0.9799 5.1e-6

7 4-[40-40]-1 0.9523 6.9e-6

8 4-[10-10-10]-1 0.9664 2.7e-7

9 4-[15-10-15]-1 0.9767 4.2e-7

10 4-[25-10-25]-1 0.9814 6.8e-7

11 5-[10-10-10]-1 0.667 8.3e-4

Figure 10 shows the percentage error for each of the 11 networks described.
Percentage error
90
80
Percentage Error 70
60
50
40
30
20
10
0
0 2 4 6 8 10 12
ANN Model number (Refer Table 90 )

Figure 10 Error percentage for ANN models

Input parameters for 4-[10-10-10]-1 are velocity, DPM concentration, surface shear stress and turbulent

kinetic energy and target are experimentally observed erosion rates. A network representation can be seen

in Figure 11. Weights are represented by ‘w’ and biases by ‘b’

Figure 11 Neural network representation for 4-[10-10-10]-1

A regression plot for ANN with three hidden layers and structure 4-[15-10-15]-1 with and without

Bayesian regularization is presented next. Bayesian regularization avoids over-fitting the model and

significantly improves results. Consequently, it is important to appreciate the comparison. A clearly

visible better fit is obtained by Bayesian regularization. The mean percentage error is reduced from 27%

to as low as 7% after regularization is forced on the Lavernberg-Marquardt training algorithm. The 5-[10-
10-10]-1 network represents the inclusion of impact angle as an input parameter, but it has a negative

effect on network performance.

Figure 12and Figure 13 show a graph, plotting measured and predicted values for all points without and

with Bayesian regularization.

Without Bayesian Regularization


4-[10-10-10]-1
Predicted value

Experimental value

Figure 12 ANN model without Bayesian regularization


With Bayesian Regularization
4-[10-10-10]-1

Predicted Value

Experimental value

Figure 13 ANN model with Bayesian regularization

To verify the reproducibility of the results by ANN , a bootstrap sampling on input parameter was

performed. A 100-sample input was created from the current data set. The mean percentage error is

presented in a histogram in Figure 14.


100
ANN Percentage Error Bootstrap
90
80
70
60
Frequency

50
40
30
20
10
0
1
6

56
11
16
21
26
31
36
41
46
51

61
66
71
76
81
86
91
96
Precentage Error

Figure 14 ANN mean percentage error bootstrap check

The best of each model was selected, and their prediction capability was compared regarding percentage

error (Figure 15) and error ratio (Figure 16). The Baker Hughes model was the base case for comparison.
Overall Percentage Error comparision

ANN 7.1

TKE 29

Shear stress 42

Baker Hughes 57

0 10 20 30 40 50 60
Mean percentage error

Figure 15 Percentage error comparison

Error Ratio plot


Baker
100 Hughes-
Predicted value/Experimental value

New

10
TKE Model

0.1 ANN Model

Upstream Downstream
0.01
0 2 4 6 8 10 12
Non dimensional point location (0-5 upstream, 5-12 downstream)

Figure 16 Error ratio plot


All models were developed based on experimental results of an elbow case. Two different sand

concentration and velocities were tested, but it was still the same elbow case. Consequently, there was no

change in geometry. Verification was necessary with a different geometric case. A hold out data

validation is vital for any machine learning approach. Experimental results for a test on a completion tool

were compared with the models developed. Due to proprietary data restrictions, the geometry of the tool

is not presented here. To give a comparative idea, flow conditions will be discussed. The erosion rate of

the tool was measured for 2% sand loading with 53-micron sand size and 66.47 ft/sec inlet velocity.

Erosion rate at 243 points was measured and compared with simulated results from the new Baker

Hughes model, the shear stress model, the turbulent kinetic energy model, and the ANN model.

Regression results are presented below in Figures 16. Mean percentage errors are represented in a graph

below.
Figure 17 Predicted vs. Experimental Erosion Rate (mm/hr.) for Model verification

It can be observed that in the verification case that each model performs better in comparison to elbow

erosion rate prediction. This can be attributed to the fact that the verification model geometry does not

have major flow path variations. The model has an essentially undisturbed streamline flow and is more
comparable to upstream points in case of the elbow. Even in the elbow, the mean error for upstream

points is much lower compared to downstream where complex flow structures like vortices are generated.

Conclusion

CFD best practice was followed to simulate fluid flow and particle transport in a 90-degree elbow.

Experimental results were considered as the target parameter, and three new models based on statistical

analysis of CFD output parameters were developed. Correlation analysis was performed on eight different

CFD output parameters to identify ones affecting erosion rate prediction the most. Velocity, surface shear

stress, DPM concentration and turbulent kinetic energy were found to have a significant correlation.

A new shear stress-based erosion rate prediction model was developed by implementing robust least-

square curve fitting. The new shear stress-based model had an average percentage error of 42%. A

turbulent kinetic energy-based erosion model was also developed, and it exhibited and error of 29%,

which is an improvement of about 20% from the baseline model of Baker Hughes ERC -2008.

Artificial neural network models were explored to act as a black box where CFD output was taken as

input parameter and erosion was the output. More than 45 different networks with a different number of

neurons in hidden layers as well as a different number of hidden layers were modeled. A multilayer feed-

forward network with a back-propagation algorithm was implemented for each model. Training

algorithms like Scaled Conjugate gradient and Levenberg-Marquardt were implemented. A Bayesian

regularization was implemented on Levenberg-Marquardt to avoid over-fitting while training. This

approach gave the best results. The results from the ANN model were excellent, with the mean percentage

error of 7%, which is a further improvement of more than 20% points and 300% from the shear stress or

the TKE model. Overall, it brings down error from 57 percent to 7 percent compared to the baseline.

ANN’s have potential to be developed further as a tool to predict erosion rate. The ANN model was also

verified on a hold out data set consisting of CFD analysis on a completion tool. Developed model had not

seen the data from completion tool at all and hence served as a good candidate for model verification.
ANN model showed improvement in predicting erosion rates on completion tool when compared to the

elbow case. This is counter-intuitive, but the verification model has very simple flow path with no

circulation regions. It can be compared to upstream data points of the elbow where erosion rate prediction

capabilities were found to be more accurate when compared to downstream points. Consequently, the

models were successfully verified on a different geometry.

All the models were developed based on experimentation on elbows. This is just one case, so care should

be taken not to generalize it. Experimentation should be performed on more complex tools that are used in

the oil and gas industry and which face daily erosion problems. This approach will help in the

development of a better model.

For any statistical model, more data points lead to better models, and this can be achieved by more

experimentation.

Material properties of the tool are believed to have the great effect on erosion rate. Dependence of erosion

rate on material properties like hardness or tensile strength should be explored.

Lastly, although our studies show no significant effect of impact angle, other experimental studies have

found it to be a significant parameter affecting erosion (Finnie et al. 1992, Sheldon 1970). More effort

should be made to implement a custom impact-angle function into erosion models.

Appendix A – Correlation coefficients

Pearson’s correlation coefficient:

For paired x and y values in a sample, Pearson’s correlation coefficient r is given by:

r=
r=

r
 ( x  X )( y  Y )
 ( x  X )    ( y  Y )
2 2


Nonlinearity and outliners are a major factor that affects the value of r.

Spearman’s rank coefficient:

Spearman rank coefficient is calculated using the following formula:

Where d is the difference in rank between the two variables.

After the data are collected, it is ranked by giving highest rank to the highest value, and so on. Then the
difference in rank of each data set is calculated, and the values are inserted in the formula above.

Kendall’s Tau

The general formula for calculating the Kendall’s Tau is given below:

For a given sample of variables x and y, both with sample size of ‘n’, there are nC2 combinations possible
for selecting distinct pairs (xi., yi) and (xj., yj). These pairs are defined as concordant if (xi. > xj and yi >
yj ) or (xi. < xj and yi < yj ); discordant if (xi. > xj and yi < yj ) or (xi. < xj and yi > yj ) and neither if
xi = xj or xi = yj .

Then the number of concordant pairs and discordant pairs are inserted into the above formula to derive a
coefficient that lies between -1 to 1.
Appendix B – MATLAB implementation of Robust fitting with Bisqare
weights

Following steps are implemented for Robust fitting with Bisquare weights (2013).

1. Fit the model by weighted least square.

2. Compute the adjusted residuals given by:

ri is the usual least-square residuals and hi are the leverages that adjust the weight reducing weight of data
points that had significant effects on the least-square fit.

3. Standardize the adjusted residues by:

Where K is the tuning constant with a value 4.685 and s is the robust variance given by:

4. Compute robust weights as a function of u:

| |

| |

5. Stop if fit converges, else go back to step one for the next iteration.
A nonlinear least-squares method is employed to fit a model that has an equation with nonlinear
coefficients or is a combination of linear and nonlinear coefficients. Exponential and power function fits
were obtained using nonlinear methods.

The general matrix from of the model is given by:

y = f(X,β) +ε

where,

y is a n-by-1(one-dimensional vector of n elements) vector of responses

β is an m-by-1(one-dimensional vector of m elements) vector of coefficients

X is a n-by-m (matrix with n rows and m columns) design matrix for the model

And ε is a n-by-1 (one dimensional vector of n elements) error vector.

An iterative approach is required to solve for the coefficients in a nonlinear model. The following steps
are performed by a MATLAB function to solve for the coefficients in a nonlinear least-square regression
for curve fitting (2013):

1) Estimate the value of each coefficient.

2) Calculate the fitted response values given by

3) Adjust the coefficients following a fitting algorithm, and see if the fit improves. Two different
fitting algorithms are available in MATLAB:

a. Trust-Region – This algorithm was described by More and Sorensen (1983) and Cartis et
al. (2009). It is used if the constraints on coefficients are known.

b. Levenberg-Marquardt – This is one of the most popular algorithms used for decades.
Proposed by Levenberg (1944) and improved upon by Marquardt (1963), it has proven to
give a good fit for a wide range of nonlinear models.

4) Stop if the convergence criteria are reached, else return to step 2 for next iteration.

Appendix C – ANN algorithms

Back –Propagation algorithm

The steps of back-propagation for a connection between hidden layer neuron A and output neuron B are
explained in below (MacLeod, 2013):
1) Initialize random weights, apply inputs and calculate the outputs.

2) Calculate the error for neuron B. Remember that The transfer function is sigmoid function, so the
error is:

3) Update the weights by the following equation.

4) Calculate the error for the hidden layer neuron by back-propagating them from the output layer.
This is done by taking errors from output neurons and running them back through the weight to
get hidden layer errors. If a neuron is connected to two output neurons, B and C, then:

5) Go back to 3 to update the weight of hidden layer on the left.

Levernberg- Marquardt training

The training process for the Lavenberg-Marquardt algorithm is described in the following steps (Yu and
Wilamowski, 2012):

1) With the initial weights (randomly generated), evaluate the total error (SSE).

2) Perform the update as directed by the Levenberg-Marquardt update rule to adjust weights with the
new weights and evaluate the total error.

3) If the current total error increases as a result of the update, then retract the step (such as reset the
weight vector to the previous value) and increase the combination coefficient μ by a factor of 10
or by some other factor. Go to step ii and try an update again.

4) If the current total error decreases as a result of the update, then accept the step (such as keep the
new weight vector as the current one) and decrease the combination coefficient μ by a factor of
10 or by the same factor as step 4.
5) Go to step 2 with the new weights until the current total error is smaller than the required value.

Lavenberg-Marquardt with Bayesian regularization:

An expanded Levenberg-Marquardt Algorithm with Bayesian regularization is presented below (Souza,


2009):

1) Compute the Jacobian.

2) Compute the error gradient.

3) Approximate the Hessian matrix (matrix of second partial derivative).

4) Calculate the performance function.

5) Solve to find δ.

6) Update the weights w using δ.

7) Recalculate performance function using updated weights.

8) If the value has decreased, then

a. Discard the new weights, increase λ using γ and go to step 5.

b. Else, decrease λ using γ.

9) Upgrade the Bayesian hyperparameter using MacKay’s formulae.

a. where W is a number of network parameter (weights and


biases) and a tr= trace of the matrix

b. ⁄ where N is the number of entries in training set

c. ⁄
Reference

2006. ANSYS FLUENT user's manual. Software release 6: 449-56

2013 MATLAB Curve Fitting Toolbox Documentation. Version 3.4. Natick, Massachusetts: The

MathWorks Inc.

2013 MATLAB R2013b Documentation. Version 3.4. Natick, Massachusetts: The MathWorks Inc.

Ahlert, K.R. 1994. Effects of particle impingement angle and surface wetting on solid particle erosion of

AISI 1018 Steel. MS Thesis, University of Tulsa, Tulsa, Oklahoma (1994)

Andrews, D. and Horsfield, N. 1983. Particle collisions in the vicinity of an eroding surface. Journal of

Physics D: Applied Physics 16 (4): 525.

Andrews, D. and Horsfield, N. 1983. Particle collisions in the vicinity of an eroding surface. Journal of

Physics D: Applied Physics 16 (4): 525.

Begg, R. and Palaniswami, M. 2006. Computational intelligence for movement sciences: neutral networks

and other emerging techniques: IGI Global.

Bolboaca, S. and Jäntschi, L. 2006. Pearson versus Spearman, Kendall's tau correlation analysis on

structure-activity relationships of biologic active compounds. Leonardo Journal of Sciences 5 (9): 179-

200.

Carpenter, G.A. and Grossberg, S. 1985. Category learning and adaptive pattern recognition: A neural

network model. Presented at the Proceedings, Third Army Conference on Applied Mathematics and

Computing, ARO Report.

Curtis, C., Gould, N.I. and Toint, P.L. 2009. Trust-region and other regularizations of linear least-squares

problems. BIT Numerical Mathematics 49 (1): 21-53.


Clark, H. and Burmeister, L. 1992. The influence of the squeeze film on particle impact velocities in

erosion. International Journal of Impact Engineering 12 (3): 415-26.

Danaher, S., Datta, S., Waddle, I. and Hackney, P. 2004. Erosion modeling using Bayesian regulated

artificial neural networks. Wear 256 (9): 879-88.

Demuth, H. and Beale, M. 1993. Neural network toolbox for use with MATLAB. Version 4. Natick,

Massachusetts: The MathWorks Inc

Drukker, D.M. 2003. Testing for serial correlation in linear panel-data models. Stata Journal 3 (2): 168-

77.

Eskandari, H., Rezaee, M., and Mohammadnia, M. 2004. Application of multiple regression and artificial

neural network techniques to predict shear wave velocity from wireline log data for a carbonate reservoir

South-West Iran. CSEG recorder 42: 48.

Evans, A., Gulden, M. and Rosenblatt, M. 1978. Impact damage in brittle materials in the elastic-plastic

response regime. Proceedings of the Royal Society of London.A.Mathematical and Physical Sciences 361

(1706): 343-65.

Finnie, I. and McFadden, D. 1978. On the velocity dependence of the erosion of ductile metals by solid

particles at low angles of incidence. Wear 48 (1): 181-90.

Finnie, I. 1960. Erosion of surfaces by solid particles. Wear 3 (2): 87-103. doi: 10.1016/0043-

1648(60)90055-7.

Finnie, I., Stevick, G. and Ridgely, J. 1992. The influence of impingement angle on the erosion of ductile

metals by angular abrasive particles. Wear 152 (1): 91-8.


Foresee, F. and Hagan, M.T. 1997. Gauss-Newton approximation to Bayesian learning. Presented at the

International Conference on Neural Networks. Houston, Texas 9-12 June,

https://dx.doi.org/10.1109/ICNN.1997.614194.

Gibbons, J.D. 1993. Nonparametric measures of association: Sage University Paper series on

Quantitative Application in Social Sciences, 07-091, Newbury Park, California: Sage.

Grewal, H.S., Arora, H.S., Agrawal, A. et al. 2013. Slurry Erosion of Thermal Spray Coatings: Effect of

Sand Concentration. Procedia Engineering 68: 484-90.

doi: http://dx.doi.org/10.1016/j.proeng.2013.12.210.

Hagan, M.T., and Menhaj, M.B. 1994. Training feedforward networks with the Marquardt

algorithm. Neural Networks, IEEE Transactions on 5 (6): 989-93.

Harrington, P.d.B. 1993. Sigmoid transfer functions in backpropagation neural networks. Analytical

Chemistry 65 (15): 2167-8.

Holland, P.W., and Welsch, R.E. 1977. Robust regression using iteratively reweighted least-squares.

Communications in Statistics-Theory and Methods 6 (9): 813-27.

Humphrey, J. 1990. Fundamentals of fluid motion in erosion by solid particle impact. International

Journal of Heat and Fluid Flow 11 (3): 170-95.

Kendall, M.G. 1948. Rank correlation methods, first edition. London: Charles Griffin & Co. Ltd.

Klopf, A.H. 1975. A comparison of natural and artificial intelligence. ACM SIGART Bulletin (52): 11-3.

Klopf, A.H. 1972 Brain function and adaptive systems: a Heterostatic Theory. Special Report No. 133

Bedford, Massachusetts: Air Force Systems Command

Lee Rodgers, J. and Nicewander, W.A. 1988. Thirteen ways to look at the correlation coefficient. The

American Statistician 42 (1): 59-66.


Levenberg, K. 1944. A method for the solution of certain problems in least squares. Quarterly of Applied

Mathematics 2: 164-8.

Levy, A.V. 1995. Solid particle erosion and erosion-corrosion of materials, second edition, Materials

Park, Ohio: ASM International.

Levy, A.V. and Chik, P. 1983. The effects of erodent composition and shape on the erosion of steel. Wear

89 (2): 151-62.

MacKay, D.J. 1992. Bayesian interpolation. Neural computation 4 (3): 415-47.

MacLeod, C. 2013. The Back Propagation Algorithm In An Introduction to Practical Neural Networks

and Genetic Algorithms For Engineers and Scientists , Chap. 3 : J. W. Fernanda

McCasland, M., Barrilleaux, M., Gai, H. et al. 2004. Predicting and mitigating erosion of downhole flow-

control equipment in water-injector completions. Presented at the SPE Annual Technical Conference and

Exhibition.

Marquardt, D.W. 1963. An algorithm for least-squares estimation of nonlinear parameters. Journal of the

Society for Industrial & Applied Mathematics 11 (2): 431-41.

McCulloch, W.S. and Pitts, W. 1943. A logical calculus of the ideas immanent in nervous activity. The

Bulletin of mathematical biophysics 5 (4): 115-33.

McLaury, B.S. 1996. Predicting solid particle erosion resulting from turbulent fluctuations in oilfield

geometries. Ph.D. Thesis, University of Tulsa, Tulsa, Oklahoma (1996)

McLaury, B.S. 1993 A model to predict solid particle erosion in oilfield geometries. M.S.. Thesis,

University of Tulsa, Tulsa, Oklahoma (1993)


McLaury, B. and Shirazi, S. 1999. Generalization of API RP 14E for Erosive Service in Multiphase

Production. Presented at the SPE Annual Technical Conference and Exhibition, Houston, Texas, 3-6

October. SPE-56812-MS http://dx.doi.org/10.2118/56812-MS

McLaury, B. and Shirazi, S. 1998. Predicting sand erosion in chokes for high pressure wells. Presented at

the SPE Annual Technical Conference and Exhibition. New Orleans, Louisiana, 27-30 September. SPE-

49308-MS http://dx.doi.org/10.2118/49308-MS

McLaury, B., Wang, J., Shirazi, S., Shadley, J. and Rybicki, E. 1997. Solid particle erosion in long radius

elbows and straight pipes. Presented at the SPE Annual Technical Conference and Exhibition. San

Antonio, Texas, 5-8 October. SPE-38842-MS http://dx.doi.org/10.2118/38842-MS

Mohaghegh, S. and Ameri, S. 1995. Artificial neural network as a valuable tool for petroleum engineers.

Presented at the SPE Petroleum Computer Conference, Texas, SPE.

Moré, J.J. 1978. The Levenberg-Marquardt algorithm: implementation and theory. In Numerical analysis,

Chap. 105-116: Springer.

Moré, J.J. and Sorensen, D.C. 1983. Computing a trust region step. SIAM Journal on Scientific and

Statistical Computing 4 (3): 553-72.

Oka, Y.I., Okamura, K. and Yoshida, T. 2005. Practical estimation of erosion damage caused by solid

particle impact: Part 1: Effects of impact parameters on a predictive equation. Wear 259 (1): 95-101.

Oka, Y. and Yoshida, T. 2005. Practical estimation of erosion damage caused by solid particle impact:

Part 2: Mechanical properties of materials directly associated with erosion damage. Wear 259 (1): 102-9.

Pandya, D., Dennis, B. and Russell, R. 2014. An Improved Computational Fluid Dynamics (CFD) Model

for Erosion Prediction in Oil and Gas Industry Applications. Presented at the ASME 2014 33rd

International Conference on Ocean, Offshore and Arctic Engineering.


Pearson, K. 1901. LIII. On lines and planes of closest fit to systems of points in space. The London,

Edinburgh, and Dublin Philosophical Magazine and Journal of Science 2 (11): 559-72.

Pourahmadi, F. and Humphrey, J. 1983. Modeling solid-fluid turbulent flows with application to

predicting erosive wear. PhysicoChemical Hydrodynamics 4: 191-219.

Rosenblatt, F. 1958. The perceptron: a probabilistic model for information storage and organization in the

brain. Psychological review 65 (6): 386.

Rumelhart, D.E. and McClelland, J.L. 1986. Parallel distributed processing: explorations in The

microstructure of cognition. Volume 1. Foundations.

Russell, R., Shirazi, D. and Macrae, J. 2004. A new computational fluid dynamics model to predict flow

profiles and erosion rates in downhole completion equipment. Presented at the SPE Annual Technical

Conference and Exhibition. Houston, Texas, 26-29 September. SPE-90734-MS

http://dx.doi.org/10.2118/90734-MS

Russell, R., Barrilleaux, M., Gai, H. et al. 2004. Design, analysis, and full-scale erosion testing of a

downhole flow control device for high rate water injection wells. Presented at the SPE Annual Technical

Conference and Exhibition.

Russell, R. and Marsis, E. 2013. A State-of-the-Art Computational Fluid Dynamics Simulation for

Erosion Rates Prediction in a Bottom Hole Electrical Submersible Pump. Presented at the SPE Heavy Oil

Conference-Canada. Calgary, Alberta, 11-13 June. SPE-165452-MS http://dx.doi.org/10.2118/165452-

MS

Saputelli, L., Malki, H., Canelon, J. and Nikolaou, M. 2002. A critical overview of artificial neural

network applications in the context of continuous oil field optimization. Presented at the SPE Annual

Technical Conference and Exhibition. San Antonio, Texas, 29 September-2 October. SPE-77703-MS

http://dx.doi.org/10.2118/77703-MS
Shahab, M., Bogdan, B. and Samuel, A. 1997. Permeability determination from well log data. SPE

Formation Evaluation 12 (3): 170-4.

Shamshirband, S. 2015. "Performance investigation of micro- and nano-sized particle erosion in a 90°

elbow using an ANFIS model", Powder Technology 284 (2015) 336–343.

Sheldon, G. 1970. Similarities and differences in the erosion behavior of materials. J BASIC ENG TRANS

ASME 92 (3): 619-26

Sivanandam, S., Sumathi, S. and Deepa, S. 2006. Introduction to neural networks using Matlab 6.0: Tata

McGraw-Hill Education.

Smeltzer, C., Gulden, M. and Compton, W. 1970. Mechanisms of Metal Removal by Impacting Dust

Particles. Journal of Basic Engineering 92: 639-52.

Souza, C. 2009. Neural Network Learning by the Levenberg-Marquardt Algorithm with Bayesian

Regularization (part 2).

Suresh, A., Harsha, A. and Ghosh, M. 2009. Solid particle erosion studies on polyphenylene sulfide

composites and prediction on erosion data using artificial neural networks. Wear 266 (1): 184-93.

Tilly, G.P. 1973. A two-stage mechanism of ductile erosion. Wear 23 (1): 87-96. doi: 10.1016/0043-

1648(73)90044-6.

Turenne, S., Chatigny, Y., Simard, D. et al. 1990. The effect of abrasive particle size on the slurry erosion

resistance of particulate-reinforced aluminium alloy. Wear 141 (1): 147-58

Velten, K., Reinicke, R. and Friedrich, K. 2000. Wear volume prediction with artificial neural networks.

Tribology International 33 (10): 731-6.

Vogl, T.P., Mangis, J., Rigler, A., Zink, W. and Alkon, D. 1988. Accelerating the convergence of the

back-propagation method. Biological Cybernetics 59 (4-5): 257-63.


Werbos, P. 1974. Beyond regression: New tools for prediction and analysis in the behavioral sciences. .

Ph.D. thesis, Harvard University, (1974)

Werbos, P.J. 1988. Backpropagation: Past and future. Presented at the IEEE International Conference on

Neural Networks. San Diego, California, 24-27 July. 10.1109/ICNN.1988.23866

Yu, H. and Wilamowski, B. 2012. Neural Network Training with Second Order Algorithms. In Human–

Computer Systems Interaction: Backgrounds and Applications, Chap.2, 463-476: Springer.

Zhang, Y., McLaury, B.S. and Shirazi, S.A. 2009. Improvements of particle near-wall velocity and

erosion predictions using a commercial CFD code. Journal of fluids engineering 131 (3).

Zhang, Z., Barkoula, N., Karger-Kocsis, J. and Friedrich, K. 2003. Artificial neural network predictions

on erosive wear of polymers. Wear 255 (1): 708-13.

Highlights

 Turbulent kinetic energy and shear stress based erosion models are established
 CFD – Artificial neural network hybrid model to predict erosion rate is developed
 Significant performance boost is obtained compared to baseline Baker Hughes model
 Novel CFD – machine learning method for erosion prediction is presented

You might also like