Download as pdf or txt
Download as pdf or txt
You are on page 1of 54

Author's Accepted Manuscript

A Review of Drillstring Vibration Modeling


and Suppression Methods
Ahmad Ghasemloonia, D. Geoff Rideout, Ste-
phen D. Butt

www.elsevier.com/locate/petrol

PII: S0920-4105(15)00179-5
DOI: http://dx.doi.org/10.1016/j.petrol.2015.04.030
Reference: PETROL3039

To appear in: Journal of Petroleum Science and Engineering

Received date: 9 June 2014


Accepted date: 20 April 2015

Cite this article as: Ahmad Ghasemloonia, D. Geoff Rideout, Stephen D. Butt, A
Review of Drillstring Vibration Modeling and Suppression Methods, Journal of
Petroleum Science and Engineering, http://dx.doi.org/10.1016/j.petrol.2015.04.030

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 Review of Drillstring Vibration Modeling and Suppression Methods

Ahmad Ghasemloonia, Corresponding Author

Address: Advanced Drilling Laboratory, Faculty of Engineering, Memorial University, St. John's, NL, Canada, A1B 3X5

Tell: +1(709)8642573, Fax: +1 (709)8648975, E-mail: a.ghasemloonia@mun.ca

D. Geoff Rideout, Associate Professor

Address: Advanced Drilling Laboratory, Faculty of Engineering, Memorial University, St. John's, NL, Canada, A1B 3X5

Tell: +1(709)8643746, Fax: +1 (709)8648975, E-mail: g.rideout@mun.ca

Stephen D. Butt, Professor

Address: Advanced Drilling Laboratory, Faculty of Engineering, Memorial University, St. John's, NL, Canada, A1B 3X5

Tell: +1(709)8648955, Fax: +1 (709)8648975, E-mail: sdbutt@mun.ca

Abstract

Drilling is one of the most costly and risky activities in oil and gas reservoir exploration and field

development. A portion of this high cost is related to unwanted vibrations of the drillstring. With the

advancement of “Measurement While Drilling“ (MWD) tools and their real-time implementation,

vibration models are still a valuable tool for pre-drilling analyses, designing the “Bottomhole Assembly”

(BHA) and sensitivity analyses of the input parameters. In the last 70 years, a wide variety of models has

been developed for design and analysis of drilling structures. Due to the complexity of downhole

interactions and excitations a variety of modeling simplifications and assumptions is typically made. This

paper presents a broad survey of the drillstring vibration modeling literature. The state-of-the-art of

models for predicting axial, torsional and bending vibrations (uncoupled and coupled) , boundary

condition assumptions, equation formulation methods, and applications to vibration mitigation is

reviewed. Moreover, the challenges of drillstring vibration modeling in the presence of modern drilling
techniques such as deviated drilling and use of vibrating downhole tools are discussed. This survey is

intended to organize and summarize the vast amount of literature in the field, and to aid engineers in both

industry and academia in developing, choosing, or critically assessing the limitations of drillstring

vibration models.

Keywords

Drillstring Vibration; Coupled/Uncoupled Modes; Boundary Conditions; Vibration Suppression;

Vibration-Assisted Drilling; Directional Drilling

1. Introduction

Drilling is an essential component of exploration and exploitation of oil and gas reservoirs in petroleum

engineering, and minerals in mining engineering. Failure of drilling facilities, and in particular the

drillstring, has enormous cost implications. The subject of drillstring vibration is an ongoing challenge for

drillers in oil fields and drillstring vibration is assumed as the primary cause of drillstring components'

premature failure, bit failure, deterioration of the well trajectory, excessive bit and stabilizer wear, lower

penetration rate, reduced accuracy of “Measurement While Drilling” (MWD) tools and decreased

efficiency (Bailey et al., 2008). Also, unwanted vibrations of the drillstring dissipate a portion of the

energy that is intended to be delivered to the bit. Unwanted vibrations of the drillstring are said to increase

the drilling cost by 10% (Jardine et al., 1994). Drilling companies are continuously working to utilize

methods and technologies to identify the source of unwanted vibrations and suppress them. Although

MWD tools provide downhole data and help towards real-time adjusting of the drilling parameters to

avoid severe downhole vibrations, their failure due to successive lateral shocks in conjunction with their

high cost has led drilling companies to develop sophisticated drillstring vibration models for “bottom-hole

assembly” (BHA) design and pre-drilling analyses. Moreover, dynamic models are the first essential step

towards developing control strategies for a faster and efficient drilling without premature component

failures.
The complexity of the drillstring dynamic response is due to factors such as formation properties and

heterogeneity, BHA imbalance, misalignment and friction. Such factors are usually unknown or not

measurable in real-time. Mud damping, drillstring-wellbore contact, stochastic bit-rock interaction force

and excitation sources add further complexity to the problem (Spanos et al., 2002). Models to analyze the

vibration pattern of the drillstring can be formulated either in the frequency (Hakimi and Moradi, 2009) or

the time domain (Yigit and Christoforou, 1998; Khulief and Al-Naser, 2005). The most common

application of models is to predict the effects of adjusting the parameters of vibration generator tools, or

the main drilling parameters at the surface such as rotary speed, torque, and WOB (Chin, 1994). Fidelity

of the models depends on the realism of assumptions regarding drillstring geometric configuration,

interacting forces, linear and non-linear excitation sources, damping effects, contact behavior with the

wellbore, and boundary conditions including multiple stabilizers located along the BHA (Clayer et al.,

1990). Despite the challenges in model formulation, and the compromise between model fidelity and

computation speed, models are still considered to be a powerful and emerging method to compare the

sensitivity of different BHAs to vibrations, and to study the propagation of near-bit vibration along the

drillstring to the surface.. Model predictions in conjunction with MWD tool data are considered as reliable

tools to predict vibration patterns or analyze suspected vibration-related failures, and to decrease

unwanted vibrations of the drillstring.

The drillstring is composed of a long, thin-walled interval (drill pipes)—which can be up to several

kilometres—and a heavier, thick-walled bottom section (consisting of drill collars, heavy weight drill

pipe, or both) constrained by the stabilizers inside the wellbore (BHA) with an approximate length

typically up to hundreds of metres. The BHA plays the dominant role in the vibration behavior of the

drillstring (Dareing, 1984a; Yaveri et al., 2010). The top of the drillstring is suspended from the hook of

hoisting system suspended from the derrick. The stabilizers are fined subs placed with the BHA interval at

multiple locations to centralize the drillstring inside the wellbore, to increase the load carrying capacity of

the BHA, and to control well trajectories in deviated wells. The annulus between the drillstring and the
cased or uncased wellbore is filled with circulating drilling fluid which is used to provide well control,

flush and lift bit cuttings, and transmit hydraulic power to the bit. Moreover, the mud plays an important

role in stabilizing the lateral vibrations of the BHA as a nonlinear damping media. The drillstring is under

interaction of several axial forces, such as WOB (to provide the axial bit load), hook load, self weight,

mud hydrostatic effects (both upward and downward), torque and drag loads for non-vertical well

intervals, and excitation forces (e.g. bit-formation interaction, multiple contact loads, and vibration

generator tools) (Aadnoy and Kaarstad, 2006). The normal practise for drill string design is to keep the

pipe section under tension with the transition to axial compression occurring within the stiffer BHA

interval. The length and material properties of the BHA, along with WOB and mud density, are

controllable parameters to keep the BHA under compression (Mitchell and Miska, 2011). The rotary speed

of the drillstring is typically between 50 and 200 rpm.

The vibration of the drillstring can be divided into three primary modes: axial, lateral (also referred to as

transverse, or bending), and torsional modes, with different destructive natures. Bit bouncing, stick-slip

and whirling are extreme examples of coupled vibration dominated by axial, torsional and lateral motions

respectively. Bit-formation interaction, multiple point drillstring-wellbore contacts, mass imbalance, and

vibration generation tools are the main sources of these vibration modes. The axial vibration causes lateral

vibration in the BHA, while severe downhole lateral vibration can cause axial and torsional vibrations that

are detectable at the surface (Christoforou and Yigit, 2001). While lateral vibrations are very severe in

vertical wells compared to torsional vibrations, axial vibrations become very important while

implementing downhole axial generator tools or drilling hard formations (Schlumberger, 2010). A typical

drillstring vibrates in 3 major coupled modes: lateral-axial, lateral-torsional and axial-torsional. Rotary

speed, driving torque, and curvature of the drillstring are causes of the coupling phenomenon.

Axial vibration excited by the bit-formation interaction causes bit bounce, which results in cutting tooth

wear and bearing failure (Li et al., 2007). Bit bounce is the most severe manifestation of axial vibration, in
which the bit loses contact with the hole bottom, usually as a result of resonance in the axial direction.

Axial vibration is most common in vertical wells while drilling hard formations and can at least be

detected at the surface (Chin, 1994). Accelerated bearing and tool wear, seal failure, broken tooth cutters,

failure of the MWD tools and reduction in the ROP are most common catastrophic outcomes of axial

vibration and bit bounce (Ashley et al., 2001).

Among the primary modes of drillstring vibration, the transverse mode is said to be responsible for 75%

of drillstring failures (Kriesles et al., 1999; Sotomayor et al., 1997; Berlioz et al., 1996). Lateral vibration

causes large high-frequency bending moment fluctuations in the BHA, which ends in premature fatigue

failure of the BHA components, wellbore washout, and wear of stabilizers (Zhu and Liu, 2013). Bending

waves are not propagated up to the surface via the drillstring as are torsional and longitudinal waves (this

is due to the difference in the wave speed for different type of modes). As a result, damaging lateral

vibration near the bit can go undetected. Furthermore, transverse vibration has a larger damping value

with respect to the other modes, which is the result of mud damping and the drillstring-wellbore contact

(Chin, 1994). Therefore, there could be severe bending vibrations deep in the hole which the surface

measuring tools do not indicate. Therefore, modeling this mode of vibration, extracting the natural

frequencies and analyzing the dynamic behavior of the BHA are important for failure prevention.

Two types of frequency-induced whirling are possible in the rotating BHA, namely forward whirl and

backward whirl. When the BHA center of gravity is not located along the center of the hole, a centrifugal

force causes the collar to bend. Forward whirl is the rotation of the deflected drill collars around the

borehole axis in the same direction as spin rotation (Leine et al., 2002). Backward whirl is a rolling

motion of the drill collars over the borehole wall in the direction opposite to spin. Whirling modes are

excited due to wellbore contact in low strength formations and as a result, an over gauge hole can be

drilled (Leine et al., 2002). Eccentricity, bent drill collars and high compressive loads at the bit, nonlinear

effects of mud forces, and stabilizer clearance are other sources of whirl excitation. Vandiver et al. (1990)
and Jansen (1991) studied the whirling motion of the drillstring. The heavier collar section is easily

excited by contact with the wellbore in the lower modes. In those modes, the collars are vibrating

transversely, while the pipes do not vibrate and remain approximately motionless. This is due to the axial

load distribution along the drillstring, the collars of which are mainly under compression while the pipes

are under tension. As a result of the tension, the natural frequencies of the pipe section increase, while the

natural frequencies of the collar section are reduced (Thomsen, 2003). The lateral vibration behavior of

the drillstring is strongly influenced by BHA vibration.

The drillstring is a structure with low torsional stiffness, which is easily excited during rotation. The BHA

generates torsional oscillations which are transmitted through the drillstring. Torsional vibration often

occurs in hard formations (Ashley et al., 2001) and is excited by the friction force between the drillstring

and the wellbore (Sananikone et al., 1992), or by the reaction torque between the bit and rock. Torsional

vibration causes irregular downhole rotation, which causes fatigue failure of the drill collar connections,

damage to the bit and slows down the drilling process (Brett, 1992; Elsayed et al., 1997). In general,

torsional vibrations can be detected at the surface by fluctuations in the power needed to maintain a

constant rate of rotation (Ashley et al., 2001; Wu et al., 2012). MWD devices have assisted researchers in

obtaining a better understanding of this type of drillstring vibration and its effects on downhole tools and

drilling performance. Stick-slip is a severe form of torsional vibration of the drillstring. It involves

periodic fluctuation in the bit rotational speed varying almost from zero velocity to more than twice the

surface rotary speed (Chen et al., 2002). . In this state, the BHA can over rotate and builds up reverse

torque, causing the BHA and the bit to rotate backwards. Early studies on stick-slip include

Belokobyl’skii and Prokopov (1982), in which friction induced self-excited stick-slip motions of the

drillstring were studied.

In the last 60 years, drillstring vibration modeling has evolved and been enhanced to precisely predict

downhole vibrations and determine the effect of input parameters on the drillstring vibrations. Static
modeling, basic and enhanced elastodynamic modeling, and numerical modeling (e.g., finite element or

finite difference) are common approaches towards analyzing the dynamic behavior of the drillstring. Static

models were first developed in the 1950s to investigate the stability, load carrying capacity of the

drillstring, reaction loads at the bit, BHA-wellbore side forces, and mud hydrostatic forces. These models

were mostly implemented to design the length and geometry of the entire drillstring, and predict the

deformed shape of the drillstring inside the wellbore (Walker and Friedman, 1977); however, they were

not able to reveal any information about the modal contents of the drillstring, dynamic response of the

BHA, and contact behavior. Ultimately, the shortcoming of these models directed the researchers towards

developing basic dynamic models.

Basic elastodynamic models were first developed in the 1960s to investigate the natural frequencies and

mode shapes of the drillstring. Classical uncoupled, non-rotating lateral, axial, and torsional-beam

vibration equations with simplified boundary conditions, without major excitations (e.g., bit-rock

excitation, contact forces) were developed for the drillstring problem (Dareing, 1984a, b; Chen and

Geradin, 1995). However, these models were not capable of extracting transient and steady-state response

of the drillstring down the hole.

Then, enhanced uncoupled dynamic models were developed to investigate precise critical rotary speeds,

the transient and steady state response of the drillstring, contact behavior and the resulting forces, cutting

forces at the bit, and reaction forces at the surface. The motivation for the enhanced dynamic models was

to adjust the primary working parameters of drilling (combination of WOB, driving torque, rotary speed,

pump pressure and mud characteristics), develop remedial guidelines for severe vibration levels of the

drillstring, and design controllers for in-time suppression actions based on investigation of the transient

and steady state response of the drillstring. Formulating various coupling effects and contact forces,

setting up equations of motion, and implementing approximate solution methods (approximate shape
functions and simplification of boundary conditions) are significant modeling challenges. Vibration

coupling effects were eventually added to the enhanced uncoupled models as discussed in Section 3.

This study is a comprehensive review of the state-of-the-art in drillstring vibration models with different

vibration modes (including coupled modes), excitation sources, modeling approaches, and boundary

condition assumptions. Application of modeling to vibration mitigation and emerging drilling

technologies is also reviewed. Sections 2 and 3 review models with uncoupled and coupled modes,

respectively. Section 4 focuses on two important aspects of drillstring vibration modeling, i.e., boundary

condition assumptions and excitation sources in analytical and numerical models. Section 5 is a review of

modeling methods and equation solution schemes. Section 6 is a review of experimental investigations of

drillstring vibrations. Section 7 is a review of the challenges in vibration modeling to incorporate

emerging drilling technologies such as “Vibration-Assisted Rotary Drilling” and “directional drilling”.

Section 8 is a review of techniques and models applied to mitigate or reduce the unwanted vibrations of

the drillstring, and is followed by the Conclusions section.

2. Uncoupled axial, lateral and torsional vibration models

Since the commencement of drillstring vibration modeling, a variety of models has been developed to

study each of the three possible modes in an uncoupled manner. An uncoupled model can reveal insight

into response of a single mode for a limited range of operating conditions, and has the advantages of

simplicity and high computation speed. Many early models were uncoupled out of necessity, given the

limited computing resources available at the time. In the following section, uncoupled drillstring vibration

models for each type of vibration will be reviewed.

2.1 Axial vibration models

A variety of models has been developed to investigate the axial vibration of the drillstring. Mostly,

uncoupled axial vibration is modeled using the linear partial differential equation governing the undamped

longitudinal oscillation of a bar (Kreisle and Vance, 1970). An early study was conducted by Bailey and
Finnie (1960), in which an undamped partial differential equation for longitudinal bar vibration was used

to derive the axial natural frequencies of the drillstring. The drillstring was assumed as a stepped bar at the

collar-pipe intersection. Also, the effect of different boundary conditions in the axial vibration models of

the drillstring was discussed in their study. Finnie and Bailey (1960) explained experimental methods to

measure the resulting force and displacement at the top of the drillstring due to axial vibrations. Paslay

and Bogy (1963) investigated the intermittent bit tooth contact on the axial vibration of the drillstring and

proved that axial vibrations of the drillstring can be detected at the surface. Dareing and Livesay (1968)

implemented the damped linear axial wave equation to study axial vibrations of the drillstring. One of the

most important early studies on shock sub effects was conducted by Kreisle and Vance (1970). A classical

linear-damped axial wave equation was considered as the governing equation and the Laplace

transformation was implemented to solve the equations. A shock sub with lower stiffness was suggested in

their study as a mean to attenuate the unwanted axial vibrations of the drillstring. Dareing (1984a,1984b)

implemented a damped axial wave equation and studied the effect of the length of the BHA on the axial

resonance of the drillstring. Decoupling the axial resonance from drill bit excitation frequencies was

suggested either with a shock absorber, or by changing the BHA length or rotary speed. Dareing (1985)

investigated the positive effect of axial vibration of the drillstring on the cutting force at the bit and

concluded that BHA length has a great effect on the axial vibration energy transferred to the bit. Skaugen

et al. (1986, 1987) investigated the effect of shock subs on axial vibrations of the drillstring using a free

oscillation equation of a non-distributed mass and considered structural damping of the drillstring.

Dunayevsky et al. (1993) investigated the stability of the drillstring under fluctuating WOB using the

undamped-uncoupled classical axial vibration equation of a bar. Parfitt and Abbasian (1995) implemented

a finite difference method to assess the effect of a shock sub on the axial vibration of the drillstring.

Viscous damping was considered between the elements and a non-linear bit-formation interaction model

was considered for the first time. The effect of process damping, caused by the interference between the

cutter and the previously cut surface, on the axial vibrations of the drillstring and its stability was
investigated by Elsayed et al. (1994). Niedzwecki and Thampi (1988) studied the response of a drillstring

excited by the heave motion of a drill ship. The drillstring was modeled as an assembly of continuous rod

segments of varying cross sections. A linearized form of the classical axial wave equation along with

viscous damping and hydrodynamic added mass was implemented to study the heave response of the

drillstring. Elsayed and Phung (2005) modeled the axial vibrations of the drillstring using discrete mass

segments and springs and the FRF function was used to prove the similarity of the response of the model

to a real drillstring. It was concluded that retaining the first 6 modes is accurate for the axial response

determination. Li et al. (2007) modeled the axial vibrations of the drillstring using the damped classical

axial vibration equation of a bar and investigated the lower boundary condition effects on the accuracy of

the axial response. A bit displacement function, as opposed to a prescribed force, was suggested to be the

most accurate lower boundary condition.

Extreme-depth wells pose a potential axial vibration problem during tripping. The sudden change of

velocity at the top of the drillstring and axial velocity gradient prior to reaching a constant tripping speed

can create a transient axial impulse excitation. Lubinski (1988) was one of the first researchers who

investigated the dynamic phenomena of the drill pipe during tripping motion. A simplified non-coupled

model of the axial vibration of a bar was implemented to investigate the dynamic load of the drillstring

during tripping. Tool joint stiffness and mud damping were also considered in his axial vibration model. It

was concluded that the dynamic load during tripping (in and out) could be greater and lower than the static

weight of the drillstring and could reach the yield stress of the drillstring material. In a series of studies,

Aadnoy et al. (2001; 2010) studied the drillstring-wellbore friction during tripping motion using static

models of the drillstring and proposed rotation of the drillstring to reduce drillstring-wellbore friction. A

quasi-static analysis for friction models during the tripping motion may result in a more realistic model

compared to static modeling.


2.2 Lateral vibration models

Mathematical models have an important role in the investigation of lateral vibrations of drillstrings; both

analytical and finite element models have been developed to study this potentially catastrophic mode of

vibration. For the boundary value problem of transverse motion of the drillstring, a beam element is often

used to derive the mathematical equations. Due to the high slenderness ratio of the drillstring and low

rotational speed, Euler-Bernoulli beam theory is mostly used to model the lateral vibration of drillstrings

(Rao, 2007). The use of Euler-Bernoulli beam theory to model flexure-dominated (long) beams and the

Timoshenko theory for shear-dominated (short) beams was also recommended by Beck and da Silva

(2010).

Lateral vibration models of the drillstring can be grouped into two major categories: single plane and

three-dimensional (two orthogonal lateral directions). Single-plane models are mostly derived in polar

coordinates. Ghasemloonia et al. (2012) derived the equations of motion for two lateral directions and

investigated the effect of driving torque and WOB on the resonance rotary speeds of the drillstring. Jansen

(1991) studied the dual lateral motion between the drillstring and the wellbore at the contact point, using a

lumped segment to model the contact point in polar coordinates. Yigit et al. (1997) studied single mode

lateral vibrations of the drillstring through modal analysis and investigated the effect of mode localization

and the resulting BHA failure. Chen and Geradin (1995) studied the three-dimensional lateral vibration of

a rotating BHA in the presence of a constant WOB with the transfer matrix technique to study the

influence of rotary speed, mud damping and WOB on the resonance rotary speeds and load carrying

capacity (buckling load) of the drillstring. Spanos et al. (1997) implemented the transfer function method

based on modal superposition to investigate the effects of frequency-dependent fluid added mass and

nonlinear wellbore constraints on the lateral response of the BHA. Gulyaev et al. (2006; 2007) derived the

equations of motion of the drillstring in dual lateral modes and studied the buckling capacity and effects of

input parameters on the stability of the drillstring. Vaz and Patel (1995) implemented the Galerkin

technique to study single mode lateral vibrations of the drillstring and extract natural frequencies and
buckling mode shapes in the vertical and horizontal wells. Spanos et al. (2002) investigated lateral

vibrations in the presence of bit-rock interaction force using a stochastic dynamic approach and the finite

element method. In another study, Spanos and Payne (1992) used FEM with a frequency-dependent mass

matrix to study the effect of fluid added mass and Rayleigh damping on the lateral BHA response. Apostal

et al. (1990) implemented FEM to investigate the forced frequency response of the BHA lateral vibrations

and to investigate the effect of damping due to formation friction and mud. A complete discussion of

lateral mode damping and impulsive drillstring-wellbore contact can be found in Section 4. The lateral

vibration models that have been described in this section assume that bending motions are decoupled from

axial and torsional motions. The reader is referred to Section 3 for a review of models in which coupling

effects are considered.

2.3 Torsional vibration models

A variety of models have been used to study torsional vibration of drillstrings. The most prevalent model,

which has been widely used for many years, implements a torsional pendulum. In this model, the BHA is

assumed as a rigid body and the pipes are assumed inertialess compared to the rotary table and the collars

(Lin and Wang, 1991; Jansen 1995; Tucker and Wang, 1999). Modifications have been made to adopt this

model for other investigations; for example, the rotary table has been considered as an added mass (Jansen

and Steen, 1995; Lin and Wang, 1991; Zamanian et al., 2007). In most studies, viscous damping has been

applied between the masses and the borehole wall (Jansen and Steen, 1995; Lin and Wang, 1991). The

mud damping effect can be modeled as a torsional dashpot (Tucker and Wang, 1999). Halsey et al. (1986)

derived a mathematical model of drillstring torsional vibration and studied the influences of damping and

rotary speed on the stick-slip vibration of the drillstring. Lin and Wang (1991) investigated the effect of

dry friction on the torsional vibrations of the drillstring. Baumgart (2000) investigated the effect of mud

damping on the stick-slip vibrations of the drillstring. The effect of bit-rock interaction force on the

torsional vibrations of the drillstring was investigated by Gulyaev et al. (2009), and the effect of bit-rock

interaction on stick-slip motion of the bit by Challamel et al. (2000). Besselink et al. (2011) investigated
stick-slip vibrations with a semi-analytical approach and limit cycle stability analysis. Barton et al. (2010)

implemented a nonlinear torsional model of the drillstring and performed bifurcation analysis to study the

stick-slip phenomenon. Navarro-Lopez and Cortes (2007) developed a lumped parameter model to

investigate the influence of sliding motion on self-excited stick-slip oscillations and bit sticking

phenomena. Hoop bifurcations was used to investigate the range of rotary speeds where undesired

torsional vibrations of the drillstring happen. In a recent study, Arjun Patil et al. (2013) developed an

uncoupled 2 DOF torsional vibration model of the drillstring and conducted a parametric study on the

effect of rotary speed, WOB and drillstring stiffness on the stick-slip motion of the drillstring and also

effect of rock strength on the rate of penetration and verified that the results are inline with the field trials.

More discussion of stick-slip and whirling phenomena can be found in the section on coupled drillstring

vibration models.

3. Coupled vibration models

The change of axial force from tension to compression along the drillstring, the coupling nature of bit-

rock interaction, high static driving torque and the curvature of the drillstring are major causes of coupling

of vibration modes of the drillstring. While each individual mode of drillstring vibration is of great

academic interest, coupled vibration study is very important in practical drilling engineering for higher-

fidelity prediction of the dynamic behavior. Such models may be linear or nonlinear. Coupling between

torsion and bending (Khulief and Al-Naser, 2005; Al-Hiddabi et al., 2003; Yigit and Christoforou, 2000;

Yigit and Christoforou, 1998), axial motion and bending (Ghasemloonia et al., 2013; Yigit and

Christoforou 1996) and axial motion and torsion (Sampaio et al., 2007; Elsayed and Raymond, 2002) have

been extensively studied.

3.1 Coupled axial-torsional vibration models

Sampaio et al. (2007) used nonlinear finite element analysis, considering nonlinear strain displacement

and geometrical stiffening, to analyze axial-torsional vibration coupling of the drillstring. Energy
variational methods were used to generate the equations, assuming both linear and nonlinear strain energy.

The non-linear model reduced to a linear model; however, the response of the drillstring differed after the

first period of stick-slip and the non-linear model was recommended for designing control strategies. Also,

the nonlinear model was more predictive of stick-slip compared to a linear model. Zamanian et al. (2007)

used a discrete model with two degrees of torsional freedom and one degree of axial freedom. Frictional

contact and nonlinear bit-rock interaction were considered and this interaction was said to be the cause of

axial-torsional coupling. The Euler-forward finite difference technique was used to solve the numerical

equations and to study the role of mud damping on the stick-slip phenomenon. Elsayed and Raymond

(2002) investigated the axial-torsional coupling of the drillstring with PDC bits. The measurement of

chatter showed evidence of stick-slip as well as coupling between axial and torsional modes. The

sidebands in the frequency spectrum (frequency modulation) around the torsional natural frequency was

considered as the axial-torsional coupling indicator. Gulyayev et al. (2009) investigated the quasi-static

stability of an elongated drillstring under torque, axial load, rotary inertia force and internal flow of the

mud. The natural frequencies and mode shapes of the elongated drillstring were extracted using their

model. Drillstring fatigue life under coupled axial-torsional vibration was studied by Chi et al. (2006) and

the coupling was attributed to bit-rock interaction. The effect of torsional vibrations on axial vibrations

was studied by Varonov et al. (2007) with a two degree of freedom model. The nonlinearity due to the

cutting action of the bit was considered and the influence of axial and torsional dynamics on the chip

formation process was studied. Poincare maps were sketched to develop regions of the operating

parameters for stable continuous rock cutting. Germay et al. (2009) also developed a coupled axial-

torsional vibration model of the drillstring to investigate the effect of axial vibrations on the stick-slip

phenomenon.

3.2 Coupled bending-torsional vibration models

Coupled torsional-bending vibration of the drillstring was studied by Yigit and Christoforou (1998). Bit-

rock interaction assumptions led to a highly nonlinear set of equations. Newton’s method in polar
coordinates was implemented to solve the governing equations for the lumped mass of the drillstring at the

contact point with the wellbore. Stick-slip for these coupled modes was also investigated in their research.

Al-Hiddabi et al. (2003) applied Newton’s method in polar coordinates to derive the equations for coupled

torsional-bending vibrations. Using their mathematical model, a nonlinear control method to suppress this

vibration was suggested. Leine et al. (2002) investigated the stick-slip whirl interaction using bifurcation

theory. The fluid interaction and contact were considered, and the disappearance of stick-slip with the

appearance of whirl was verified by experimental data from a full scale drilling rig. Melakhessou et al.

(2003) developed a four degree of freedom system model incorporating bending, torsion, whirling motion

and drillstring-wellbore friction; and investigated the effect of initial position of the drillstring and friction

coefficient on the contact behavior of the drillstring. Liao et al. (2011) developed a reduced order coupled

lateral-torsional FEM model at the contact point of the drillstring and wellbore. Based on a qualitative

analysis, an optimum friction coefficient value for stable drillstring behavior at the contact point was

suggested. The effect of mud damping on the lateral motion at the contact point was neglected and the

model was restricted to predicting lateral motion at the contact point. Richard et al. (2007) derived a

coupled lateral-bending model to study stick-slip vibrations and the cause of this phenomenon in drag bit

drilling.

3.3 Coupled axial-lateral vibration models

Hakimi and Moradi (2009) investigated single plane lateral-axial natural frequencies of a single span

BHA. The “Differential Quadrature Method” (DQM) was used to solve the nonlinear set of equations and

to derive the natural frequencies and mode shapes of the BHA. Yigit and Christoforou (1996) investigated

the axial-transverse behavior of a non-rotating BHA. Dunayevsky et al. (1993) modelled coupled axial-

lateral vibrations of the drillstring subject to fluctuating WOB and axially induced bit excitations to

predict stable rotary table speeds. Mahyari et al. (2009) developed a continuous coupled axial-lateral

model to study the influence of the locations of multiple stabilizers on the stability of the drillstring. The

equilibrium position of the system was determined by finding the roots of the first derivative and the sign
of the second derivative of the potential energy of the system. The best location of one, two or three sets

of stabilizers to give stable lateral motion and the largest WOB was investigated. Sahebkar et al. (2011)

implemented the multiple scale perturbation technique to solve coupled-dual lateral-axial vibrations of the

drillstring and study the stability region of the system. The effects of rotary speed, axial compression load,

and imbalance mass on the drillstring response were studied. The effect of nonlinear terms on the

parametric resonance phenomenon was also discussed in their study. Christoforou and Yigit (1997)

developed a coupled axial-lateral model of a rotating drillstring using the Lagrangian approach. The

gyroscopic effect, contact, mud lateral damping and axial excitation of the bit-rock interaction were

considered. Simultaneous parametric resonance and whirling was indicated in the results. Trindade et al.

(2005) implemented a non-linear axial-bending finite element model, considering nonlinear strain

displacements. The self weight of the drillstring yielded a large number of vibration impacts between the

drillstring and wellbore. The Karhunen-Loeve decomposition technique was implemented to derive the

proper number of orthogonal retained modes to precisely reconstruct the drillstring-borehole impacts.

Ghasemloonia et al. (2013) developed a coupled axial-lateral model to investigate the effect of vibration

generator tools on the stability of the drillstring and its contact behavior. The equations were derived using

the Lagrangian approach and the Bypassing PDEs method, and the results were verified using an FEM

model. The effects of torque, non-linear axial stiffening, and spatially varying axial force were included.

3.4 Fully coupled vibration models

One of the early studies on fully coupled drillstring vibrations was conducted by Baumgart (2000), who

derived nonlinear differential equations of the drillstring in three planes to study the effect of mud flow

and mud pump pressure on the vibrations of the drillstring. The model was capable of predicting WOB

and TOB as a result of rock cutting. Bit bounce, drillstring buckling and stick-slip were shown to be

influenced by the mud properties. The transverse vibration results were not discussed in the paper. In

2001, Christoforou and Yigit developed a fully coupled model to design a controller for stick-slip

mitigation. The model accounted for dependence of all three modes, contact, and bit-formation
interaction. It was suggested that a certain level of axial vibration helps mitigation of stick-slip. Khulief

and Al-Naser (2005) developed an FEM model for finite shaft elements with 12 degrees of freedom. The

model accounted for gyroscopic effects, torsional-bending inertia coupling, inertia-axial stiffening

coupling and gravity. Modal transformation was used to obtain a reduced order model. The modal

contents and lateral time response of the drillstring without contact and torsional response at the bit were

extracted in both full order and reduced-order models. Khulief et al. (2008) developed a coupled lateral-

torsional elastodynamic model at the impact point of the drillstring, using a Lagrangian approach in

conjunction with the FEM method. Torsional-bending inertia coupling, axial-bending geometric nonlinear

coupling, gyroscopic effects and gravity were assumed in the model. The wellbore impact was modelled

based on a continuous force-displacement law which represents the impulsive load during short impact

intervals. The material stiffness and damping at the contact location were determined from energy balance

equations and the contact behavior in the presence of stick-slip was studied. The model was verified by a

laboratory test rig by the same authors (Khulief and Al-Sulaiman, 2009). In 2009, Ritto et al. developed a

fully coupled model of the drillstring to study the effect of mud flow. A simplified fluid-structure

interaction model that considered mud flow inside and outside of the drillstring was developed, and all the

modes were coupled via finite strain terms. The bit-rock interaction and the drillstring-wellbore contact

were also included in the model. The results showed that the mud internal and external flow has a great

impact on the vibrations of the drillstring, especially in the lateral mode.

It is apparent that a wide range of model complexities has been used in the prediction of drillstring

dynamic response. The types of vibration to include, and the formulation method, must be carefully

considered based on the simulation goals and computation time constraints of the investigator. The

drillstring is not simply a beam vibrating in air with simple external forcing. High-fidelity dynamic

predictions require appropriate boundary conditions and treatment of mud damping, wellbore contact, and

the complex interactions between the bit and formation.


4. Boundary conditions, contact, bit-rock interaction and mud damping models

Due to the complex dynamics of the drillstring, especially during coupled vibration, closed form solutions

of the governing differential equations is generally not possible. Approximate solution techniques are

required to numerically solve the equations of motion. The proper assumption and simplification of

boundary conditions (BC) for different modes is an important step towards getting a precise approximate

function—and thus a precise solution. In this section the assumptions and simplifications of the boundary

conditions, excitation sources and damping in the drillstring will be discussed.

4.1 Boundary conditions for axial vibration

Axial drillstring boundary conditions (BCs) are treated in a variety of ways in the literature. Fixed at the

top-free at the bottom BC (Mahyari et al., 2009; Yigit and Christoforou, 1996), fixed-fixed BC (Dareing

and Livesay, 1968) and free at top-fixed at the bottom BC (Bailey and Finnie, 1960; Paslay and Bogy,

1963) are commonly used. Jogi et al. (2002) suggested that the free-fixed axial BC does not match well

with major axial frequencies observed in the field, and that a free end assumption gives predictions closer

to field data. Clayer et al. (1990) suggested an equivalent mass-spring-damper at the top and verified that

this model is sufficiently accurate for rig surface modeling. Arrestad and Kyllingstad (1993) also

suggested an equivalent mass spring damper for the top boundary condition, but recommended a nonlinear

coupled axial model to study the role of this boundary condition on the axial vibration of the drillstring.

The methods of measuring the damping, stiffness and concentrated mass values at the top point of the

drillstring can be found in the study by Aarrestad and Kyllingstad (1993). Bit displacement, the magnitude

of which depends on the rock formation properties and the bit type, has also been suggested as an accurate

lower boundary condition in the axial direction. Kreisle and Vance (1970) was the first to apply a

sinusoidal bit displacement as the lower BC in the axial motion. The frequency of displacement was

assumed to be three times the rotary speed of the drillstring for tri-cone bits and the same as the rotary

speed for PDC bits (Christoforou and Yigit, 1997). Macpherson et al. (2001) suggested the same boundary

condition proposed by Kreisle and Vance (1970), with a phase shift relative to the drillstring rotary speed.
Dareing and Livesay (1968) also suggested a constant amplitude sinusoidal function as the bit

displacement. He also indicated that the various types of drill bits (e.g. roller cone and PDC bits) generate

different loading conditions to the bottom end of the drillstring. Due to the rolling of the bit, a multi-lobed

surface is formed on the formation; the number of lobes formed depends on the number of cones on the

bit. This lobed pattern can be defined by a profile with sinusoidal angular variation elevation (Spanos et

al., 1995).

On the other hand, however, in a number of studies the bit in the axial direction was assumed as a free BC,

and an excitation force (bit-rock force) was used. Elsayed et al. (1997) proposed a force excitation at the

bit which depends on the width of the cut and cutting stiffness of the rock. Yigit and Christoforou

(1997,1998,2000) and Dunayevsky et al. (1993) assumed a varying component of the WOB with the same

frequency as the drillstring rotary speed for PDC bits. To decide between force and displacement

boundary conditions at the bit, Li et al. (2007) used a mathematical model of the drillstring and field data

and concluded that the bit displacement BC model agrees more closely with field data. Skaugen et al.

(1987) also recommended that the bit displacement BC is the most appropriate BC for shock sub design.

In summary, three sets of boundary condition were assumed in the literature for axial vibration of the

drillstring: fixed at top-fixed at the bottom; fixed at top-free at the bottom; equivalent mass-spring-damper

at top and a sinusoidal displacement at the bit location, with the last proposed BC being the most realistic

assumption according to multiple sources.

4.2 Boundary conditions for lateral vibration

Simply-supported BCs for lateral vibrations at the stabilizer locations were suggested by Khulief and Al-

Sulaiman (2009), Dareing (1984a), Heisig and Neubert (2000) and Yigit and Christoforou (1996,1998).

Free-free (Bailey et al., 2008), fixed-fixed (Hakimi and Moradi, 2009), deflected springs (Al-Hiddabi et

al., 2003; Zhao and Lian, 2008), mass loaded boundary conditions (Berlioz et al., 1996) were also

considered in lateral vibration models of the drillstring. Field investigation by Jogi et al. (2002) validated
an assumption of simply supported BCs at the location of stabilizers. The top BC in the lateral direction is

suggested to be fixed at the location of the rotary table and free at the bit location (Bailey et al., 2008; Al-

Hiddabi et al., 2003; Gulyaev et al., 2006,2007).

4.3 Boundary condition for torsional vibration

Choosing appropriate boundary conditions is another important step in the modeling of torsional

vibrations of drillstrings. The upper node represents the rotary table and a constant rotary speed has been

considered in some investigations (Dawson et al., 1987; Leine et al., 2002; Richard et al., 2007; Serrarens

et al., 1998; Jansen and Steen, 1995). In some cases, a control relationship between the rotary table torque

and the rotary speed has been considered in order to maintain desired rotary speed (Brett, 1992).

4.4 Drillstring-wellbore contact models

Treatment of the contact behavior of the drillstring and wellbore is an ongoing challenge (Yaveri et al.,

2010; Burgess et al., 1987). If the working conditions are near a resonance state, lateral vibrations are

amplified and impact with the wellbore results. Catastrophic collisions of the BHA with the wellbore lead

to wear of the drillstring, reduction in ROP and reduction of "mean time between failures" (MTBF)

(Placido et al., 2002). MWD tools can also be catastrophically damaged (Reckmann et al., 2010). Mitchell

and Allen (1987) presented case studies of BHA vibration failure and suggested a sophisticated 3D model

capturing successive contact behavior.

Hsu et al. (1965) was the first to model contact behavior using Hertzian contact theory in studying the

lateral behavior of the drillstring. Jansen (1991) used a lumped representation at the contact point of the

drillstring and wellbore, assuming a mud damping force exerted in two orthogonal directions. The effects

of fluid damping and stabilizer clearance on the resonance lateral frequencies were studied. Stability at the

contact point was also studied using the concept of phase planes. Christoforou and Yigit (1997) and

Ghasemloonia et al. (2010, 2014) used the classical Hertzian contact law with the Dirac delta function to

model the contact behavior of the drillstring and wellbore, assuming a constant axial force along the
drillstring. The assumed mode method with a single mode was used and contact was modeled at the

middle of a single-span BHA. Further investigation using multi-span multi-mode assumptions was

suggested. Yigit and Christoforou (1998) modeled the contact behavior in coupled torsional-bending

motion, using the momentum balance method. Their impulse friction model included a compression phase

and restitution phase, with assumed friction and restitution coefficients. The model was capable of

analyzing both rolling and slip-rolling behavior. The nonlinear coupling significantly affected response

and at some resonance rotary speeds there was significant energy transfer between two modes. Hakimi

and Moradi (2009) and Mahyari et al. (2009) modelled the drillstring-wellbore contact as arising from a

series of constant stiffness springs.

Approaches to deal with the impulsive motion of elastodynamic systems with contact, in a finite element

environment, fall into two categories. The first approach is based on assuming a smooth impulsive force

distribution during the impact interval. In this approach the impact force is presented by a force-

displacement law, where the material stiffness is estimated or assumed using approximate energy

relationships (e.g. Khulief et al., 2008). In other words, in this method the contact location could be

modeled by an interface spring (e.g. the penalty algorithm in ABAQUS). The other approach which is

numerically more efficient (Khulief, 2000) is based on an impulse-momentum balance equation, since the

impulsive forces cause an abrupt change in system velocities or momentum (e.g. the kinematic contact

algorithm in ABAQUS) (Wong et al., 2001; ABAQUS, 2012). Melakhessou et al. (2003) modeled only

the contact point of the drillstring in his FEM model. Four independent degrees of freedom were assumed

in their drillstring model (unbalanced rotor within two bearings). Contact was modeled with a Coulomb

friction law and a qualitative sensitivity analysis for the friction coefficient was conducted. Their model

was capable of considering both rolling and sliding motions. The location of the contact was also found in

their model. The initial position of the string in the well was an important parameter in its future behavior.

Khulief et al. (2008) implemented a continuous force-displacement law to model impulsive contact force

in their FEM model. The material stiffness and damping coefficients were determined at the contact zone
according to an energy balance relation. This type of contact modeling was suggested to prevent jump

discontinuities in numerical solution of discontinuous models. Impact was provoked by increasing the

WOB, and time histories at the contact point were qualitatively studied. The FEM model derived by

Khulief et al. (2008) was verified and tuned by a laboratory scale test rig (Khulief and Al-Sulaiman,

2009). The rig was able to excite both stick-slip and lateral contact. The WOB was implemented by a

shaker and the axial, lateral and torsional natural frequencies for different damping media inside the

wellbore were compared. The transient response was not compared between the FEM and the rig data. As

well, Liao et al. (2011) developed a reduced order FEM model at the contact point of the drillstring and

wellbore. Based on a qualitative analysis, an optimum friction coefficient value for the stable drillstring

behavior at the contact point was suggested.

4.5 Mud damping, mud flow-drillstring interaction and drillstring structural damping models

Damping in drillstring vibrations arises primarily from interaction with the drilling fluid, and primarily

affects lateral vibration; however, damping has been included in axial and torsional models. The two most

common types of damping used in axial models include: viscous damping (Kreisle and Vance, 1970;

Paslay and Bogy, 1963; Spanos and Payne, 1992), structural damping (Spanos et al., 1995) and frequency-

dependent damping (Spanos and Payne, 1992).

The mud damping behavior can be considered as Rayleigh damping (a quadratic expression for the energy
dissipation rate), which is proportional to the mass and stiffness of each mode. In the absence of a major
source of dissipation, such as inelastic material or dashpots, Rayleigh damping is an appropriate model
providing a convenient abstraction for damping low-frequency range behavior (mass dependent) and
higher-frequency range behavior (stiffness dependent). The Rayleigh damping depends on two damping
factors, namely ∝ r as the mass proportional damping and β r as the stiffness proportional damping (
Ci =∝i mi + β i ki ) (Apostal et al., 1990). Appropriate mass-proportional damping does not have a great
effect on the stability limit, while the other factor significantly reduces the stability limit .The assumption
of Rayleigh damping for mud has been shown to deliver accurate results, especially near natural
frequencies (Apostal et al., 1990). The mass proportional damping factor introduces damping forces that
are caused by absolute velocities at each node. This phenomenon could model a structure moving through
a viscous fluid (such as drillstring inside mud), in a way such that any point in the model triggers damping
forces (Sampaio et al., 2007). Rayleigh damping can be converted to mode- proportional damping ratio
(Spanos and Payne, 1992; Apostal et al., 1990).
Christoforou and Yigit (1997), Jansen (1991) and Ghasemloonia et al. (2013) modeled the mud damping

as a hydrodynamic drag force (velocity-squared proportional force). Rayleigh damping was implemented

in drillstring vibration models by Sampaio et al. (2007), Spanos and Payne (1992,1997), Apostal et al.

(1990) and Ghasemloonia et al. (2013). Spanos et al. (1997) investigated an equation for damping as a

function of the working frequency and the mud density in an FEM model. He divided the damping matrix

into two dissipative and non-dissipative matrices to account for both Rayleigh damping and gyroscopic

effects. Sampaio et al. (2007) and Ghasemloonia et al. (2013) assumed mass proportional Rayleigh

damping for both axial and torsional modes. Apostal et al. (1990) divided the damping matrix in his FEM

model to structural damping of the drillstring and structural damping due to the mud damping effect on

the BHA motion.

The drillstring-mud interaction effect on drillstring dynamic response is also discussed in the literature.

Paidoussis et al. (2008) investigated the effect of mud flow both inside and outside the drillstring. They

found that the effect of internal and external mud flow is sensitive to the annular space between the

drillstring and wellbore. Zhang and Miska (2005) studied the effects of mud flow on the load carrying

capacity of the drillstring. A two-dimensional model for a pinned-pinned vertical pipe without wellbore

contact was developed. The critical flow rate for pipe buckling was found, and a relationship for the

length of the pipe section and flow rate in the drill pipe buckling analysis was found. The effect of fluid

density on pipe buckling was also studied. However, the role of damping on the stability analysis and

multiple lateral contacts was not investigated. Ritto et al., (2009) analyzed the influence of mud flow on

the natural frequencies and dynamic behavior of the drillstring. They investigated that the axial and

torsional behaviors are not sensitive to the mud flow. They found lateral natural frequencies changed by a

maximum of 0.53 rpm when the fluid flow is considered. When the fluid flow rate was considered in the

dynamic equations, the lateral dynamic response was a bit larger initially. However, steady-state response

was unchanged.
5. Drillstring vibration models: Derivation of the governing equations and

solution methods

Extracting modal contents (Hakimi and Moradi, 2009; Gulyayev et al., 2007, 2009; Khulief and Al-Naser,

2005; Ghasemloonia et al., 2010, 2012), transient and steady state time response (Yigit and

Christophorou, 1996; Ghasemloonia et al., 2013), qualitative contact analysis (Ghasemloonia et al., 2014;

Spanos et al.,1997) and determining the resulting cutting force, WOB and sensitivity analyses of the

drillstring vibration to the input parameters are the main goals of drillstring vibration modeling. From a

design perspective, drillstring dynamic models enable sensitivity analyses for controllable parameters to

determine their optimum values. The following section is a review of the modeling methods and solution

schemes of drillstring vibrations models.

5.1 Derivation of the governing equations

In order to investigate the dynamic behavior of the drillstring, either in the frequency domain or the time

domain, it is essential to derive the equation of motion of the drillstring. The Newtonian approach and the

energy variational approach (Hamilton's principle with the variational approach) were extensively used by

researchers in the field. Ghasemloonia et al. (2010), Yigit and Christoforou (1998, 2000), Gulyaev et al.

(2006), and Hakimi and Moradi (2009) implemented the force-balance equation concept (Newtonian

modeling) and derived the equations of motion of the drillstring. An energy variational approach with

Hamilton's principle was applied by Mahyari et al. (2009), Melakhessou et al. (2003), Sahebkar et al.

(2011), Sampaio et al. (2007), Heisig and Neubert (2000), Christopherrou and Yigit (1997) and Khulief

and Al-Sulaiman (2009). Ghasemloonia et al., (2013, 2014) derived the Lagrangian of the drillstring

motion in the coupled axial-two orthogonal lateral directions and equations of motions were derived using

the “Bypassing PDE's” method, which is based on combining the expanded Galerkin's technique with the

Lagrange's equation, instead of the conventional Hamiltonian approach for the continuous systems. The

energy variational approach is preferred to the Newtonian approach for complicated systems having

multiple degrees of freedom, systems having complex coordinate systems and systems having coupled
motion in certain directions. This fact is due to the scalar measuring of energy variables compared to the

vector form of the forces in the Newtonian approach. For problems with a rotary coordinate system or for

nonlinear systems the energy method is also preferred.

The difficulties and limitations of analytical models to model complex boundary conditions and forces,

and the need to reconfigure such models for new drillstring geometries, coupled with the development of

fast processing computers, have attracted investigators to the use of recognized powerful numerical

methods, such as finite element method (FEM), finite difference method (FDM), and differential

quadrature method (DQM). One of the early attempts at FEM analysis of the drillstring was conducted by

Millheim et al. (1978), where the drillstring was modeled as a straight beam with beam and gap elements

using an FEM package called MARC-CDC. The nodal displacement was derived using the variational

principle to study drillstring deflection and bit forces. Straight and curved beam elements were compared

in their model and the results were verified by field data. Transient response of the rotating BHA,

assuming bit-rock interaction force, was modeled by Baird et al. (1985) using the FEM technique. Apostal

et al. (1990) studied the forced frequency response of the non-rotating BHA with a 3D finite element

model. Burgess et al. (1987) modeled the lateral vibration of the drillstring by FEM, considering only the

length of the drillstring that was not lying along the wellbore. A static nonlinear analysis was performed to

predict that length. Uncoupled transverse-torsional behavior of the drillstring was modeled by Axisa and

Antunes (1992) with FEM. The drillstring was modeled as a straight rod and a sensitivity analysis for the

fluid elastic effect was conducted. The nonlinear gravitational axial stiffening effect was not considered in

their model. Parametric instability of the rotating drillstring was studied by Berlioz et al. (1996), where

the FEM equations were derived for a 6 DOF shaft element using rotor dynamics equations. Axial

stiffening was neglected in their model. The results were verified by test rig data. Spanos et al. (1997)

modeled the BHA assuming the added fluid mass with a frequency dependent mass matrix. The excitation

forces were modeled as monochromatic functions of time and nonlinear contact force at the stabilizers

location was assumed. The frequency response model was suggested as a powerful tool to understand the
complex behavior of the BHA. In another work by Spanos et al. (2002) a transfer function representation

for the BHA based on modal superposition was derived. Lateral vibrations of the drillstring were studied,

while the lateral displacement of the drill bit was defined as an equivalent linear system and predicted

subsequently by Monte Carlo simulation. Model uncertainties of the bit-rock nonlinear interaction were

studied by Ritto et al. (2009), using a non-parametric probabilistic approach. They found that the

uncertainties in the bit-rock interaction model play an important role in the coupling between the axial and

torsional responses. Khulief and Al-Naser (2005) used Lagrange’s equation to derive an FEM model of a

rotating drillstring. Shaft elements with 12 DOF were used in their model, capturing torsional-bending

inertia coupling and gyroscopic effects, while the contact with the wellbore was not considered. The

reduced order model, derived with the modal transformation, was compared with the full order model.

Time response analysis of the BHA and transverse natural frequencies were produced. Other numerical

techniques, such as “Differential Quadrature Method” (DQM) (Hakimi and Moradi, 2009), “Finite

Difference Method” (FDM) (Chin, 1994), and transfer matrix technique (Chen and Geradin, 1995) were

also used to study the vibration behavior of the drillstring.

5.2 Solutions of the governing differential equations

Using either Newtonian or energy approaches, one will arrive at an “analytical” formulation (a set of

coupled PDE’s) of the boundary value problem. Applying a proper solution method is the next step to

extract the dynamic time response of the drillstring vibration. Due to the coupling effect and nonlinearity

of the drillstring vibration problem, closed form solution of the resulting PDEs is not feasible.

Approximate methods must be applied to the set of resulting PDE’s to convert them into a set of ODEs.

This idea is based on integrating the developed equations over the space domain in “Eigenvalue

Problems” (EVP's), and assumption of the approximate space domain functions (mode shapes) to

eliminate space variables, while considering the orthogonality of the mode shapes. Due to the complexity

of the geometric configuration and boundary conditions, deriving the exact mode shapes of the drillstring

problem is not feasible and approximate shape functions, e.g. comparison functions (Ghasemloonia et al.,
2010) and admissible functions (Yigit and Christoforou, 1996), are instead implemented. Approximate

solution methods have been widely used for the nonlinear coupled vibration analysis of the drillstring.

Assumed modes method was implemented by Yigit and Christoforou (1996), Christoforou and Yigit

(1997) and Mahyari et al. (2009). The Galerkin method was applied by Ghasemloonia et al. (2010,2013)

and Vaz and Patel (1995). Other techniques, such as Laplace transform (Kreisle and Vance, 1970), DQM

(Hakimi and Moradi, 2009) and transfer matrix method (Chen and Geradin, 1995) were also applied as

approximate solution techniques in the vibration analysis of the drillstring.

The number of modes which should be retained in the approximation methods depends on the upper

bound for the frequency of interest and the expected accuracy of the results. Most of the studies in

drillstring vibration assumed a single mode approximation (Yigit and Christoforou, 1996), while some

others conducted a multi-mode analysis of the problem by considering higher number of retained modes

in the approximation methods (Ghasemloonia et al., 2013, 2014). The mass participation factor, the

effective mass and the total modal effective mass are three factors which can be extracted from finite

element models to indicate modes which are contributing more to the motion in a specific direction

(Ghasemloonia et al., 2013). The Karhunen-Loeve decomposition technique was implemented by Al-

Hiddabi et al. (2003) to develop a reduced order FEM model of axial-lateral vibrations of the drillstring

from which retention of 15 orthogonal modes for a precise time response of the drillstring was

recommended.

Once one of the above approximation methods is implemented to give coupled nonlinear PDE’s, and after

integration over the drillstring length and use of mode orthogonality, the result is a set of coupled ordinary

time differential equations. The initial conditions are then required to extract the resonant frequencies and

external forcing to excite the system of ODEs and derive motion time histories. Any compatible set of

initial conditions will suffice, as natural frequencies and steady state responses are not initial condition

dependent. Several solution methods, such as Fehlberg fourth-fifth order adaptive Runge-Kutta method
(Ghasemloonia et al., 2012), Runge-Kutta method (Ghasemloonia et al., 2013), finite difference method

(Chin, 1994), perturbation method (Sahebkar et al., 2011) and Euler-forward finite difference technique

(Zamanian et al., 2007) have been applied to solve the resulting drillstring model algebraic equations. The

great advantage of adaptive solvers compared to fixed-step solvers is the dynamic time step reduction

strategy. For stiff ODEs, it is suggested to implement fixed step solvers with very small time increments

to avoid jump discontinuities.

6. Experimental vibration analysis of the drillstring

Uncertainties and difficulties of analytical and numerical modeling of drillstring vibrations have motivated

researchers in the field to develop experimental and laboratory scale test rigs. Test rigs and field

measurements have been used extensively to validate and tune models, and to better represent the

complicated excitation sources such as bit-rock interaction forces and drillstring-borehole friction

coefficients.

Khulief and Al-Sulaiman (2009) developed a test rig which was capable of varying drilling fluid

interaction and WOB to excite borehole contact and stick-slip. The test rig measurements were used to

find the fluid damping effects and tune a FEM model. Berlioz et al. (1996) developed a test rig to study

the lateral instabilities of the drillstring at the contact points and tune their FEM model. Elsayed and

Raymond (2002), Elsayed and Phung (2005), Elsayed and Aissi (2006) implemented the “Hard Rock

Drilling Facility” (HRDF) in Sandia National Laboratories to conduct a series of experiments on the effect

of PDC bits on the vibration behavior of the drillstring, shock absorber analysis, bit force and

displacement analyses. Liao et al. (2011) developed a test rig to augment their developed 4 and 5 DOF

system based on qualitative contact analysis and stick-slip motions and the trajectory of the drillstring

inside the wellbore. Melakhessou et al. (2003) developed a test rig equipped with two opto-lineic devices

(for measuring the lateral displacement of the drillstring) to validate the whirling motion and tune the

drillstring-wellbore friction coefficient in their coupled lateral-torsional analytical model. Experimental


drillstring vibration analysis studies due to scaling issues and difficulties in simulating the real downhole

vibrations are of limited interest and field investigations are preferred to experimental studies. However,

field experiments are not widely used for cross validations of drillstring vibration models, except for

resonant frequency and vibration levels comparisons. Wolf et al. (1985) compared the uncoupled natural

frequencies of axial and lateral modes with the field data and concluded that predicated axial natural

frequencies are lower than those measured in field. Jogi et al. (2002) verified the natural frequency

analysis of several modeling packages with field results. They proved that the simplification of boundary

conditions in mathematical models agrees well with the field results for almost all modes, especially for

pinned-pinned boundary conditions in the lateral mode. Macpherson et al. (1993) compared resonant

frequencies and axial and torsional wave velocities in the drillstring between an analytical model and field

tests in Oklahoma. It was concluded that axial bit displacement in tricone bits is slightly less than three

cycles per bit revolution assumed in analytical models. Bailey et al. (2008) validated a dynamic model

(VybsTM) of the BHA developed using a lumped parameter, frequency domain approach with eight field

tests. It was concluded that the model derived vibration indices are highly correlated with MSE values

from field tests and the developed model (despite its simplicity) is potential to predict downhole vibration

levels. In another study, Dykstra et al. (1994) conducted laboratory and field experiments to evaluate anti-

whirl and PDC bit effects on the dynamics of the drill bit and drillstring and concluded that anti-whirl bits

vibrate an order of magnitude less than PDC bits. In field trials it was concluded that the location of the

stabilizers plays an important role on the bit vibration. Based on the experimental and field tests,

operational guidelines and drillstring configurations were suggested to reduce downhole vibrations. Arjun

Patil et al. (2013) published a very comprehensive review of the drillstring torsional vibration and

addressed experimental studies performed in laboratories to reproduce modes of vibrations in the

experimental setups.
7. Emerging drilling technologies and vibration modeling challenges

As companies strive to extract resources from more difficult formations, and with lower environmental impact at

the surface, the need for more advanced and efficient drilling methods is increasing. Directional drilling is one

practice that has been widely implemented in recent years, predominantly for “Extended Reach Drilling” (ERD) and

horizontal drilling, with both having long, high-inclination well intervals that are prone to high levels of torque and

drag resulting in reduced weight and torque applied at the bit. Another emerging drilling technology is vibrating

drilling tools that can increase drilling ROP by one of several mechanisms. However, these tools act as a source of

active vibration in the drillstring or BHA which, under specific conditions, may negatively interact with the

drillstring and cause increased shock and damage to drillstrings, drilling tools, and other BHA components. This

section will detail the vibration modelling challenges for these drilling scenarios.

7.1 Vibrating drilling tools

Vibrating drilling tools impart oscillating force vibrations in the tool by various types of hydraulic or

mechanical mechanisms. One class of these vibrating tools are “agitators” which used to mitigate the

effects of torque and drag as described in the preceding section. These agitators use an axially compliant

element, such as a shock sub, in-line with the force generator to produce high frequency axial

displacement oscillations (10 to 30 Hz) close to the bit (Barton et al., 2011). This imposed vibration

significantly reduces the drillstring-wellbore friction thus reducing torque and drag friction and resulting

in improved weight transfer, improved steering, increased power at the bit, and consequently a higher

ROP (Al Ali et al., 2011). Newman et al. (2009) presented a theoretical torque-drag model to verify the

experimental results of the friction reduction through the use of vibration generator tools, while

implementing the National Oilwell “CT AG-itatorTM“ tool. The use of this tool increased the depth of

penetration by approximately 1000 feet and the model and field test results were in agreement in

predicting the amount of friction forces eliminated by the tool. Their positive effects on ROP

notwithstanding, these agitator tools act as a source of excitation for the entire drillstring and as a result,

unintended vibration-related failures of drillstring components can result. Al Ali et al. (2011) investigated
axial oscillation generator (AGT) tools in both vertical and horizontal drilling in several field tests. In

vertical drilling, a 60% increase in ROP, with 63% less required WOB, extended bit life and less stick-slip

were reported. However, increase in vibration levels of the drillstring was reported in all field trials.

Another class of axially vibrating tools is used to apply vibration at the bit to enhance penetration

mechanisms and thus ROP. Manko et al. (2003) and Kolle (2004) separately developed and evaluated

hydrovibrator tools which applied short duration force impacts on the bit and, although ROP was

enhanced under atmospheric drilling conditions, there was no significant improvement reported under

increased well bore pressures at depth. More recently, Li et al. (2010), Khorshidian et al. (2012) and

Babapour et al. (2014) showed from numerical and laboratory studies that enhanced ROP can occur under

pressurized borehole conditions when axial displacement is applied at the bit, however, this has yet to be

evaluated under field drilling conditions. Ghasemloonia et al. (2013) developed an analytical and FEM

model of the drillstring in the presence of these axial vibration generators to predict and evaluate potential

axial and lateral instabilities of the drillstring. Time histories and phase plane plots for several points on

the BHA were generated and severity of the drillstring-wellbore contacts was investigated. The most

severe lateral vibration occurred in the uppermost of the three BHA spans and axial motion was most

severe near the bit.

7.2 Directional drilling and drillstring vibration modeling

Directional drilling is used to reach drilling targets where vertical access is difficult, to hit multiple

subsurface targets, for ERD to reach targets at significant horizontal departure from the drill rig

(particularly for offshore drilling), and for horizontal well bores that have increased exposure in the

reservoirs. In 1970, mud motors and bent subs were implemented to directional drilling to facilitate the

change of direction of the well. However, directional drilling with a downhole motor requires occasionally

stopping rotation of the drill pipe and sliding the pipe through the wellbore as the motor cuts a curved

path. Sliding motion is difficult in certain formations. Significant friction can build up between the
drillstring and the wellbore, which causes poor weight transfer and lower WOB and TOB, resulting in a

low ROP, and thus increased drilling costs. In recent years, rotary steerable drilling tools have facilitated

steering while maintaining drillstring rotation; however, torque and drag is only partially reduced and

problems with reduced weight and torque transfer to the bit remain. In deviated wellbores, modeling the

contact between the low side of the drillstring and the wellbore (friction modeling), and implementing

torque and drag forces in analytical or numerical models are challenging (Hu et al., 2012). A large number

of studies were undertaken since 1987 to develop models capturing these complex phenomena in vibration

modeling of inclined drillstrings.

Torque and drag are caused by tight-hole conditions, differential sticking, cuttings build-up, and most

importantly, sliding wellbore friction (Mirhaj et al., 2011; Wang and Yuan, 2012). In most prior torque

and drag modeling studies, the drillstring was assumed as a string (cable) and normal forces due to the

bending moment were neglected. In all of these models, the normal contact force and the mud upward

pressure were assumed and the coefficient of friction between the contact surfaces was found based on

Coulomb's friction model (Sheppard et al., 1987). The model developed by Sheppard et al. (1987) has

been verified through several field trials (Liu and Samuel, 2009). Later on, hydrodynamic viscous drag

force in wellbore drag modeling was investigated by Meidla and Wojtanowicz (1988). The effect of mud

characteristics on the friction coefficient was investigated by He et al. (1995). The torque and drag model

developed by Aarrestad and Blikra (1994) was implemented in a numerical model by Payne and Abbasian

(1997) to study the buckling of the drillstring and wellbore trajectory. In 1988, Ho applied the stiff beam

model instead of the string formulation for modeling torque and drag. Aadnoy et al. (2006) derived

equations for different sections of the wellbore profile, including straight inclined sections and drop-off

and built-up sections and verified his model in a field test (Adanoy and Djurhuus, 2008). Schamp et al.

(2006) suggested methods to reduce the frictional torque produced between the drillstring stabilizers and

the wellbore. Mitchell and Samuel (2009) applied the torque and drag model established by Kaarstad and

Aadnoy (2009) for beam-like drillstrings. Fazaelizadeh et al. (2010) developed a 3-D analytical model for
directional wellbore friction for different drilling conditions, such as rotating and tripping, and verified the

results by a field test. The torque and drag models by Aadnoy et al. (2006), Aadnoy and Djurhuus 2008)

and Fazaelizadeh et al. (2010) are static torque and drag models, useful for pulling pipe out, lowering it

down, or drilling without vibration. They are of primary use in well planning, as opposed to capturing

dynamic nonlinear friction effects in vibrating drillstrings.

8. Application of dynamic models to mitigation of drillstring vibrations

Unwanted vibrations of the drillstring can diminish the life of the drillstring and cause wellbore washout

and premature failure of the drillstring components. As a result of vibrations, the rate of penetration and

bit longevity will be reduced. Vibration isolation methods for drillstring axial lateral and torsional

vibrations are based primarily on three strategies: proper design of drillstring configuration (BHA length

and stabilizer locations) to stay far away from the resonance state (Mahyari et al., 2009; Bailey et al.,2008;

Jogi and Macpherson, 2002), passive isolators such as shock subs (Elsayed and Aissi, 2006), active

controllers (Zamanian et al., 2007; Jansen and Steen, 1995) and real-time drilling input parameters

optimization based on MWD data (Nicholson, 1994; Finger et al., 2003; Wise et al., 2003; Hutchinson and

Dubinsky, 1995; Macpherson et al., 1993).

Vibration analysis at the design step of the BHA was proposed by Dareing (1984b) and Bailey et al.

(2008) to avoid resonance states in the drillstring through the adjustment of the BHA length. From a

practical drilling standpoint, the rotational speed should be adjusted so that it does not correspond to one

of the natural frequencies (Gulyaev et al., 2007). Certain combination of the WOB and rotary speed can be

used to avoid or reduce vibrations for any configuration of the BHA. In another study, Gulyaev et al.

(2006) and Gulyaev and Borshch (2011) derived a mathematical model to describe the critical quasi-static

equilibrium of the rotating drillstring. The critical frequencies of this coupled mode were extracted and it

was stated that the rotating speed of the drillstring should not be located in the resonance range. Mahyari

et al. (2009) studied the effect of stabilizer locations on the stability of the drillstring and proposed
stabilizer locations for a set of four stabilizers on a specific size BHA. Jogi et al. (2002) used static

buckling analysis to predict wellbore touch points and span length for estimating lateral natural

frequencies of a BHA. Bailey et al. (2008) redesigned a BHA based on the results of a frequency domain

BHA vibration model for stick-slip reduction and field investigation of both BHAs showed that the model

effectively predicted stick-slip in the original BHA and its reduction in the redesigned BHA. Fear et al.

(1997) studied the effect of a near-bit stabilizer on the stick-slip vibration of the drillstring and

investigated that lateral movement of the bit is another reason for stick-slip besides bit-rock reaction

torque. Increasing the BHA stiffness was also proposed as a method to reduce the risk of stick-slip

(Pastusek et al., 2005; Davis et al., 2012) and was verified by Jaggi et al. (2007). The studies by

Ghasemloonia et al. (2012), Chen and Geradin (1995), Hakimi and Moradi (2009) and Khulief and Al-

Naser (2005) predict resonance rotary speeds of the drillstring complicated BHA’s in different modes, to

aid in vibration isolation through avoidance of resonance.

Passive control through the use of shock subs is an efficient method of decupling the source of vibration

from the rest of the drillstring. Shock subs are installed above the bit and below the drill collars, and lower

the natural frequency of the drillstring. Shock subs reduce the displacement or force amplitude transmitted

to the BHA through frequency detuning and phase shift. Elsayed and Aissi (2006) investigated the effect

of shock sub-parameters on vibration isolation in hard rock drilling through an experimental setup, and an

optimum range for the stiffness and damping values of the shock sub to reduce the bit bouncing

phenomena was found. Warren et al. (1998) investigated the role of shock subs on the reduction of axial

vibrations, and suggested to install the shock subs with softer springs below the collars for a more efficient

isolation. A reduction in lateral vibration levels with the use of shock subs was also reported. Skaugen et

al. (1987) proposed a simple linear axial model for designing shock sub parameters and recommended a

nonlinear model for a precise design of shock subs. Moreover, it was concluded that decreasing the

stiffness value of a shock sub to an optimum point increases its efficiency in reducing axial vibration

amplitudes above the shock sub. Kreisle and Vance (1970) investigated the effect of shock subs on the
axial vibration of the drillstring through an uncoupled axial model and the Laplace solution scheme, and it

was proven that the efficiency of the shock sub is due to the change in the phase angle rather than a

change in natural frequency of the system, and recommended a softer spring to increase efficiency.

The potential of active controllers to suppress harmful oscillations of the drillstring in different modes has

been extensively studied. Active control work has focused on stick-slip (Christoforou and Yigit, 2001;

Zamanian et al., 2007; Kriesels et al., 1999), torsional-bending (Al-Hiddabi et al., 2003; Yigit and

Christoforou, 2000; Navarro-Lopez and Cortes, 2007), axial-bending (Trindae et al., 2005), uncoupled

axial (Elsayed et al., 2006) and uncoupled bending vibrations (Yigit and Christoforou, 1997). One of the

early attempts to actively control drillstring vibration was by Halsey et al. (1988), where torque and speed

were controlled for a smooth drillstring motion. Al-Hiddabi et al. (2003) stated that linear controllers are

not effective in suppressing the vibrations of the drillstring and suggested a nonlinear inversion controller

for damping the lateral and torsional vibrations of the drillstring (the controller was more effective in

damping torsional vibrations than lateral vibrations). Yigit and Christoforou (1999) developed a linear

quadratic regular controller based on a linearized model to control stick-slip vibrations. Use of the

controller reduced the occurrence of stick-slip, while exciting some lateral vibration modes for certain

operational parameters. Christoforou and Yigit (2001) designed an optimal state feedback control to

mitigate stick-slip in his developed fully coupled model. He also suggested that axial vibration control to

some extent has positive effects on reducing the chance of stick-slip. Zamanian et al. (2007) developed an

active damping controller which prevents stick-slip vibrations and proposed higher mud damping ratios to

avoid stick-slip.

The real time controllers have practical limitations, as the states cannot be measured in a real-time

manner. One of other practically accepted works in the field of active vibration control of the drillstring

was conducted by Kriesels et al. (1999), where the “Soft Torque Rotary System” (STRS) was applied to

control stick-slip vibrations of the drillstring. STRS actively controls top drive speed to absorb torque
fluctuations based on a transfer function of a drillstring represented by a low-order lumped segment

model.

One of the best options for drillstring vibration isolation is to monitor the controllable parameters during

the drilling process through MWD tools. When encountering high vibration levels, the driller may adjust

these parameters to reduce the vibration based on the provided remedial guidelines (Nicholson, 1994). The

benefits of real time measurement of the drilling parameters and real-time adjusting of the input

parameters are discussed in the study by Shuttleworth et al. (1998). One of the recent studies in this field

is the SCADA system which adjusts the WOB and RPM of the system based on the real-time data

provided by the rig controller (Chapman et al., 2012). Arjun Patil et al. (2013) also conducted a

comparative review of drillstring torsional models in laboratory setups and their implementation in

vibration suppression of the drillstring.

9. Conclusions

Premature failure of drill rig components, and in particular the drillstring, adds to drilling costs. Unwanted

vibration waves along the drillstring are the main cause of drillstring failures. Anticipating problematic

vibration patterns and designing mitigation techniques requires vibration modeling of the drillstring. Over

the years, drillstring vibration models have evolved from basic static models to sophisticated dynamic

models to predict drillstring motions down the hole. Given the complexity of the drillstring and its

interactions with its surroundings, all models are limited by certain assumptions and simplifications. More

enhanced and realistic models need to be developed to more effectively predict the downhole vibrations

for real field applications. Also, with the development of new drilling techniques such as directional and

vibration-assisted rotary drilling, new challenges are appearing in vibration modeling. In this paper, the

state-of-the-art of drillstring vibration models and their evolution has been presented. The assumptions

made for boundary conditions, excitation and damping sources have been reviewed, along with modeling

methods and solutions schemes. Challenges related to drillstring vibration models incorporating emerging
drilling technologies were discussed. Finally, vibration mitigation methods developed with the aid of

models were reviewed.

Acknowledgment

This research was conducted at the Advanced Drilling Technology Laboratory at Memorial University of

Newfoundland and was funded by Husky Energy, Suncor Energy, The Atlantic Canada Opportunities

Agency and The Research and Development Corporation of Newfoundland and Labrador under AIF

Contract No. 781-2636-1920044.

References

Aadnoy, B. S., Andersen, K, 2001. Design of Oil Wells Using Analytical Friction Models. Journal of

Petroleum Science and Engineering. 32, 53-71.

Aadnoy, B. S., Kaarstad, E., 2006. Theory and application of buoyancy in wells. IADC/SPE#101795,

IADC/SPE Asia Pacific Drilling Technology Conference and Exhibition, Bangkok, Thailand.

Aadnoy, B. S., Fabiri, V. T., Djurhuus, J., 2006. Construction of ultralong wells using a catenary well

profile. SPE/IADC#98890, SPE/IADC Drilling Conference, Miami, Florida.

Aadnoy, B. S., Djurhuus, J., 2008. Theory and application of a new generalized model for torque and

drag. SPE/IADC#114684, SPE/IADC Asia Pacific Drilling Technology Conference and Exhibition,

Jakarta, Indonesia.

Aadnoy, B. S., Fazaelizadeh, M., Hareland, G., 2010. A 3D analytical model for wellbore friction. Journal

of Canadian Petroleum Technology. 49(10), 25-36.

Aarrestad, T. V., Kyllingstad, A., 1993. Rig suspension measurements and theoretical models and the

effect on drillstring vibrations. SPE Drilling and Completion. 8(3), 201-206.


Aarrestad, T. V., Blikra, H., 1994. Torque and drag-Two factors in extended-reach drilling. Journal of

Petroleum Technology. 46(9), 800-803.

ABAQUS Theory Manual (version 6.10), 2012, Dassault Systèmes Simulia Corp., RI, USA.

Al Ali, A., Barton, S., Mohanna, A., 2011. Unique axial oscillation tool enhances performance of

directional tools in extended reach applications. SPE Brasil Offshore, Macaé, Brazil.

Al-Hiddabi, S. A., Samanta, B., Seibi, A., 2003. Nonlinear control of torsional and bending vibrations of

oil well drillstrings. Journal of Sound and Vibration. 265(2), 401-415.

Apostal, M. C., Haduch, G. A., Williams, J. B., 1990. A Study to determine the effect of damping on finite

element based forced frequency response models for nottomhole assembly vibration analysis. SPE#20458,

SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana.

Arjun Patil, P, Teodoriu, C., 2013. Model development of torsional drillstring and investigating

parametrically the stick-slips influencing factors. Journal of Energy Resources Technology. 135, 013103-

1- 013103-7.

Arjun Patil, P, Teodoriu, C., 2013. A comparative review of modeling and controlling torsional vibrations

and experimentation using laboratory setups. Journal of Petroleum Science and Engineering. 112, 227-

238.

Ashley, D. K., McNary, X. M., Tomlinson, J. C., 2001. Extending BHA life with multi-axis vibration

measurements. SPE/IADC# 67696, IADC/SPE Drilling Conference, Amsterdam, Netherlands.

Axisa, F., Antunes, J., 1992. Flexural vibration of rotors immersed in dense fluids: Part I- theory. Journal

of Fluids and Structures. 6(1), 3-21.


Babapour, S., Butt, S. D., 2014. Investigation of enhancing drill cuttings cleaning and penetration rate

using cavitating pressure pulses. 48th US Rock Mechanics / Geomechanics Symposium, Minneapolis,

MN, USA.

Bailey, J. J., Finnie, I., 1960. An analytical study of drillstring vibration. ASME Journal of Engineering

Industry. 82(2), 122 – 128.

Bailey, J. R., Biedigner, E., Sundararaman, S., Carson, A. D., Elks, W. C., Dupriest, F.E., 2008.

Development and application of BHA vibrations model. International Petroleum Technology Conference,

Kuala Lumpur, Malaysia.

Baird, J. A., Caskey, B. C., Wormley, D.N., Stone, C. M., 1985. GEODYN2: A bottomhole assembly

geological formation dynamic interaction computer program. SPE#14328, SPE Annual Technical

Conference and Exhibition, Las Vegas, Nevada.

Barton, D., Krauskopf, B., Wilson, R. E., 2010. Nonlinear dynamics of torsional waves in a drill-string

model with spatial extent. Journal of Vibration and Control. 16, 1049-1065.

Barton, S., Baez, F., Al Ali, A., 2011. Drilling performance improvement in gas shale plays using a novel

drilling agitator device. SPE#144416, SPE North American Unconventional Gas Conference and

Exhibition, Woodlands, Texas.

Baumgart, A., 2000. Stick-slip and bit-bounce of deep-hole drillstrings. ASME Journal of Energy

Resources Technology. 122(2), 78-82.

Beck, A. T., da Silva Jr., C. R. A., 2010. Timoshenko versus euler beam theory: pitfalls of a deterministic

approach. Structural Safety. 33(1), 19-25.

Belokobyl’skii, S.V., Prokopov, V. K., 1982. Friction induced self excited vibration of drill rig with

exponential drag law. Soviet Applied Mechanics. 18(12), 98-101.


Berlioz, A., Der Hagopian, J., Draoui, E., 1996. Dynamic behavior of a drillstring: Experimental

investigation of lateral instabilities. Journal of vibration and acoustics. 118, 292-298.

Besselink, B., van de Wouw, N., Nijmeijer, H., 2011. A semi-analytical of stick-slip oscillations in

drilling systems. ASME Journal of Computational Nonlinear Dynamics. 6(2), 021006-1–021006-9.

Brett, J. F., 1992. The genesis of torsional drillstring vibrations. SPE Drilling Engineering. 7(3), 168-174

Burgess, T. M., McDaniel, G. L., Das, P. K., 1987. Improving BHA tool reliability with drillstring

vibration models: Field experience and limitations. SPE#16109, SPE/IADC Drilling Conference, New

Orleans, Louisiana.

Challamel, N., Sellami, H., Chenevez, E., 2000. A stick-slip analysis based on rock/bit interaction:

theoretical and experimental contribution. IADC/SPE#59230, SPE/IADC Conference, New Orleans,

Louisiana.

Chapman, C. D., Sanchez, J. L., Perez, R. D. L., Yu, H., 2012. Automated closed-loop drilling with ROP

optimization algorithm significantly reduces drilling time and improves downhole tool reliability.

SPE#151736, IADC/SPE Drilling Conference and Exhibition, San Diego, CA.

Chen, S. L., Geradin, M., 1995. An improved transfer matrix technique as applied to BHA lateral

vibration analysis. Journal of Sound and Vibration. 185(1), 93-106.

Chen, S. L., Blackwood, K., Lamine, E., 2002. Field investigation of the effects of stick-slip, lateral and

whirl vibration on roller cone bit performance. SPE Drilling & Completion. 17(1), 15-20.

Chi, A., Zhang, J., Ge, W., Guo, B., 2006. Prediction of drillstring fatigue life under axial-torsional

combined vibration. SPE Gas Technology Symposium, Calgary, Canada.

Chin, W. C., 1994. Wave Propagation in Petroleum Engineering. Gulf publishing Co., Houston.
Christoforou, A. P., Yigit, A. S., 1997. Dynamic modeling of rotating drillstrings with borehole

interactions. Journal of Sound and Vibration. 206 (2), 243-260.

Christoforou, A. P., Yigit, A. S., 2001. Active control of stick-slip vibrations: the role of fully coupled

dynamics. SPE#68093, SPE Middle East Oil Show, Bahrain.

Clayer, F., Vandiver, J. K., Lee, H. Y., 1990. The effect of surface and downhole boundary conditions on

the vibration of drillstrings, SPE#20447, SPE Annual Technical Conference, New Orleans, Louisiana.

Dareing, D. W., 1984a. Guidelines for controlling drillstring vibrations. Journal of Energy Resources

Technology. 106(2). 272-277.

Dareing, D.W., 1984 b. Drill collar length is a major factor in vibration control. Journal of Petroleum

Technology. 36(4), 637-644.

Dareing, D. W., Livesay, B. J., 1968. Longitudinal and angular drillstring vibrations with damping. ASME

Journal of Manufacturing Science and Engineering. 90(4), 1-9.

Dareing, D. W., 1985. Vibration increase power at the bit. Journal of Energy Resources Technology.

107(4), 138-141.

Davis, J., Smyth, G. F., Bolivar, N., Pastusek, P. E., 2012. Eliminating stick-slip by management bit depth

of cut and minimizing variable torque in the drillstring. SPE#151133, SPE/IADC Drilling Conference,

San Diego, CA.

Dawson, R., Lin, Y. Q., Spanos, P. D., 1987. Drill-string stick-slip oscillations. SEM Spring Conference

on Experimental Mechanics, Houston, Texas.

Dunayevsky, V. A., Abbassian, F., Judzls, A., 1993. Dynamic stability of drillstrings under fluctuating

weight on bit. SPE Drilling and Completion. 8(2), 84-92.


Dykstra, M. W., Chen, D. C. K., Warren, T. M., 1994. Experimental evaluations of drill bit and drill string

dynamcis. SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana.

Elsayed, M. A., Wells, R. L., Dareing, D. W., Nagirimadugu, K., 1994. Effect of process damping on

longitudinal vibrations in drillstrings. ASME Journal of Energy Resources Technology. 116, 129-135.

Elsayed, M. A., Dareing, D. W., Vonderheide, M. A., 1997. Effect of torsion on stability, dynamic forces

and vibration characteristics in drillstrings. ASME Journal of Energy Resources Technology. 119(1), 11-

19.

Elsayed, M. A., Raymond, D. W., 2002. Analysis of coupling between axial and torsional vibration in a

compliant model of a drillstring equipped with a PDC bit. ASME 2002 Engineering Technology

Conference on Energy (ETCE), Houston, Texas.

Elsayed, M. A., Phung, C. C., 2005. Modeling of drillstrings. 24th ASME International Conference on

Offshore Mechanics and Arctic Engineering (OMAE), Halkidiki, Greece.

Elsayed, M. A., Aissi, C., 2006. Analysis of shock absorber characteristics for drillstrings. 8th Biennial

ASME Conference on Engineering System Design and Analysis (ESDA 2006), Torino, Italy.

Fazaelizadeh, M., Hareland, G., Aadnoy, B. S., 2010. Application of new 3-D analytical model for

directional wellbore friction. Modern Applied Science, 4(2), 1-22.

Fear, M. J., Abbasian, F., Parfitt, S. H. L., McClean, A., 1997. The destruction of PDC bits by severe

stick-slip vibration. SPE#37639, SPE/IADC Drilling Conference and Exhibition, San Diego, CA.

Finger, J. T., Mansure, A. J., Knudsen, S. D., Jacobson, R. D., 2003. Development of a system for

diagnostic-while-drilling (DWD). SPE# 79884, SPE/IADC Drilling Conference, 19-21 February,

Amsterdam, Netherlands.
Finnie, I., Bailey, J. J., 1960. An experimental study of drill-string vibration. ASME Journal of

Engineering for Industry, 82(2), 129-135.

Germay, C., Van de Wouw, N., Nijmeijer, H., Sepulchre, R., 2009. Nonlinear drillstring dynamics

analysis. SIAM Journal of Applied Dynamic Systems. 8(2), 527-553.

Ghasemloonia, A., Rideout, D. G., Butt, S. D., 2010. The effect of weight-on-bit on the contact behavior

of drillstring and wellbore. 9th International Conference on Bond Graph Modeling (ICBGM2010),

Orlando, Florida.

Ghasemloonia, A., Rideout, D. G., Butt, S. D., 2012. Coupled transverse vibration modeling of drillstrings

subjected to torque and spatially varying axial load. Journal of Mechanical Engineering Science (IMechE,

part C). 227(5), 946-960.

Ghasemloonia, A, Rideout, D. G. Butt, S. D., 2013. Vibration analysis of a drillstring in vibration-assisted

rotary drilling: Finite element modeling with analytical validation. ASME Journal of Energy Resources

Technology. 135(3), 032902-1-032902-18.

Ghasemloonia, A, Rideout, D. G. Butt, S. D., 2014. Analysis of multi-mode nonlinear coupled axial-

transverse drillstring vibration in vibration assisted rotary drilling. Journal of Petroleum Science

Engineering. 116, 36-49.

Gulyaev, V. I., Lugovoi, P. Z., Belova, M. A., Solov'ev, I. L., 2006. Stability of the equilibrium for

rotating drillstrings. International Applied Mechanics. 42(6), 692-698.

Gulyaev, V. I., Lugovoi, P. Z., Gaidaichuk, V. V., Solov'ev, I. L., 2007. Effect of the length of a rotating

drillstring on the stability of its quasistatic equilibrium. International Applied Mechanics. 43 (9), 1017-

1023.
Gulyayev, V. I., Gaidaichuk, V. V., Solovjov, I. L., Gorbunovich, I. V., 2009. The buckling of elongated

rotating drillstrings. Journal of Petroleum Science and Engineering. 67(3-4), 140-148.

Gulyayev, V. I., Borshch, O. I., 2011. Free vibrations of drillstrings in hyper deep bore-wells. Journal of

Petroleum Science and Engineering. 78(3-4), 759-764.

Hakimi, H., Moradi, S., 2009. Drillstring vibration analysis using differential quadrature method. Journal

of Petroleum Science and Engineering. 70, 235-242.

Halsey, G.W., Kyllingstad, A ,Aarrestad, T.V. Lysne, D., 1986. Drill string torsional vibrations:

Comparison between theory and experiment on a full-scale research drilling rig. SPE#15564, 61st SPE

Annual technical Conference and Exhibition, New Orleans, LA.

Halsey, G. W., Kyllingstad, A., Kylling, A., 1988. Torque feedback used to cure stick-slip motion.

SPE#18049, SPE Annual Technical Conference and Exhibition, Houston, TX.

He, X., Halsey, G. W., Kyllingstad, A., 1995. Interactions between torque and helical buckling in drilling.

SPE#30521, SPE Annual Technical Conference & Exhibition, Dallas, Texas.

Heisig, G., Neubert, M., 2000. Lateral drillstring vibrations in extended-reach wells. IADC/SPE# 59235,

IADC/SPE Drilling Conference, New Orleans, Louisiana.

Ho, H. S., 1988. An improved modeling program for computing the torque and drag in directional and

deep wells. SPE#18047, SPE Annual Technical Conference and Exhibition, Houston, Texas.

Hsu, F., Wilhoit, J., James, C., 1965. Lateral vibration of drill pipe including wall reaction. SPE#1046,

Conference on Drilling and Rock Mechanics, Austin, Texas.

Hu, Y., Di, Q., Zhu, W., Chen, Z., Wang, W., 2012. Dynamic characteristics analysis of drillstring in the

ultra-deep well with spatial curved beam finite element. Journal of Petroleum Science and Engineering.

82-83, 166-173.
Hutchinson, M., Dubinsky, V., Henneuse, H., 1995. An MWD downhole assistant driller. SPE#30523,

SPE Annual Technical Conference and Exhibition, Dallas, Texas.

Jaggi, A., Upadhaya, S., Chowdhury, A. R., 2007. Successful PDC/RSS vibration management using

innovative depth of cut control technology: Panna Field, Offshore India. SPE#104388, SPE/IADC

Drilling Conference, Amsterdam, Netherlands.

Jansen, J. D., 1991. Nonlinear rotor dynamics as applied to oil well drillstring vibrations. Journal of Sound

and Vibration. 147(1), 115-135.

Jansen, J. D., Steen, L. V. D., 1995. Active damping of self excited torsional vibrations in oil well

drillstrings. Journal of Sound and Vibration. 179(4), 647-668.

Jardine, S., Malone, D., Sheppard, M., 1994. Putting dampers on drilling’s bad vibrations. Oilfield

Review. 6(1), 15-20.

Jogi, P. N., Macpherson, J. D., Neubert, M., 2002. Field verification of model-derived natural frequencies

of a drillstring. ASME Journal of Energy Resources technology. 124(3), 154-162.

Kaarstad, E., Aadnoy, B. S., 2009. A study of temperature dependent friction in wellbore fluids.

SPE/IADC#119768, SPE/IADC Drilling Conference and Exhibition, Amsterdam, The Netherlands.

Khorshidian, H., Mozaffari, M., Butt, S. D., 2012. The role of natural vibrations in penetration mechanism

of a single PDC cutter. 46th U.S. Rock Mechanics/Geomechanics Symposium (ARMA), Illinois, USA.

Khulief, Y. A., 2000. Spatial formulation of elastic multibody systems with impulsive constraints.

Multibody System Dynamics. 4(4), 383-406.

Khulief, Y. A., Al-Naser, H., 2005. Finite element dynamic analysis of drillstrings. Journal of Finite

Element in Analysis and Design. 41, 1270-1288.


Khulief, Y. A., Al-Sulaiman, F. A., Bashmal, S., 2008. Vibration analysis of drillstrings with string-

borehole interaction. International Journal of Mechanical Engineering Science, IMechE Part C. 222, 2099-

2110.

Khulief, Y. A., and Al-Sulaiman, F. A., 2009. Laboratory investigation of drillstring vibrations.

International Journal of Mechanical Engineering Science (IMechE Part C). 223(10), 2249-2262.

Kolle, J. J. 2004. HydroPulseTM drilling, Final report, http://www.netl.doe.gov/technologies/oil-

gas/publications/.../Final_34367.pdf.

Kreisle, L. F., Vance, J. M., 1970. Mathematical analysis of the effect of shock sub on the longitudinal

vibrations of an oilwell drillstring. SPE Journal. 10(4), 349-356.

Kriesels, P. C., Keulties, W. J. G., Dumont, P., Huneidi, I.,Owoye, O. O., Hartman, R. A., 1999. Cost

saving through an integrated approach to drillstring vibration control. SPE/IADC#57555, SPE/IADC

Middle East drilling technology conference, Abu Dhabi, UAE.

Leine, R. I., Van Campen, D. H., Keultjes, W. J. G., 2002. Stick-slip whirl interaction in drillstring

dynamics. ASME Journal of Vibration and Acoustics. 124(2), 209-220.

Li, H., Butt, S. D., Munaswamy, K., Arvani, F., 2010. Experimental investigation of bit vibration on

rotary drilling penetration rate. 44th US Rock mechanics symposium (ARMA), Utah, USA.

Li, Z., Yanshan, U., Guo, B., 2007. Analysis of longitudinal vibration of drillstring in air and gas drilling.

SPE#107697, SPE Rocky Mountain Oil and Gas Technology Symposium, Denver, Colorado.

Liao, C., Balachandran, B., Karkoub, M., Abdel-Magid, Y. L., 2011. Drillstring dynamics: Reduced-order

models and experimental studies. ASME Journal of Vibration and Acoustics. 133(4), 041008-1041008-8.

Lin, Y. Q., Wang, Y. H., 1991. Stick-slip vibration of drillstrings. ASME Journal of Engineering for

Industry. 113, 38-43.


Liu, X, Samuel, R., 2009. Catenary well profile for extended and ultra-extended reach wells. SPE#

124313, SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana.

Lubinski, A., 1988. Dynamic loading of drillpipe during tripping. Journal of Petroleum Technology.

40(8), 975-983.

Macpherson, J. D., Mason, J. S., Kingman, J. E. E., 1993. Surface measurement and analysis of drillstring

vibrations while drilling. SPE/IADC#25777, SPE/IADC Drilling Conference, Amsterdam, Netherlands.

Macpherson, J. D., Jogi, P. N., Kingman, J. E. E., 2001. Application and analysis of simultaneous near bit

and surface dynamics measurements. SPE Drilling and Completion. 16(4), 230-238.

Mahyari, M. F., Behzad, M., Rashed, G. R., 2009. Drillstring instability reduction by optimum positioning

of stabilizers. International Journal of Mechanical Engineering science, IMechE (Part C). 224, 647-653.

Maidla, E. E., Wojtanowicz, A. K., 1988. Field method of assessing borehole friction for directional well

casing. Journal of Petroleum Science and Engineering. 1(4), 323-333.

Manko, K. I., Pilipenko, V. V., Zapols'ky, L. G., 2003. Use hydrodynamic cavitation for increase of

efficiency of process of well drilling. Fifth International Symposium on Cavitation, Osaka, Japan.

Melakhessou, H., Berlioz, A., Ferraris, G., 2003. A nonlinear well-drillstring interaction model. ASME

Journal of Vibration and Acoustics. 125, 46-52.

Millheim, K., Jordan, S., Ritter, C. J., 1978. Bottom hole assembly analysis using the finite element

method. Journal of Petroleum Technology. 30(2), 265-274.

Mirhaj, S. A., Karstaad, E., Aadnoy, B. S., 2011. Improvement of torque-and-drag modeling in long-reach

wells. Modern Applied Science, 5(5), 10-28.


Mitchell, R. F., Allen, M. B., 1987. Case studies of BHA vibration failure. SPE#16675, SPE Annual

Technical Conference and Exhibition, Dallas, Texas.

Mitchell, R. F., Samuel, R., 2009. How good is the torque/drag model?. SPE Drilling & Completion.

24(1), 62-71.

Mitchell, R. F., Miska, S. Z., 2011, Fundamentals of Drilling Engineering, SPE textbook series, Texas,

USA.

Navarro-Lopez, E. M., Cortes, D., 2007. Avoiding harmful oscillations in a drillstring through dynamical

analysis. Journal of Sound and Vibration. 307, 152-171.

Newman, K., Burnett, T., Pursell, J., Ouahab Gouasmia, S., 2009. Modeling the affect of a downhole

vibrator. SPE/ICo TA Coiled Tubing and Well Intervention Conference, Texas, USA.

Nicholson, J. W., 1994. An integrated approach to drilling dynamics planning, identification and control.

IADC/SPE#27537, SPE/IADC Drilling Conference, Dallas, Texas.

Niedzwecki, J. M., Thampi, S. K., 1988. Have compensated response of long multi-segment drill strings.

Applied Ocean Research. 10(4), 181-190.

Paidoussis, M. P., Luu, T. P., Prabhakar, S., 2008. Dynamics of a long tubular cantilever conveying fluid

downwards, which then flows upwards around the cantilever as a confined annular flow. Journal of Fluids

and Structures. 24, 111-128.

Parfitt, S .H. L., Abbasian, F., 1995. A model for shock performance qualification. SPE/IADC Drilling

Conference, Amsterdam, Netherlands.

Paslay, P., Bogy, D. B., 1963. Drill string vibrations due to intermittent contact of bit teeth. ASME Journal

of Manufacturing Science and Engineering. 85(2), 1-8.


Pastusek, P., Brackin, V. J., Lutes, P. J., 2005. A fundamental model for prediction of hole curvature and

build rates with steerable bottomhole assembly. SPE#95546, SPE Annual Technical Conference and

Exhibition, Dallas, TX.

Placido, J. C. R., Santos, H. M. R., Galeano, Y. D., 2002. Drillstring vibration and wellbore instability.

Journal of Energy resources Technology, 124, 217-222.

Rao, S. S., 2007. Vibration of Continuous Systems, JOHN WHILEY & SONS, Inc, New Jersey.

Reckmann, H., Jogi, P., Kpetehoto, F. T., Chandrasekaran, S., Macpherson, J. D., 2010. MWD failure

rates due to drilling dynamics. SPE#127413-MS, IADC/SPE Drilling Conference and Exhibition, New

Orleans, Louisiana.

Richard, T., Germay, C., Detournay, E., 2007. A simplified model to explore the root cause of stick-slip

vibrations in drilling systems with drag bits. Journal of Sound and Vibration. 305(3), 432-456.

Ritto, T. G., Sampaio, R. Soize, C., 2009. Drill-string nonlinear dDynamics accounting for drilling fluid.

30º CILAMCE-Iberian-Latin-American Congress on Computational Methods in Engineering, Armação

dos Búzios, Rio de Janeiro.

Sahebkar, S. M., Ghazavi, M. R., Khadem, S. E., Ghayesh, M. H., 2011. Nonlinear vibration analysis of

an axially moving drillstring system with time dependent axial load and axial velocity in inclined well.

Journal of Mechanisms and Machine Theory. 46(5), 743-760.

Sampaio, R., Piovan, M. T., Lozano, G. V., 2007. Coupled axial torsional vibrations of drillstring by

means of nonlinear model. Journal of Mechanics Research Communications. 34, 497-502.

Sananikone, P., Kamoshima, O., White, D. B., 1992. A Field method for controlling drillstring torsional

vibrations. IADC/SPE#23891, SPE/IADC Drilling Conference, New Orleans, Louisiana.


Schamp, J. H., Estes, B. L., Keller, S. R., 2006. Torque reduction techniques in ERD wells.

SPE/IADC#98969, SPE/IADC Drilling Conference, Miami, Florida.

Schlumberger Technical Report, 2010. Drillstring vibrations and vibration modeling.

http://www.slb.com/~/media/Files/drilling/brochures/drilling_opt/drillstring_vib_br.pdf

Serrarens, A.F.A., van deMolengraft, J. G., Kok, J. J., van den Steen, L., 1998. H-Infinity control for

suppressing stick-slip in oil well drillstrings. IEEE Control Systems, 18(2), 19-30.

Sheppard, M. C., Wick, C., Burgess, T., 1987. Designing well paths to reduce drag and torque. SPE

Drilling Engineering. 2(4), 344-350.

Shuttleworth, N. E., van Kerkoerle, E. J., Folmer, D. R., Foekema, N., 1998. Revised drilling practices,

VSS-MWD tool successfully addresses catastrophic bit/drillstring vibrations. SPE#39314, IADC/SPE

Drilling Conference, Dallas, TX.

Skaugen, E., Kyllingstad, A., 1986. Performance testing of shock absorbers. 61st SPE Annual Technical

Conference and Exhibition, New Orleans, LA.

Skaugen, E., Kyllingstad, A., Aarrestad, T. V., Tonnesen, H. A., 1987. Experimental and theoretical

studies of vibrations in drillstrings incorporating shock absorbers. 12th World Petroleum Congress,

Houston, Texas.

Sotomayor, G. P. G., Placido, J. C., Cunha, J. C., 1997. Drillstring vibration: How to identify and

suppress. SPE#39002, Latin American and Caribbean Petroleum Engineering Conference, Rio de Janerio,

Brazil.

Spanos, P. D., Payne, M. L., 1992. Advances in dynamic bottomhole assembly modeling and dynamic

response determination. SPE/IADC# 23905, SPE/IADC Drilling Conference, New Orleans, Louisiana.
Spanos, P. D., Sengupta, R. A., Pasley, P. R., 1995. Modeling of roller cone bit lift-off dynamics in rotary

drilling. ASME Journal of Energy Resources Technology. 117(3), 115-124.

Spanos, P. D., Payne, M. L., Secora, C. K., 1997. Bottom-hole assembly modeling and dynamic response

determination. ASME Journal of Energy Resources Technology. 119(3), 153-158.

Spanos, P. D., Chevalllier, A. M., Politis, N. P., 2002. Nonlinear stochastic drillstring vibrations. ASME

Journal of Vibration and Acoustics. 124(4), 512-519.

Thomsen, J. J., 2003. Vibrations and Stability, second ed. Springer, Berlin.

Trindade, M. A., Wolter, C., Sampaio, R., 2005. Karhunen-Loeve decomposition of coupled axial-

bending vibrations of beams subject to impacts. Journal of Sound and Vibration. 279, 1015-1036.

Tucker, W. R., Wang, C., 1999. An integrated model for drillstring dynamics. Journal of Sound and

Vibration. 224, 123-165.

Vandiver, J. K., Nicholson, J. W., Shyu, R. J., 1990. Case studies of the bending vibration and whirling

motion of drill collars. SPE Drilling Engineering. 5(4), 282-290.

Vaz, M. A., Patel, M. H., 1995. Analysis of drillstrings in vertical and deviated holes using the Galerkin

technique. Engineering Structures. 17(6), 437-442.

Voronov, S. A., Gouskov, A. M., Kvashnin, A. S., Butcher, E. A., Sinha, S. C., 2007. Influence of

torsional motion on the axial vibrations of a drilling tool. Journal of Computational and Nonlinear

Dynamics. 2, 58-64.

Walker, B. H., Friedman, M. B., 1977. Three dimensional force and deflection analysis of a variable cross

section drillstring. ASME Journal of Pressure Vessels. 99(5), 367-373.


Wang, X., Yuan, Z., 2012. Investigation of frictional effects on the nonlinear buckling behavior of a

circular rod laterally constrained in a horizontal rigid cylinder. Journal of Petroleum Science and

Engineering. 90-91, 70-78.

Warren, T. M., Oster, J. H., Sinor, L. A., Chen, D. C. K., 1998. Shock sub performance tests. SPE#39323,

IADC/SPE Drilling Conference, Dallas, Texas.

Wise, J. L., Finger, J. T., Mansure, A. J., Knudsen, S. D., Jacobson, R. D., Grossman, J. W., Pritchard, W.

A., Mathews, O., 2003., Hard-rock drilling performance of a conventional PDC drag bit operated with,

and without, benefit of real-time downhole diagnostics. Geothermal Resources Council Transactions. 27,

197-205.

Wolf, S. F., Zacksenhouse, M., Arian, A., 1985. Field measurement of downhole drillstring vibrations.

SPE#14330, SPE 60th Annual Technical Conference and Exhibition, Las Vegas, NV.

Wong, S. V., Hamuda, A. M. S., Hasmi, M. S. J., 2001. Kinematic contact-impact agorithm with friction.

International Journal of Crashworthiness. 6(1), 65-82.

Wu, X., Karuppiah, V., Nagaraj, M., Partin, U. T., Machado, M., Franco, M., Duvvuru, H. K., 2012.

Identifying the root cause of drilling vibration and stick-slip enables fit for purpose solution. SPE#151347,

IADC/SPE Drilling Conference, San Diego, CA.

Yaveri, M., Damani, K., Kalbhor, H., 2010. Solutions to the downhole vibrations during drilling.

SPE#136956, 34th Annual SPE International Conference and Exhibition, Tinapa, Nigeria.

Yigit, A. S., Christoforou, A. P., 1996. Coupled axial and transverse vibrations of oilwell drillstrings.

Journal of Sound and Vibration. 195(4), 617-627.

Yigit, A. S., Al-Ansary, M. D., Khalid, M., 1997. Active control of drillstring vibrations by mode

localization. Journal of Structural Control. 4(1), 47-63.


Yigit, A.S., Christoforou, A.P., 1998. Coupled torsional and bending vibrations of drillstrings subject to

impact with friction. Journal of sound and vibration. 215 (1), 167-181.

Yigit, A. S., Christoforou, A. P., 2000. Coupled torsional and bending vibrations of actively controlled

drillstrings. Journal of Sound and Vibration. 234 (1), 67-83.

Zamanian, M., Khadem, S. E., Ghazavi, M. R., 2007. Stick-slip oscillations of drag bits by considering

damping of drilling mud and active damping system. Journal of Petroleum Science and Engineering. 59(3-

4), 289-299.

Zhang, Q., Miska, S., 2005. Effects of flow-pipe interaction on drill pipe buckling and dynamics. ASME

Journal of Pressure Vessels. 127, 129-136.

Zhao, G. H., Lian, Z., 2008. Analysis of collision between drillstring and well sidewall. International

Journal of Modern Physics B. 22(31-32), 5459-5464.

Zhu, X, Liu, W., 2013. The effects of drill string impacts on wellbore stability. Journal of Petroleum

Science and Engineering. 109, 217-229.

Contribution and Highlights of the Paper

• Review of all uncoupled-coupled models of drillstring vibration and mitigation


• Review of boundary conditions for all modes, contact and bit-rock force
• Review of mud damping models, mud flow interaction and structural damping
• Review of derivation of governing equations and solution methods
• Review of of drillstring vibration modeling in emerging technologies

You might also like