You are on page 1of 22

Journal of Petroleum Science and Engineering 53 (2006) 203 – 224

www.elsevier.com/locate/petrol

Optimal determination of rheological parameters for


Herschel–Bulkley drilling fluids and impact on pressure drop,
velocity profiles and penetration rates during drilling
V.C. Kelessidis a,⁎, R. Maglione b , C. Tsamantaki a , Y. Aspirtakis a
a
Mineral Resources Engineering Department, Technical University of Crete, Chania, Greece
b
Vercelli, Italy
Received 1 June 2005; accepted 16 June 2006

Abstract

Drilling fluids containing bentonite and bentonite–lignite as additives exhibit non-Newtonian rheological behavior which can
be described well by the three parameter Herschel–Bulkley rheological model. It is shown that determination of these parameters
using standard techniques can sometimes provide non-optimal and even unrealistic solutions which could be detrimental to the
estimation of hydraulic parameters during drilling. An optimal procedure is proposed whereby the best value of the yield stress is
estimated using the Golden Section search methodology while the fluid consistency and fluid behavior indices are determined with
linear regression on the transformed rheometric data. The technique yields in many cases results which are as accurate as these
obtained by non-linear regression but also gives positive yield stress in cases where numerical schemes give negative yield stress
values. It is shown that the impact of the values of the model parameters can be significant for pressure drop estimation but less
significant for velocity profile estimation for flow of these fluids in drill pipes and concentric annuli. It is demonstrated that very
small differences among the values of the model parameters determined by different techniques can lead to substantial differences
in most operational hydraulic parameters in oil-well drilling, particularly pressure drop and apparent viscosity of the fluid at the
drilling bit affecting penetration rates, signifying thus the importance of making the best simulation of the rheological behavior of
drilling fluids.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Drilling fluids; Rheology; Herschel–Bulkley; Pressure drop; Velocity profile; Penetration rates

1. Introduction that above around 120 °C and in conditions of high


salinity, bentonite slurries begin to thicken catastrophi-
In oil-well drilling, bentonite is added in drilling cally (Gray and Darley, 1980; Bleler, 1990; Elward-
fluids for viscosity control, to aid the transfer of cuttings Berry and Darby, 1992). Attempts to describe and to
from the bottom of the well to the surface, and for predict the gelling tendencies of bentonite suspensions
filtration control to prevent filtration of drilling fluids have not yet resulted in a concise method which could
into the pores of productive formations. It is long known predict rheological and filtration properties, given the
amount of added bentonite and its physical character-
⁎ Corresponding author. istics. The flocculation of bentonite suspensions at high
E-mail address: kelesidi@mred.tuc.gr (V.C. Kelessidis). temperatures could be resolved with the addition of
0920-4105/$ - see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.petrol.2006.06.004
204 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

thinners to reduce the rheology of the mixture but many the accuracy in the calculations and the simplicity of the
thinners degrade over the same temperature range. A use is required and the best way to achieve this is with the
thinner with high thermal stability is lignite (Clark, use of the Herschel–Bulkley rheological model. The
1994; Briscoe et al., 1994; Miano and Rabaioli, 1994) three parameter Herschel–Bulkley model has not been
and recent evidence (Mihalakis et al., 2004; Kelessidis used widely until very recently, although it was not only
et al., 2005) demonstrated the stabilizing effect of Greek proposed almost at the same time as the Bingham plastic
lignite in terms of rheological and filtration control of model but it also describes most drilling fluid rheological
bentonite slurries. Their measurements also showed that data much better (Fordham et al., 1991; Hemphil et al.,
the three parameter Herschel–Bulkley model describes 1993; Maglione and Ferrario, 1996; Kelessidis et al.,
well the rheology of these bentonite–lignite water 2005). The reason for the nonfrequent use is that
suspensions. derivation of the model's three parameters is complex
Various rheological models have been proposed to (Nguyen and Boger, 1987; Hemphil et al., 1993).
describe the rheological behavior of bentonite mixtures, Furthermore, analytical solutions for laminar flow in
particularly for drilling applications. The two parameter pipe and annuli are not possible, requiring either
Bingham plastic model (Bingham, 1922) or the power graphical or trial-and-error solutions (Hanks, 1979;
law model (Govier and Aziz, 1972; Bourgoyne et al., Govier and Aziz, 1972; Fordham et al., 1991). The
1991) are used most often because of their simplicity advent of personal computers and their online use in the
and the fair agreement of predictions with the field, however, made trial-and-error solutions trivial
rheograms. The power law model, although useful as a tasks, hence, more and more investigators opt to use
first correction to Newtonian behavior, it may lead to Herschel–Bulkley rheological models in fluid me-
substantial errors if the fluid exhibits yield stress. Other chanics computations of drilling fluids (Maglione et
two parameter models like the Casson model (Casson, al., 1999a; Maglione et al., 2000; Becker et al., 2003). A
1959; Hanks, 1989) or the Prandl–Eyring model search in the Society of Petroleum Engineers electronic
(Govier and Aziz, 1972) have not found wide library of scientific articles, covering the period of 1975–
acceptance. Three constant parameter models have 2003, resulted in 319 articles having as keywords ‘power
been proposed by Herschel and Bulkley (1926), by law’, 131 articles with keywords ‘Bingham’, 51 articles
Graves and Collins (1978), by Gucuyener (1983) and by with keywords ‘Herschel–Bulkley’, and 16 articles with
Robertson and Stiff (1976). More complex four keywords ‘Casson’.
parameters models (Shulman, 1968; Mnatsakanov et Viscometric data reduction procedures applicable to
al., 1991) or even five parameter models (Maglione et various rheological models have been proposed by
al., 1996) have also been proposed. Detailed description many investigators, addressing also some of the inherent
of the various rheological models proposed and problems associated with data reduction (Krieger, 1968;
derivation of the appropriate flow equations have been Darby, 1985; Borgia and Spera, 1990; Yeow et al.,
given by Bird et al. (1982) and by Maglione and 2000). The standard procedure for the estimation of the
Romagnoli (1999). three rheological parameters for Herschel–Bulkley
The more complex rheological models are deemed liquids, with rheological equation,
more accurate in predicting the behavior of drilling fluids
than the two parameter models that are widely accepted s ¼ sy þ Kγn ð1Þ
at present. However, there is not wide acceptance and
where τ, τy are the shear stress and the yield stress
wide application of the more complex models because of
respectively, K, n are the fluid consistency and fluid
the difficulty in finding analytical solutions for the
behavior indices respectively and γ is the shear rate, is
differential equations of motion and because of the
through non-linear regression of the viscometric data
complexity of the calculations for the derivation of the
from concentric cylinder geometry. This is normally
appropriate hydraulic parameters such as Reynolds
done using a numerical package, minimizing the sum of
number, flow velocity profiles, circular and annular
error squares and judging the goodness of fit through the
pressure drops and criteria for transition from laminar to
value of the correlation coefficient Rc2 from the
turbulent flow. Simulation of rotational viscometer data
linearized form of Eq. (1), as in Eq. (2),
of non-Newtonian fluids appears to be better when a
larger number of rheological parameters is used but in lnðs−sy Þ ¼ lnK þ nlnðγÞ ð2Þ
this case, the hydraulic parameters can be obtained only
by numerical methods for most of the more complex However, non-linear fit to various data in this
rheological models. As of today, a compromise between laboratory with a numerical package sometimes has
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 205

given the best fit (highest correlation coefficient Rc2) with the annular sections in a drilling circuit can all be largely
negative values for the yield stress, which is mean- influenced by this choice and can have desirable or
ingless. In these situations, the condition τy > 0 must be undesirable carry-over effects. For instance, a variation
imposed to get meaningful results affecting thus the in the velocity profile and the Reynolds number can
optimum determination of all three parameters, while affect the cuttings carrying capacity of the drilling fluid.
there is also a possibility of non-unique solutions. So, the same drilling fluid can be estimated to perform
Moreover, there is concern over the use of the correlation better if a rheological parameter set is preferred to
coefficient as an indicator of the goodness of fit because another, even though both may have very similar values
it should be used only for linear functions (Helland, of Rc2, Q2 or BIV.
1988; Ohen and Blick, 1990). In addition, consideration As most flows during drilling oil and gas wells are
should be given to differences of rheological parameters laminar, for which analytical solutions exist for the flow
and their effect on subsequent computations, when of Herschel–Bulkley fluids in pipes and annuli, although
choosing data with similar correlation coefficients (Rc2), not explicit, effort should be made to determine the
like for example data with Rc2 = 0.95 versus data with impact of the use of different sets of rheological para-
Rc2 = 0.96, or between Rc2 = 0.98 and Rc2 = 0.99. meters, derived with different methodologies, on the
Many times, especially when curve fitting non-linear main parameters of interest, pressure drop and velocity
functions, the sum of square errors (Q2) is utilized for profiles, thus providing a more robust indicator for the
the goodness of fit, which is defined as, proper choice of the appropriate rheological parameters.
X Prior work (Maglione and Romagnoli, 1998; Maglione,
Q2 ¼ ðyi − ŷ i Þ2 ð3Þ 1999) has suggested that the variation of the flow
i behavior index of the Herschel–Bulkley model could be
the most important factor because it can influence all
where yi, ŷi are the measured and the predicted
hydraulic parameters of the drilling hydraulic circuit,
quantities. Another method to determine the goodness
from flow regimes, to velocity profiles, to pressure drop
of fit of a particular rheological model to rheometric data
and to rates of penetration. It was shown that the lower
is the method based on the best index value (BIV). It is
the flow behavior index is, the lower the pressure drop is,
defined as the ratio between the sum of the squares of
in the circular and the annular sections. As the maximum
the deviation of the predicted value from the average
pressure available is normally constant, the gained
value ȳ and the sum of the squares of the deviation of
pressure drop (due to a lower n) can be used to increase
the measured value from the average value (Maglione
the flow rate thus providing more hydraulic power at the
and Romagnoli, 1999).
drilling bit or allowing the drilling of longer sections.
X The authors concluded, however, that more studies were
ð ŷ i − ȳ Þ2
required to better define and resolve this problem.
BIV ¼ X i
ð4Þ The scope of the present work is to propose a different
ðyi − ȳ Þ2
i
and optimal methodology to determine the three
Herschel–Bulkley rheological parameters of drilling
The closer the value of BIV to one, the better is the fluids from concentric cylinder viscometric data, avoid-
capacity of the analytical equation (or the rheological ing the observed pitfalls of current non-linear regression
model) to approximate viscometer data. Values of BIV techniques, which may give meaningless negative yield
larger than one indicate tendency for over-prediction stress values. Furthermore, the paper aims to demon-
with the model while values lower than one indicate strate that the particular choice of the rheological
tendency for under-prediction. parameters, among similarly performing data fit curves,
The final choice of the particular parameters can can greatly affect the determination of pressure drop and
always be questioned when the choice could be, for velocity profiles of drilling fluids flowing in drilling
example, between data fits with Rc2 = 0.995 or Rc2 = 0.998, hydraulic circuits (pipe and annuli) and to present field
between Q2 = 0.9 or Q2 = 1.1, and between BIV = 0.99 or cases of these implications. An evaluation is also
BIV = 1.01, as it will be demonstrated later. Most of the attempted about the conditions for preferring a particular
time, two or even more rheological parameter sets can rheological parameter set, whether it is the best fit of a
approximate well and in a similar way rotational particular rheological model to rheometric data or also
viscometer data, but in turn they could have large effects the impact of the specific choice of rheological parameter
on the end results. For example, parameters like Reynolds values on pressure drop and velocity profile estimation
number, velocity profile, pressure drop for the circular and for flow of drilling fluids during drilling.
206 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

2. Experimental data

The rheometric data used in this work were taken


from an on-going research project on the use of Greek
lignite as thinning agent in bentonite suspensions,
particularly when exposed to high temperatures (Miha-
lakis et al., 2004; Kelessidis et al., 2005), samples S1
through S12, as well as from published field data from
drilling operations reported by Merlo et al. (1995) and
drilling fluid data of Blick (1992), as reported by Al-
Zahrani (1997).
The experimental data of Mihalakis et al. (2004) and Fig. 1. Rheograms of 6.42% w/v bentonite suspensions in water,
Kelessidis et al. (2005) were taken with water–bentonite hydrated (sample S1) and thermally aged at 177 °C for 16 h (sample
suspensions at 6.42% w/v, either hydrated for 24 h at S2). The rheological measurements were taken at 25 °C.
room temperature or aged statically in an aging cell for
16 h at 177 °C. Various types of lignites from different standard methodology, using non-linear regression with
places in Greece were added at 0.5% w/v and 3% w/v in an appropriate numerical package. These values are
the suspension in a similar fashion (hydrated or aged given in Table 1 together with the correlation coeffi-
thermally). The bentonite used was a commercial cient, the sum of errors squared and the best index value,
product used in oil-well drilling (Zenith©) provided by as defined above. From the rheograms presented in Figs.
S&B Industrial Minerals S.A. All samples were prepared 1 and 2 above, as well as from the values of the
following American Petroleum Institute procedures rheological parameters listed in Table 1, it is evident that
(2000). The samples were agitated vigorously for the rheological behavior of these suspensions is typical
5 min before testing and measurements were made of yield-pseudoplastic fluids and modeling their rheo-
with a continuously varying rotational speed rotating logical behavior as Herschel–Bulkley fluids is more
viscometer (Grace, M3500) at two sample temperatures, than appropriate.
25 °C and 65 °C and 12 speeds: 3, 6, 10, 20, 30, 60, 100, There exist, however, situations where application of
200, 300, 400, 500 and 600 rpm. The data is reproduced non-linear regression to rheometric data gives as optimal
in Appendix B (Table B1). set of the three rheological parameters, a set with
The data of Merlo et al. (1995), reproduced in Table negative values of the yield stress, which of course is
B2, were derived with drilling fluid from field meaningless. Reported data showing this behavior is
operations during drilling circulation tests at various presented in Table 2 for data of Kelessidis et al. (2005),
sections of the well, taking fluid samples from the outlet of Merlo et al. (1995) and of Blick (1992). There is the
of the drilling circuit. The rheometric data were derived possibility of imposing the condition τy > 0 when
with a Huxley–Bertran high pressure high temperature deriving the rheological parameters with non-linear
rotational viscometer for samples S13, S14, S15, S16 regression which, however, leads to non-optimal solu-
and S17, while data for sample S18 were derived with a tions, as will be demonstrated later.
Fann VG 35 six speed rotational viscometer (Merlo et It is evident that the use of a numerical package to
al., 1995). Al-Zahrani (1997) reports that the rheological apply non-linear regression to rheometric data in order
data of Blick (1992), samples S19 to S21 reproduced in to derive Herschel–Bulkley parameters is not always
Table B3, were taken with a rotary viscometer for the optimal procedure since the method can lead to
different suspensions prepared by adding various inappropriate negative values for the yield stress. For
quantities of Wyoming bentonite in water. yield-pseudoplastic fluids, as most bentonite suspen-
sions are, a yield value exists either as an engineering
3. Rheological parameter estimation for reality or as an inherent fluid property, although there
Herschel–Bulkley drilling fluids is considerable dispute about this issue (Barnes and
Walters, 1985; Chen, 1986; Hartnett and Hu, 1989;
3.1. Current methodology Evans, 1992; Schurz, 1992; De Kee and Chan Man
Fong, 1993; Barnes, 1999). The yield value of the
Some of the rheometric data, given in Table B1, are fluid should be derived first using an appropriate
presented in Figs. 1 and 2. The three rheological technique, followed by the estimation of the other two
Herschel–Bulkley parameters were derived according to parameters with non-linear regression applied to the
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 207

a set of parameters derived with a correlation coefficient


of 0.996, and the final choice, which will have an impact
on the end result, pressure drop and velocity profiles,
can be a dilemma.
The procedure could be improved if a proper initial
value of τy is assumed, either taken from the original
rheogram or from a semi-log plot, with τy the y-
intercept, as shown in Fig. 3. In this semi-log plot of
[τ − log(γ)], the last points at low shear rates should lie
normally on a straight line, which when extended to
very low values of the shear rate (τ → 0) it should cross
Fig. 2. Rheograms of bentonite–lignite–water suspensions (samples the τ-axis at the value of the yield stress, a procedure
S7 and S12) both thermally aged at 177 °C for 16 h. The rheological followed by some investigators (Benna et al., 1999;
measurements were taken at 25 °C. Barnes, 1999). This method, however, is not deemed
very accurate because measurements at very low shear
linearized form of Herschel–Bulkley rheological rates may suffer from wall slip effects (Guillot, 1990;
equation (Eq. (2)). Barnes, 1995).
This could be accomplished through a trial-and-error Other investigators have proposed to measure
procedure, where various values of τy are assumed and separately τy by other means, for example utilizing the
the original (τ − γ) rheological data is transformed to [ln vane method (Nguyen and Boger, 1987; Alderman et
(τ − τy) − ln γ] data in order to establish the best linear al., 1991; Barnes and Nguyen, 2001) which, although it
curve, determined from the correlation coefficient and leads to better determination for the yield stress, it is
the sum of error squares, giving as y-intercept, (ln K), time consuming and not applicable for drilling
while the slope of the curve will be the flow behavior operations.
index, (n) (Hemphil et al., 1993; Turian et al., 1997).
With this approach, however, there is no guarantee that 3.2. New methodology
the optimum values have been derived since the
determination of (K,n) depends on the correct estimation A new methodology is proposed, amenable to
of τy. Turian et al. (1997) found the procedure very computer implementation, which determines the three
tedious and the results very sensitive to the assumed rheological Herschel–Bulkley parameters avoiding
value of the yield stress. They further stated that a major potential errors of deriving negative yield stress values.
weakness of the Herschel–Bulkley model has been the The methodology is based on an initial optimal
inability to unambiguously establish the model para- determination of τy, using a near optimum form of the
meters since different sets of these values can provide Golden Section search, sometimes called the Fibonacci
equivalent fits of the experimental data, exactly the kind search, followed by linear regression of the linearized
of behavior described in this work. Furthermore, a set of form of the Herschel–Bulkley rheological equation. The
Herschel–Bulkley rheological parameters derived with technique has been proposed by Ohen and Blick (1990)
a correlation coefficient of 0.998 could be different than for Robertson and Stiff fluids (Robertson and Stiff, 1976)

Table 1
Herschel–Bulkley rheological parameters of bentonite–water and bentonite–lignite–water suspensions (data of Kelessidis et al., 2005)
Sample τy (Pa) K (Pa sn) n R2c Q2 (Pa2) BIV
S1 8.4748 3.4010 0.2556 0.9876 3.8520 0.9980
S2 11.3025 5.9115 0.2645 0.9885 12.6820 0.9883
S3 0.0788 2.3861 0.3407 0.9340 44.8843 0.9343
S4 0.0000 0.2050 0.4930 0.9450 4.1802 0.6399
S5 0.6751 0.0732 0.7001 0.9937 0.7718 0.9927
S6 6.3938 0.4498 0.5001 0.9951 1.2885 0.9945
S7 2.4095 0.1251 0.7012 0.9975 0.9188 0.9980
S8 1.1843 0.1265 0.6436 0.9965 0.5671 0.9961
S9 3.4701 0.0313 0.8045 0.9666 3.3209 0.9687
S10 0.3793 0.0567 0.6196 0.9983 0.0395 0.9982
208 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

Table 2
Herschel–Bulkley rheological parameters of bentonite–water and bentonite–lignite–water suspensions with negative τy, derived from nonlinear
regression
Sample τy (Pa) K (Pa sn) n R2c Q2 (Pa2) BIV Source
S11 − 0.2685 0.2486 0.5312 0.9957 0.5402 0.9848 Kelessidis et al. (2005)
S12 − 0.2880 0.2210 0.5841 0.9963 0.7890 0.9960 Kelessidis et al. (2005)
S14 − 0.0932 1.9580 0.3488 0.9990 0.2595 0.9991 Merlo et al. (1995)
S15 − 1.1650 2.3990 0.3158 0.9988 0.2850 0.9981 Merlo et al. (1995)
S17 − 0.6213 1.1400 0.3704 0.9994 0.0739 0.9995 Merlo et al. (1995)
S19 − 0.0906 1.1371 0.4393 0.9964 1.3320 0.9958 Blick (1992)
S20 − 3.7262 5.4440 0.3222 0.9982 2.4130 0.9985 Blick (1992)
S21 − 6.0980 9.2885 0.3208 0.9988 4.7160 0.9996 Blick (1992)

which follow rheological Eq. (5) below, with A, B, C the middle point b a fractional distance from one end, for
Robertson–Stiff rheological parameters, and it has been example a, of a value equal to the golden ratio of
extended to Herschel–Bulkley fluids in this work. 0.61803 and a fractional distance from the other end of
0.38197 (Fig. 4). At each stage of the search of the
s ¼ Aðγ þ CÞB ð5Þ bracketing interval of the minimum, the next point to be
tried is the point x which is a fractional distance 0.61803
The aim is to find the value of the yield stress, τy, from one end and a distance of 0.38197 from the other
which minimizes the error of the difference between the end (Press et al., 1992).
predicted and the measured shear stress values, in other To estimate the three rheological parameters τy, K, n,
words, to bracket a minimum for a given function f (x) a combination of iteration, to determine the yield stress
in the interval (a, c). A minimum is known to be using the Golden Section search method, and a least
bracketed only when there is a triplet of points a < b < c squares fitting of the linearized Herschel–Bulkley equa-
(or c < b < a) such that f (b) is less than both f (a) and f tion, is used (Ohen and Blick, 1990). This is accomplished
(c) (Press et al., 1992). The function is then evaluated at by first picking an initial interval of search (L, U), with L
an intermediate point, x, which is chosen, either between the lower limit and U the upper limit, defined as,
a and b or between b and c. After evaluating f (x), if f
(b) < f (x), the new bracketing triplet becomes (a,b, x), U ¼ sy0 þ tol ð6Þ
otherwise, if f (b) > f (x), the triplet becomes (b, x, c).
The process continues until the bracketing interval is L ¼ sy0 −tol ð7Þ
tolerably small. The strategy for choosing point b (or x),
given (a, c) leads to the Golden Section search, which where tol is the half width of the search interval and τy0
states that the optimal bracketing interval a, b, c has its is an initial estimate of the yield stress. The lower limit is
taken very close to zero while the upper limit can be
taken as 2 · τy0. Although the value of τy0 can be any
non-zero value and the procedure will converge, τy0 can
be estimated following the graphical procedure proposed
by Robertson and Stiff (1976) as modified by Ohen and
Blick (1990). Given a rheological data set of τ −γ, three
data sets are defined, one at the lowest shear stress, τmin
−γmin, one at the highest shear stress, τmax − γmax, while
the third set is based on the geometric mean shear stress

Fig. 3. Graphical determination of yield stress, τy, taken as the y-


intercept from a semi-log plot of the original rheogram. Fig. 4. Schematic explaining the Golden Section ratio.
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 209

from the first two points, that is, (τ̄)2 = (τmin *τmax), with quick convergence. The search interval varies according
γ̄ derived by interpolation. A system of three equations to the conditions,
with three unknowns is thus derived as,
R2c1 < R2c2 YU ¼ sy1 ð15Þ
smin ¼ sy0 þ K0 γnmin
0
ð8Þ

smax ¼ sy0 þ K0 γnmax


0
ð9Þ R2c1 > R2c2 YL ¼ sy2 ð16Þ
s̄ ¼ sy0 þ K0 γ̄ n0 ð10Þ
R2c1 ¼ R2c2 YU ¼ sy1 ; L ¼ sy2 ð17Þ
This system can be solved to get the initial value of
τy0. Combination of Eqs. (8)–(10) gives, The whole procedure can be performed with a
numerical package.
smax −smin γnmax
0
−γnmin
0

¼ n0 ð11Þ
smax − s̄ γmax −γ̄ 0
n
3.3. Computational results and comparison with
with 0 < n0 < 1. Eq. (11) can be solved easily numerically experimental data
to get n0 and it is then used to derive the initial value for
yield stress, τy0, This methodology has been applied to several data
sets, given in Tables B1–B3 in Appendix B, and has
s̄ ðγ̄ 2n0 −γnmin
0
γnmax
0
Þ proved to work extremely well. In Table 3, the
sy0 ¼ ð12Þ
γ̄ n0 ðγnmin
0
þ γmax −2γ̄ 0 Þ þ γ̄ 2n0 −γnmin
n0 n 0
γnmax
0
rheological parameters that have been calculated using
both procedures, non-linear regression with a numerical
The triplet of points (L, τy0, U) is thus established package and with the Golden Section technique are
following the above procedure. The new points, in the presented for the cases where non-linear regression
search of the minimum, are evaluated using the golden provides meaningful results, in other words, positive
ratio of 0.61803 as, yield stress values. Rheograms of some of the samples
are shown in Figs. 6 and 7 together with the curves
sy1 ¼ L þ 0:61803ðU −LÞ ð13Þ derived from non-linear regression (NL) and for Golden
Section search (GS).
for the point between (L, τy0), while for the point The results show minor differences among the two
between (τy0, U) the new value becomes, methods both in terms of all three rheological
parameters and the two of the three indices of correlation
sy2 ¼ U −0:61803ðU −LÞ ð14Þ (Rc2, Q2) while the third index (BIV) does not really
reflect the similarities between the estimated parameters.
The functional relationship for optimization is This close agreement demonstrates the ability of the new
provided by the correlation coefficient Rc2 of Eq. (2) proposed scheme to properly determine the rheological
for the chosen value of τy. Schematically, the relation- parameters with similar success as the application of
ship Rc2 − τ y is shown in Fig. 5 for successive nonlinear regression using numerical packages, for the
approximations of τy. It is a unimodal function implying cases that the latter predict positive yield stress values.
The results show also that for all fluid samples,
except sample S8, the flow behavior index, n,
determined with the GS method, is smaller when
compared to the value determined with the NL method,
while the yield point may be smaller or larger. In
practical terms, and while keeping all other rheological
parameters constant, a smaller n-value results in
flattening of the velocity profile improving the carrying
capacity of the drilling fluid, extending laminar flow
regime and decreasing pressure drop (Maglione and
Robotti, 1996; Maglione et al., 1999a,b).
The close agreement between the rheological para-
Fig. 5. Variation of correlation coefficient with assumed values of yield meters obtained by the two methodologies raises the
stress (the case represents real data). point about the optimal data set to be used when the
210 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

Table 3
Comparison of rheological parameters derived from non-linear regression (NL) and by Golden Section (GS)
Sample and method τy (Pa) K (Pa sn) n R2c Q2 BIV Source
S7
NL 2.4095 0.1251 0.7012 0.9975 0.9188 0.9980 Kelessidis et al. (2005)
GS 2.4141 0.1369 0.6842 0.9910 1.3124 0.9401
S8
NL 1.1843 0.1265 0.6436 0.9965 0.5671 0.9961 Kelessidis et al. (2005)
GS 1.3012 0.1058 0.6680 0.9967 0.6013 0.9855
S9
NL 3.4701 0.0313 0.8045 0.9666 3.3209 0.9687 Kelessidis et al. (2005)
GS 2.8973 0.1566 0.5661 0.9800 4.8284 0.8285
S10
NL 0.3793 0.0567 0.6196 0.9983 0.0395 0.9982 Kelessidis et al. (2005)
GS 0.2847 0.0839 0.5625 0.9960 0.0670 0.9615
S13
NL 1.7020 1.2063 0.4352 0.9971 1.0847 0.9959 Merlo et al. (1995)
GS 0.0000 1.9940 0.3704 0.9952 1.8105 1.0246
S16
NL 0.1747 0.9448 0.4097 0.9990 0.1563 0.9989 Merlo et al. (1995)
GS 0.0379 1.0200 0.3993 0.9995 0.1636 0.9956
S18
NL 2.6750 0.2492 0.6607 0.9982 0.7375 0.9977 Merlo et al. (1995)
GS 1.6813 0.6496 0.5173 0.9950 5.1244 0.8035
Positive yield stress values.

choice is among values giving very similar data fit. sometimes result in non-optimum solutions giving
Furthermore it raises questions about the implications lower correlation coefficients, much higher sum of
and the real impact on the operational parameters of squared errors and values of best index value signifi-
interest, pressure drop and velocity profiles. cantly different than one, compared to cases of non-
The new methodology has been applied also to linear regression but without this imposition and to
rheometric data which give negative yield stress values cases of Golden Section search methodology (samples
when applying non-linear regression, with the results S11, S12, S15). For some of the cases though, NLP
shown in Table 4. The listed rheological parameters are method may result in as good predictions as the Golden
derived using three different techniques: application of Section search (samples S14, S17, S19, S20 and S21).
non-linear regression (NL, numerical-no penalty), The new proposed scheme provides good solutions, as
application of non-linear regression with imposition of evidenced by the sum of squared errors, the correlation
τy > 0 (NLP, numerical with penalty) and application of
the Golden Section search (GS). Imposing τy > 0 may

Fig. 6. Comparison of rheograms of original data with results from Fig. 7. Comparison of rheograms of original data with results from
rheological models for bentonite–lignite–water suspensions. rheological models for drilling mud.
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 211

Table 4
Comparison of rheological parameters derived from numerical package and by proposed scheme, non-optimal solutions
Sample and method τy (Pa) K (Pa sn) n R2c Q2 BIV Source
S11
NL − 0.2685 0.2486 0.5312 0.9957 0.5402 0.9948 Kelessidis et al. (2005)
NLP 0.0000 0.2326 0.5166 0.9590 5.1540 0.9149
GS 0.0718 0.1462 0.6068 0.9942 0.7865 1.0263
S12
NL − 0.2880 0.2210 0.5841 0.9963 0.7890 0.9960 Kelessidis et al. (2005)
NLP 0.0001 1.0493 0.3218 0.9160 34.15 0.4435
GS 0.3976 0.0940 0.7036 0.9942 1.5646 0.9881
S14
NL − 0.09323 1.9580 0.3488 0.9990 0.2595 0.9991 Merlo et al. (1995)
NLP 0.0000 1.8900 0.3530 0.9990 0.2627 0.9938
GS 0.0000 1.9050 0.3523 0.9990 0.2608 0.9987
S15
NL − 1.1650 2.3990 0.3158 0.9988 0.2850 0.9981 Merlo et al. (1995)
NLP 0.0076 1.7260 0.3813 0.9959 0.9695 1.5152
GS 0.0000 1.7330 0.3561 0.9981 0.4583 0.9832
S17
NL − 0.6213 1.1400 0.3704 0.9994 0.0739 0.9995 Merlo et al. (1995)
NLP 0.0000 0.8491 0.4079 0.9987 0.1619 0.9196
GS 0.3767 0.4160 0.4083 0.9987 0.1619 0.7949
S19
NL − 0.0906 1.1371 0.4393 0.9964 1.3320 0.9958 Blick (1992)
NLP 0.0000 1.1170 0.4414 0.9964 1.3380 0.9930
GS 1.4701 0.6234 0.5203 0.9927 2.2244 1.0398
S20
NL − 3.7262 5.4440 0.3222 0.9982 2.4130 0.9985 Blick (1992)
NLP 0.0000 3.6872 0.3683 0.9974 3.3610 0.9890
GS 0.0000 3.5776 0.3739 0.9975 3.7578 1.0204
S21
NL − 6.0980 9.2885 0.3208 0.9988 4.7160 0.9996 Blick (1992)
NLP 0.0003 6.3864 0.3650 0.9980 7.2000 0.9900
GS 0.0000 6.1803 0.3712 0.9987 8.4888 1.0262
(NL) is the application of non-linear regression, (NLP) is the application of non-linear regression with the imposition of τy > 0 and (GS) is the
application of the new technique.

coefficients and the best index values. The suitability of Table 4 are given for the various approaches. Similar
the proposed scheme is further demonstrated in Figs. 8– graphs produced for the rest of the samples have further
12, where the rheograms for some of the samples of demonstrated the suitability of the proposed technique.

Fig. 8. Comparison of rheograms of original data with results from Fig. 9. Comparison of rheograms of original data with results from
rheological models for bentonite–lignite–water suspensions (sample rheological models for bentonite–lignite–water suspensions (sample
S11). S12).
212 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

Fig. 12. Comparison of rheograms of original data with results from


Fig. 10. Comparison of rheograms of original data with results from
rheological models for drilling mud (sample S21).
rheological models for drilling mud (sample S15).

rheological model is not an easy task. There will be


From the data reported in Table 4 it is evident that cases where application of non-linear regression with
the values of the flow behavior index, determined available numerical packages can give estimates of the
with the three different methods, do not differ three rheological parameters with high degree of
significantly except for sample S12. The main confidence. Caution, however, should be exercised
difference is the yield point which can affect both when these techniques give meaningless negative
cuttings carrying capacity of drilling fluids and yield stress values. Imposition of the condition that
pressure drop. The results of Tables 3 and 4 show yield stress is positive may lead to significant differ-
that when the fluid shows a positive yield point, ences of the estimated rheological parameters and the
when determined by non-linear regression, the GS choice of an inappropriate rheological model, other than
method determines quite different rheological para- the Herschel–Bulkley model which might have been
meters with the larger difference observed on the derived if determination of the rheological parameters is
value of the flow behavior index. When the fluid performed following the Golden Section search meth-
shows negative value of the yield point, when derived odology proposed here.
by non-linear regression, the only difference derived It has been demonstrated that the proposed
when using the GS method is on the yield point technique, which relies on the proper choice of the
itself. yield stress using the Golden Section search methodol-
ogy and on the minimization of the sum of squared
3.4. Discussion errors on the linearized form of the Herschel–Bulkley
rheological equation, can lead to meaningful and
The results presented above have indicated that appropriate values of all three rheological parameters
optimal determination of the rheological parameters of with high degree of confidence. The technique
drilling fluids described by the Herschel–Bulkley eliminates the ambiguity and tediousness in finding
the correct parameter set, evident by previous used
methodologies.
Data fitting with the appropriate model and the
application of the most applicable technique always
poses the question about a relevant index of “correct-
ness”. There are concerns over real differences between
parameters derived with curve fitting giving similar
correlation coefficients, as for example, differences
between Rc2 = 0.97 and Rc2 = 0.98. The values in Tables 3
and 4 indicate that the differences in Herschel–Bulkley
parameters can sometimes be very subtle and sometimes
very large, but the impact on the main parameters of
interest, pressure drop and velocity profiles is not really
Fig. 11. Comparison of rheograms of original data with results from known nor it has been investigated thoroughly in the
rheological models for drilling mud (sample S19). past.
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 213

4. Impact on pressure drop and velocity profile of


rheological parameter estimation by different
techniques for drilling operations

4.1. Theoretical considerations

The rheological parameters determined by the three


techniques described above, non-linear regression, non-
linear regression with imposition of penalty for negative
yield stress and the new methodology utilizing the
Golden Section search methodology, have been utilized
to determine pressure drop and velocity profiles in
typical oil-well drilling situations. Two pipe geometries
are used, a pipe with internal diameter of 0.0883 m and a
pipe with internal diameter of 0.1264 m. In addition,
three annulus geometries are used, an annulus with an Fig. 13. Pressure drop-flow rate graph for the three fluids, with
internal diameter of the outer pipe of 0.4445 m and an rheological parameters determined by Golden Section (GS) and by
non-linear regression with τy > 0 (NLP) in a 0.311 m by 0.127 m
external diameter of the inner pipe of 0.127 m, an
concentric annulus. Laminar flow computations.
annulus with an internal diameter of the outer pipe of
0.311 m and an external diameter of the inner pipe of
0.127 m and an annulus with an internal diameter of the the correlation coefficient for NLP is a low 0.9160
outer pipe of 0.216 m and an external diameter of the (Table 4) while for method GS is 0.9942, hence the
inner pipe of 0.089 m. Results have been derived for a variation observed in Fig. 13 may be partly explained
range of flow rates encountered in oil-well drilling by the bad fit of the rheogram when using method
situations, but keeping the flow laminar, which is the NLP. For sample S17, the variation in the predicted
normal condition encountered in drilling applications, pressure drop values with the rheological parameters
particularly for flow in annulus. The appropriate flow derived by methods GS and NLP is between 1.5 to 2.5
equations for Herschel–Bulkley fluids are summarized times. For this case, however, the correlation coeffi-
in Appendix A, both for pipe flow and for concentric cient for the estimation of the rheological parameters
annulus, utilizing slot flow solutions for the latter, a is a high 0.9987 for both methods (Table 4). For
common approach in drilling hydraulics, particularly for sample S19, for which the rheological parameters
annulus diameter ratios greater than 0.3 (Bourgoyne et were determined with GS have a correlation coeffi-
al., 1991; Maglione and Romagnoli, 1999). cient of 0.9927 and with NLP have a correlation
Computations of pressure drop-flow rate data sets coefficient of 0.9964 (Table 4), the difference in the
have been performed for all samples with rheological pressure drop predictions is smaller than the previous
data presented in Tables B1–B3. Results, however, are cases but it is still significant ranging from 10% at
presented here for samples S12, S17 and S19, for which high flow rates to 50% at low flow rates.
negative yield values have been estimated as well as for The computed velocity profiles for the 0.311 m by
some of the samples giving positive yield stress values 0.127 m concentric annulus at various flow rates for
(samples S7, S9, S10 and S18). The rheological sample S19, the sample that showed the smallest
parameters were determined with the new methodology variation in pressure drop values, are shown in Fig.
applying Golden Section search (GS) and with non- 14. Small variations on velocity profiles are observed
linear regression with penalty for τy > 0 (NLP) and are with larger variations in plug velocities, particularly at
listed in Tables 3 and 4. high flow rates. Larger variations are expected and have
In Fig. 13, data is presented for the pressure drop been determined for the other fluid samples for which
variation with the flow rate, for flow in the 0.311 m larger pressure drop variations have been observed
by 0.127 m concentric annulus. For sample S12, there when using rheological parameters derived with the two
is a 2 to 3 times variation between the computed techniques.
pressure drop using rheological parameters derived Fig. 15 shows the results of the pressure drop
with the GS method when compared with the values computations for the smaller annulus and similar trends
computed when the rheological parameters were are observed as in Fig. 13. For sample S12, the variation
obtained according to method NLP. For this sample, between the pressure drop values computed with
214 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

Fig. 14. Velocity profiles for fluid S19 with rheological parameters Fig. 16. Velocity profiles for fluid S12 with rheological parameters
determined by Golden Section (GS) and by non-linear regression with determined by Golden Section (GS) and non-linear regression with
penalty (NLP) in the 0.311 m by 0.127 m concentric annulus. Laminar penalty (NLP), for the 0.216 m by 0.089 m concentric annulus, for five
flow computations for five flow rates (379 l/min, 1136 l/min, 1893 l/ flow rates.
min, 3028 l/min and 3785 l/min).

For pipes, the results for pressure drop estimation are


rheological parameters obtained by methods GS and shown in Fig. 17 for the pipe size of 0.1264 m and Fig.
NLP, is two to three times. For sample S17, this 18 for the pipe size of 0.0883 m for samples S12, S17
variation is between 1.5 to 2 times, while for sample S19 and S19, for laminar flow. The variation observed is
there are no significant differences for flow rates higher similar to the presented cases for the concentric annulus,
than 600 l/min. The velocity profiles for sample S12, the with significant variation for samples S12 and S17 for
sample that showed the largest variation in the estimated both pipe sizes, while the variation is very small for
pressure drop values, are shown in Fig. 16, and sample S19 for both pipe sizes.
differences are observed both in velocity profiles but Similar computations, for the samples for which the
also in maximum (plug) velocities for the computed yield stress was found positive and with no significant
values using the rheological parameters from the two differences among the model parameters listed in Table
cases. 3, have revealed that the differences in pressure drop

Fig. 15. Pressure drop-flow rate graph for three fluids, with rheological Fig. 17. Pressure drop-flow rate graph for three fluids, with rheological
parameters determined by Golden Section (GS) and by non-linear parameters determined by Golden Section (GS) and by non-linear
regression with penalty (NLP), for the 0.216 m by 0.089 m concentric regression with penalty (NLP), for the 0.1264 m pipe. Laminar flow
annulus. Laminar flow computations. computations.
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 215

important for pressure drop estimation because of the


small margin for frictional pressure drop normally
allowed for flow in the annulus. It is equally important
for the estimation of velocity profiles which affect
cuttings transport and intermixing of fluids in the
annulus, where laminar flow conditions normally
prevail. The analysis further demonstrates that selec-
tion of rheological parameters based on the goodness
of fit of the rheogram, the correlation coefficient or the
best index value may not always be sufficient.
Estimation of the impact of the choice of the
rheological parameters by an appropriate methodology
should be made by computing pressure drop and
velocity profiles for typical drilling situations. This
approach can be further extended and refined in order
Fig. 18. Pressure drop-flow rate graph for three fluids, with rheological to derive an ‘index of appropriateness’ for estimating
parameters determined by Golden Section (GS) and by non-linear rheological parameters.
regression with penalty (NLP), for the 0.0883 m pipe. Laminar flow
computations.
4.2. Implications on drilling operations

computation are much smaller in this case while the The implications on drilling from use of different
velocity profiles for the cases of these samples do not rheological parameters but from same rheological data
vary much. Some of the results for the pressure drop set have been assessed with field data of Merlo et al.
estimation are shown in Fig. 19 for samples S7, S9, S10 (1995). An analysis is presented for the circulation
and S18. test performed in well A at the depths of 555 m and
The presented results demonstrate that differences in 2008 m with the flow geometry shown in Fig. 20.
computed pressure drop values and velocity profiles The circulation test at 555 m depth is deemed as the
exist when computations are performed using either the most reliable in terms of data analysis because the
rheological parameters obtained by GS (the most drill pipe was located inside the casing with a
correct one) or by NLP. The variations can be rather nominal diameter of 0.508 m (inside diameter of
large for typical drilling situations not only on pressure 0.4826 m) with a well-known geometry and because
drop but also on velocity profiles. This is particularly the circulating drilling fluid is not much affected by

Fig. 19. Pressure drop-flow rate graph for fluids S7, S9, S10 and S18,
with rheological parameters determined by Golden Section (GS) and Fig. 20. Geometry of well A for the reported drilling circulation test at
by non-linear regression (NL), for the 0.311 m by 0.127 m concentric 3200 l/min. Relevant geometrical data are given by Merlo et al.
annulus. Laminar flow computations. (1995).
216 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

the temperature of the well. For purposes of


simulation, the S13 sample was considered, for
which the rheological parameters have been deter-
mined by using both non-linear regression (NL) and
Golden Section method (GS), which exhibit small
differences in the rheological parameters but very
similar correlation coefficients, sum of square errors
and best index values (Table 3).
The calculated value of pressure drop in the drilling
hydraulic circuit (excluding the drilling bit), using the
methodology of Merlo et al. (1995), is estimated as
44.000 Pa for the NL fluid and about 37.800 Pa for
the GS fluid. Measured field data gave a value of Fig. 22. Velocity profiles of NL and GS fluids for flow of the S13
drilling fluid in a 0.483 m by 0.127 m concentric annulus, for a flow
29.100 Pa (Merlo et al., 1995). From the above data it
rate of 3202 lpm.
can be seen that the calculated pressure drop in the
drilling circuit is 29.9% and 51.2% more than the field
measured value when using the GS fluid or the NL flow rate of 3202 lpm, the results show similar velocity
fluid respectively. Furthermore, the pressure drop profile in the annular section (Fig. 22) meaning that both
calculated by using the GS fluid is less than the fluids can have the same cuttings carrying capacity.
pressure drop calculated with the NL fluid by an Cuttings transport efficiencies, calculated as per Walker
amount equal to 16.4% of the GS fluid value, and Mayes (1975) but extended for Herschel–Bulkley
indicating that the GS fluid performs better than the fluids by Gallino and Maglione (1996), are estimated at
NL fluid, in terms of pressure drop. This holds even 78.9% and 77.5% for the NL fluid and GS fluid
though both fluids have the same formulation and respectively.
characteristics and the viscometer data are very well Finally, considering the apparent viscosity of the
described by both methods ( R c2 = 0.9971 and drilling fluid at the nozzles of the drilling bit, μa,
Rc2 = 0.9952 for the NL fluid and the GS fluid which affects the rate of penetration among other
respectively). In Fig. 21, the variation of the calculated parameters, differences in the behavior of the NL and
frictional pressure loss in the drilling hydraulic circuit GS fluids are expected which can be evaluated. The
is plotted against the flow rate for the NL fluid and apparent viscosity at the drilling bit can be
GS fluid, together with measured data. computed, following standard drilling hydraulics
When the flow velocity distribution in the annulus procedures (Bourgoyne et al., 1991; Maglione and
(0.483 m by 0.127 m) is computed for the two fluids at a Romagnoli, 1999), by estimating the wall shear
stress, τw, and the wall shear rate, γw, of the drilling
fluid at the bit nozzles, and computing the apparent
viscosity, μa, as,

sw
la ¼ ð18Þ
γw

The apparent viscosity is estimated at a value of


about 1.707 cP for the NL fluid and at a value of about
1.261 cP for the GS fluid for the flow rate of 3202 lpm.
Fig. 23 shows the behavior of apparent viscosity at the
drilling bit nozzles for the NL and GS fluids, for
various flow rates. The results show that the apparent
viscosity of the GS fluid has always lower values than
the values of the NL fluid with the differences varying
between 18 and 27%, for the range of the considered
Fig. 21. Pressure drop-flow rate curves for drilling circulation test for
Well A, using NL and GS derived rheological parameters for drilling
flow rates.
fluid sample S13. The measured data at the one flow rate (3202 lpm) is The rate of penetration during drilling (ROP) is
also shown. directly affected by fluid density (ρ), fluid velocity in the
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 217

fluid, with the only difference in the approach


followed to determine the best fit rheological para-
meters from viscometer data. The variation of the
rheological parameters affects mostly frictional pres-
sure drop and apparent viscosity at the bit. Hence, at
the design phase and during oil-well drilling opera-
tions, caution should be exercised when using the
rheological parameters from the best fit curves of the
viscometer data, because, as it has been shown, very
small differences at the design phase can lead
sometimes to substantial differences in most of
operational parameters.
Fig. 23. Variation of apparent viscosity, at the drilling bit, with the flow
rate, for the NL and GS fluids, for sample S13 and Well A (concentric
5. Conclusions
annulus of 0.4826 m by 0.127 m).
In this study it was shown that the Herschel–Bulkley
rheological model properly described all rheological
drilling bit nozzles (Vn), bit nozzle diameter (dn), while data of drilling fluids obtained with rotational visc-
it is inversely proportional to drilling fluid apparent ometer data. It has been demonstrated that the three
viscosity at the drilling bit (μa,bit) (Eckel, 1967; Beck et rheological Herschel–Bulkley parameters can be
al., 1995; Gallino et al., 1996). This variation is derived with a numerical package using non-linear
described well by Eq. (19), regression but the procedure may not always lead to
optimal solutions, because sometimes, meaningless
!c negative yield stress values are determined. Imposition
qVn dn
ROP~ ð19Þ of the condition for positive values of the yield stress
la;bit gives non-optimal solutions. A methodology has been
proposed to alleviate this problem. It estimates the best
where c is a constant, which depends on the drilled value for the yield stress using the Golden Section
formation. From Eq. (19) one can notice that, keeping search methodology, and then applies linear regression
all other parameters constant, the rate of penetration is to the transformed data, yielding as accurate results as
inversely proportional to the apparent viscosity at the the numerical schemes in normal cases but also giving
bit. Considering for instance the case for well A positive values for the yield stress in situations where
reported above (Merlo et al., 1995), if a value for the numerical schemes determine negative values. The
rate of penetration of 10 m/h is assumed for the NL recommended approach leads to unique solutions and
fluid, the corresponding rate of penetration for the GS can be easily implemented.
fluid would be predicted around 12.6 m/h, an increase It was further shown that pressure drop and
of 26%. This particular example reveals that just by velocity profiles for laminar flow in pipes and
considering and studying better the best fit curves of concentric annulus can be significantly affected by
viscometer data of drilling fluids, a surprising result proper choice of rheological parameters. The most
with respect to the rate of penetration can be derived. appropriate set should be determined utilizing not only
This significant increase has been computed using the statistically best fit indices but also the impact on
rheological parameters that were very close to each pressure drop and velocity profiles. Use of the
other (Table 3, sample S13). The results could be relevant flow equations for laminar flow in typical
more dramatic, if larger differences occur among oil-well drilling situations in concentric annulus and
rheological parameters derived using the two different pipe can aid significantly in determining the impact of
techniques. the particular choice of rheological parameters on
All the above results could be considered as the pressure drop, on velocity profiles and on rate of
direct effect of the decrease of the flow behavior penetration, aiding in choosing the most appropriate
index of the drilling fluid when determined using the rheological parameters. The computed results have
GS method and its consequences on hydraulic demonstrated that it is very important to make the best
parameters. It seems that the GS fluid performs better simulation of rheological behavior of drilling fluids
than the NL fluid, even though they are the same before computing hydraulic parameters. Once the best
218 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

rheological parameters are obtained, utilization of Subscripts


appropriate calculation tools for estimation of pressure min minimum value
drop and velocity profiles can aid in avoiding many max maximum value
problems during drilling operations and can allow o initial value
better exploitation of the properties of existing drilling w wall
fluids. a apparent value
a, bit at the bit
Nomenclature
A, B, C Robertson–Stiff rheological parameters Appendix A. Flow of Herschel–Bulkley drilling
BIV Best index value fluids in pipes and concentric annuli
dp/dL Pressure drop per unit length (M/L2T2)
dn Bit nozzle diameter (L) A.1. Flow in pipes
h Gap of annulus, distance between plates in the
slot (L) The geometry for pipe flow together with the
K Fluid consistency index (M/LT2−n) appropriate parameters is shown in Fig. A1. There is a
L Lower limit for the estimation of yield stress central core of the fluid which moves as a rigid plug if
n Fluid behavior index the shear stress levels are smaller than the yield stress of
p Pressure (M/LT2) the fluid. Letting Δ = dp/dL, with dp/dL the pressure
r Radius (L) drop per unit length, and for values of the shear stress τ
R pipe radius (L) greater than the yield stress of the fluid τy, τ ≥ τy,
R1 radius of outer pipe of annulus (L) balance of forces gives,
R2 radius of inner pipe of annulus (L)
Rc2 dðsrÞ dp
Correlation coefficient ¼r ¼ rD ðA1Þ
ROP Rate of penetration (L/T) dr dL
q Flow rate (L3/T) Eq. (A1) upon integration gives,
Q2 Sum of error squares
tol Half width of search interval r C1
s¼ Dþ ðA2Þ
u Liquid velocity (L/T) 2 r
U Upper limit for the estimation of yield stress The shear stress is finite at r = 0, hence, C1 = 0. The
Vn Liquid velocity at the drill bit nozzle (L/T) radius at which there is an unsheared portion of the fluid,
ya Distance of top of inner layer from bottom rp, is given by
plate (L)
yb Distance of bottom of top layer from bottom rp
plate (L) sy ¼ D ðA3Þ
2
yi Measured parameter
ŷi Predicted parameter and the wall shear stress, τw, is given by,
ȳ Average value of the parameter
w Width of the slot (L) R
sw ¼ D ðA4Þ
2
Greek letters
γ Shear rate (T− 1)
γ̄ Geometric mean shear rate (T− 1)
Δ pressure drop per unit length (M/L2T2)
μ Liquid viscosity (M/LT)
ξ Dimensionless parameter, ratio of yield stress
to wall shear stress
ρ Liquid density (M/L3)
τ Shear stress (M/LT2)
τ̄ Geometric mean shear stress (M/LT2)
τy Yield stress (M/LT2)
τy0 Initial estimate of yield stress (M/LT2)
τo Integer constant Fig. A1. Geometry and parameters for laminar flow in pipes.
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 219

where R is the pipe radius. The flow curve of the


Herschel–Bulkley fluid is given by,
 
du n
s ¼ sy þ K − ðA5Þ
dr

where K, n are the fluid consistency and flow behavior


indices respectively. Hence,

 n Fig. A2. Representation of the concentric annulus as a slot.


du r
sy þ K − ¼ D ðA6Þ
dr 2
A.2. Flow in concentric annuli
The solution of Eq. (A6), utilizing the boundary
condition that the fluid velocity u is zero at the pipe For flow in a concentric annulus, the annulus can be
wall, u = 0 at r = R, is finally given by, approximated by a slot of gap h = R2 − R1, as represented
in Fig. A2. The solution that will be derived follows the
( mþ1   ) procedure for flow of Bingham plastic fluids in a slot,
1 1 D sy D sy mþ1 given by Bourgoyne et al. (1991).
u¼ R− − r− ;
D=ð2KÞ m þ 1 2K K 2K K There is a central core of the fluid which moves as a
sy rigid plug if the shear stress levels are smaller than the
rp ¼ VrVR yield stress of the fluid. Integration of the force balance
D=2
ðA7aÞ performed on a fluid element gives

and s ¼ yD þ s0 ðA9Þ

 
ðD=2KÞm sy mþ1 with τ0 an integration constant to be determined. Let ya,
u ¼ up ¼ R− ; 0VrVrp ðA7bÞ yb the distances of the lower sheared and upper sheared
mþ1 D=2
surfaces from the bottom plate respectively. At the inner
layer of the plug, ya, the shear stress τa must equal
with m = 1/n.
(− τy). It follows then from (A9),
The flow rate q can be derived as,
Z R
sa ¼ −sy ¼ s0 þ ya D ðA10Þ
q ¼ 2k urdr
0
kn ðDR=2−sy Þ1=nþ1 which gives,
¼
K 1=n ðD=2Þ3
" #
ðDR=2−sy Þ2 2sy ðDR=2−sy Þ s2y sy þ s0
 þ þ ya ¼ − ðA11Þ
1 þ 3n 1 þ 2n 1þn D

ðA8Þ Similarly, for the outer plug region one obtains for
the shear stress at yb,
Eq. (A8) relates pressure drop (Δ = dp/dL) with
flow rate (q) for flow of Herschel–Bulkley drilling sb ¼ sy ¼ s0 þ yb D ðA12Þ
fluids in a pipe of radius R for laminar flow. If the
pressure drop is known, the flow rate can be directly In the fluid layer enclosed by the inner layer of the
computed. On the other hand, if the flow rate is plug and the bottom plate, the shear stress is,
known, Eq. (A8) can be solved by trial-and-error for
the estimation of pressure drop. The velocity profiles  n
can be computed readily from Eqs. (A7a) and (A7b) du
s ¼ −sy −K ¼ s0 þ yD ðA13Þ
for both cases. dy
220 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

The solution to differential Eq. (A13), utilizing The plug velocity, given by Eqs. (A16) and (A20) is
boundary condition u = 0 at y = 0 is finally given by, the same, hence,
(     ) ymþ1 ¼ ðh−yb Þmþ1 ðA21Þ
−K sy þ s0 mþ1 sy þ s0 D mþ1 a
u¼ − − þ − − y ;
ðm þ 1ÞD K K K
0VyVya ðA14Þ Taking the (m + 1)th root and keeping only the
positive value, it follows that,

with m = 1/n. In terms of the parameter ya this equation ya ¼ h−yb Z ya þ yb ¼ h ðA22Þ


becomes
Substituting the values of ya, yb, given above, the
unknown τ0 can now be determined as,
ðD=KÞ n m o
u¼− −ðya Þmþ1 þ ðya −yÞmþ1 ; 0VyVya
ðm þ 1Þ h h dp
s0 ¼ − D ¼ − ðA23Þ
ðA15Þ 2 2 dL

Furthermore, the following holds,


The plug velocity, up, is the velocity at y = ya, given
h dp
by, sw ¼ −s0 ¼ ðA24Þ
2 dL
 m
ymþ1 D The summary of the equations governing flow of
up ¼ a
; ya VyVyb ðA16Þ Herschel–Bulkley drilling fluids in a slot is given below.
ðm þ 1Þ K
There is fluid flow only if,
For the fluid region enclosing the plug and the upper dp sy
plate, the shear stress is D¼ > ðA25Þ
dL h=2
  When there is flow, then,
du n
s ¼ sy þ K − ¼ s0 þ yD ðA17Þ
dy h h dp
s0 ¼ − D ¼ − ¼ −sw ðA26Þ
2 2 dL
and following similar approach as above, utilizing the
boundary condition u = 0 at y = h, the velocity is given h sy
by, ya ¼ − ðA27Þ
2 D
 h sy
1 yb ¼ þ ðA28Þ
u¼ ðs0 −sy þ hDÞmþ1 2 D
DK m ðm þ 1Þ

mþ1
−ðs0 −sy þ yDÞ ; yb VyVh ðA18Þ The velocity profile is given by,

ðD=KÞm n o
u¼− −ðya Þmþ1 þ ðya −yÞmþ1 0VyVya
In terms of yb, Eq. (A18) becomes, ðm þ 1Þ
 m   ðA29aÞ
D 1  m
u¼ ðh−yb Þmþ1 −ðy−yb Þmþ1 ; yb VyVh ymþ1 D
K ðm þ 1Þ u¼ a
ya VyVyb ðA29bÞ
ðA19Þ ðm þ 1Þ K

 m  
The velocity of the plug is given when y = yb, hence, D 1 mþ1 mþ1
u¼ ðh−yb Þ −ðy−yb Þ ; yb VyVh
K ðm þ 1Þ
 m
D ðh−yb Þmþ1 ðA29cÞ
up ¼ ; ya VyVyb ðA20Þ
K ðm þ 1Þ with m = 1/n.
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 221

The flow rate per unit width of the slot, w, is given If now ξ is defined as,
by,
sy 2sy sy
n¼− ¼ ¼ ðA32Þ
Z h Z ya Z yb Z h s0 hD sw
q
¼ udy ¼ udy þ up dy þ udy
w 0 0 ya yb
and noting that for partial or full flow it must be true that
which can be written as, ξ < 1, the flow rate is,

h Z h Z h  
q  du du  m

¼ uy − y dy ¼ 0 − y dy D 2wðh=2Þmþ2 ð1−nÞmþ1
w dy dy q¼ ½n þ ðm þ 1Þ
Z0 ya 0  Z h 0  K ðm þ 1Þðm þ 2Þ
du du
¼− y dy − y dy ¼ −I1 −I2 ðA33Þ
0 dy yb dy

An equivalent expression has been derived by


Grinchik and Kim (1974) as referenced by Fordham et
With some mathematical manipulation, it can be
al. (1991), where though, only the final result was given.
shown that,
The flow rate also can be expressed in terms of dp/dL
and τy as,
 m
D ymþ2
I1 ¼ a

K ðm þ 1Þðm þ 2Þ dp 1=n  1þ1=n
2wðh=2Þ2þ1=n 1
sy
dL K
q¼ 1−
and ð1=n þ 1Þð1=n þ 2Þ ðh=2Þðdp=dLÞ
 
sy 1
 þ þ1
ðh=2Þðdp=dLÞ n ðA34Þ
 m " #
D ðh−yb Þmþ2 yb ðh−yb Þmþ1
I2 ¼ − þ
K mþ2 mþ1 The final result for the flow rate in terms of annulus
geometry parameters, noting that, wh = π(R22 − R12), and
h = R2 − R1, is,
So finally,

 m 2  1=n 3
D 1 dp
q 1 kðR22 −R21 ÞðR2 −R1 Þ1þ1=n
¼ 6 K dL 7
w K ðm þ 1Þðm þ 2Þ q¼4 5
2 1=n
ð1=n þ 1Þð2=n þ 4Þ
 f−ymþ2 þ ðm þ 1Þðh−yb Þmþ2
a
 1þ1=n
þ ðm þ 2Þðyb Þðh−yb Þmþ1 g ðA30Þ sy
 1−
½ðR2 −R1 Þ=2ðdp=dLÞ
" sy #
The flow rate can be expressed in terms of parameters ½ðR −R Þðdp=dLÞ
þ 1n þ 1
 2 1

τy, Δ by replacing ya, yb and noting that τ0 = − hΔ/2, so 2 1=n


ð1=n þ 1Þð2=n þ 4Þ ðA35Þ
that,

 m      Eq. (A35) provides the relationship between flow


D w sy þ s0 mþ2 rate and pressure drop for Herschel–Bulkley drilling
q¼  − −
K ðm þ 1Þðm þ 2Þ D fluids flowing in laminar flow in concentric annuli,
 s −s mþ2 modeled as a slot. The flow rate can be directly
y 0
þðm þ 1Þ h−
D determined given the pressure drop but trial-and-error
s −s  s −s mþ1

þðm þ 2Þ
y 0
h−
y 0
ðA31Þ solution is required if the pressure drop is to be
D D determined for a given flow rate.
222
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224
Appendix B. Rheological data

Table B1. Rheological data for bentonite–water suspensions and bentonite–lignite–water suspensions (A: bentonite suspension, L: lignite) (from Kelessidis
et al., 2005).

Sample number S1 S2 S3 S4 S5 S6 S7 S8 S9 S10 S11 S12


Sample source A hydrated A aged A + L1, 3.0% A + L2 3.0% A + L4 3.0% A + L5 3.0% A + L6 3.0% A + L7 0.5% A + L7 0.5% A + L4 3.0% A + L8 0.5% A + L3 3.0%
25 °C 25 °C aged 25 °C aged 25 °C aged 25 °C aged 25 °C aged 25 °C hydrated 25 °C hydrated 65 °C hydrated 65 °C hydrated 25 °C aged 25 °C
Shear rate Shear Shear Shear Shear Shear Shear Shear Shear Shear Shear Shear Shear
stress stress stress stress stress stress stress stress stress stress stress stress
(1/s) (Pa) (Pa) (Pa) (Pa) (Pa) (Pa) (Pa) (Pa) (Pa) (Pa) (Pa) (Pa)
1021.38 28.50 48.42 21.92 6.92 10.25 20.75 18.17 11.75 12.83 4.50 9.33 12.17
851.15 27.75 46.83 22.75 6.17 8.92 19.25 16.67 11.08 9.75 4.17 8.58 11.00
680.92 26.83 44.58 26.08 5.58 7.83 17.92 14.75 9.75 8.58 3.58 7.92 9.83
510.69 25.00 41.58 22.42 4.75 5.92 17.08 12.75 8.58 8.00 3.08 6.83 8.33
340.46 22.58 37.83 18.58 3.67 4.75 15.17 9.92 6.58 7.33 2.42 5.58 6.83
170.23 21.00 35.67 13.58 2.17 3.42 12.33 6.50 4.25 5.58 1.75 3.42 4.00
136.18 21.00 33.25 12.00 1.83 3.08 11.67 6.17 4.08 5.42 1.50 2.92 3.42
102.14 20.08 30.83 10.50 1.33 2.92 10.75 5.75 3.75 5.08 1.42 2.42 2.75
51.07 16.67 27.42 6.92 0.58 2.00 9.00 4.00 2.58 4.00 1.08 1.50 1.58
34.05 17.50 27.83 7.42 0.50 1.67 8.83 4.17 2.42 4.17 1.00 1.33 1.42
17.02 16.00 21.90 6.67 0.08 1.25 8.08 3.33 2.08 3.67 0.67 1.00 1.17
10.21 14.67 23.75 6.25 0.00 1.00 8.17 3.25 1.83 3.67 0.58 0.58 0.92
5.11 13.20 19.90 5.10 0.00 0.50 7.70 2.80 1.60 3.25 0.50 0.50 0.10
V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224 223

Table B2: Rheological data of drilling fluids (from Merlo et al., 1995)

Sample S13 S14 S15 S16 S17 S18


number
Sample source Mud 20 °C 0.1 MPa Mud 30 °C 0.1 MPa Mud 45 °C 0.2 MPa Mud 85 °C 0.5 MPa Mud 100 °C 1 MPa Mud 20 °C 2008 m
Shear rate Shear stress Shear stress Shear stress Shear stress Shear stress Shear stress
(1/s) (Pa) (Pa) (Pa) (Pa) (Pa) (Pa)
1021.32 26.6 22.0 20.4 16.4 14.3 27.1
510.66 19.4 16.9 15.8 12.3 10.7 17.4
340.44 16.9 14.8 13.8 10.2 9.2 14.8
170.22 13.3 11.8 11.2 8.2 7.2 10.2
10.21 5.6 4.6 4.1 2.6 2.0 4.1
5.11 3.6 3.1 2.6 2.0 1.5 3.1

Table B3: Rheological data of drilling fluids (from Blick, 1992, as reported by Al-Zahrani, 1997)

Sample number S19 S20 S21


Sample source 10% bentonite 12% bentonite 28% bentonite
Shear rate Shear stress Shear stress Shear stress
(1/s) (Pa) (Pa) (Pa)
1020.80 23.46 46.44 78.52
765.60 21.07 42.61 72.78
510.40 17.72 37.35 63.20
340.27 14.84 32.56 55.06
238.19 12.93 28.25 47.40
153.12 10.05 23.46 40.22
119.09 8.62 21.55 36.39
85.07 8.14 18.67 32.56
51.04 5.75 14.84 25.86
17.01 4.31 10.53 17.72

References fluid applications. Paper SPE 84215 Presented at the SPE


Productions and Operations Symposium, Oklahoma City, OK.
Benna, M., Kbir-Ariguib, N., Magnin, A., Bergaya, F., 1999. Effect of
Alderman, N.J., Meeten, G.H., Sherwood, J.D., 1991. Vane rheometry pH on rheological properties of purified sodium bentonite
of bentonite gels. J. Non-Newton. Fluid Mech. 39, 291–310. suspensions. J. Colloid Interface Sci. 218, 442–455.
Al-Zahrani, S.M., 1997. A generalized rheological model for shear Bingham, E.C., 1922. Fluidity and Plasticity. McGraw-Hill, New
thinning fluids. J. Pet. Sci. Eng. 17, 211–215. York.
American Petroleum Institute Specifications 13I, 2000. Recommended Bird, R.B., Dai, G.C., Yarusso, B.Y., 1982. The rheology and flow of
Practice Standard Procedure for Laboratory Testing Drilling viscoplastic materials. Rev. Chem. Eng. 1 (1), 1–70.
Fluids. Bleler, R., 1990. Selecting a drilling fluid. J. Pet. Technol. 42,
Barnes, H.A., 1995. A review of the slip (wall depletion) of polymer 832–834.
solutions, emulsions and particle suspensions in viscometers: its Blick, E.F., 1992. Non-Newtonian Fluid Mechanics Notes, Engineer-
cause, character and cure. J. Non-Newton. Fluid Mech. 56, ing Library. University of Oklahoma, Norman, OK.
221–251. Borgia, A., Spera, F.J., 1990. Error analysis for reduction noisy wide-
Barnes, H.A., 1999. The yield stress — a review or ‘πάντα ρει’ — gap concentric cylinder rheometric data for nonlinear fluids: theory
everything flows? J. Non-Newton. Fluid Mech. 81, 133–188. and applications. J. Rheol. 34 (1), 117–136.
Barnes, H.A., Nguyen, Q.D., 2001. Rotating vane geometry — a Bourgoyne, A.T., Chenevert, M.E., Millheim, K.K., Young Jr., F.S.,
review. J. Non-Newton. Fluid Mech. 98, 1–14. 1991. In: Evers, J.F., Pye, D.S. (Eds.), Applied drilling
Barnes, H.A., Walters, K., 1985. The yield stress myth? Rheol. Acta engineering. SPE Textbook Series, vol. 2. Richardson, TX.
24, 323–326. Briscoe, B.J., Luckham, P.F., Ren, S.R., 1994. The properties of drilling
Beck, F.E., Powell, J.W., Zamora, M., 1995. The effect of rheology on muds at high-pressures and high-temperatures. Philos. Trans. R.
rate of penetration. Paper SPE/IADC 29368 Presented at the SPE/ Soc. Lond. Ser. A: Math. Phys. Sci. 348, 179–207.
IADC Drilling Conference, Amsterdam. Casson, N., 1959. Flow equation for pigment oil suspensions of the
Becker, T.E., Morgan, R.G., Chin, W.C., Griffith, J.E., 2003. Improved printing ink type. In: Mills, C.C. (Ed.), Rheology of Disperse
rheology model and hydraulic analysis for tomorrow's wellbore Systems. Pergamon Press, Oxford.
224 V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203–224

Chen, D., 1986. Yield stress: a time dependent property and how to Maglione, R., Ferrario, G., 1996. Equations determine flow states for
manage it. Rheol. Acta 25, 542–554. yield-pseudoplastic drilling fluids. Oil Gas J. 94, 63–66.
Clark, R.K., 1994. Impact of environmental regulations on drilling Maglione, R., Robotti, G., 1996. A numerical procedure for solving a
fluid technology. J. Pet. Technol. 46, 804–809. non-linear equations systems for determining the three rheological
Darby, R., 1985. Couette viscometer data reduction for materials with parameters of a drilling mud from experimental data. Proceedings
a yield stress. J. Rheol. 29, 369–378. of the Fourth International Conference on Integral Methods in
De Kee, D., Chan Man Fong, C.F., 1993. Letter to the Editor: a true Science and Engineering, Oulu, Finland.
yield stress? J. Rheol. 37, 775–776. Maglione, R., Romagnoli, R., 1998. The role of rheology optimisation
Eckel, J.R., 1967. Microbit studies of the effects of fluid properties and in the drilling mud design. GEAM Bull. 94, 123–132.
hydraulics on drilling rate. J. Pet. Technol. 541–546 (April). Maglione, R., Romagnoli, R., 1999. Idraulica dei Fluidi di
Elward-Berry, J., Darby, J.B., 1992. Rheologically stable, nontoxic, Perforazione. Edizioni Cusl, Torino.
high temperature water-base drilling fluid. Paper SPE 24589 Maglione, R., Ferrario, G., Rrokaj, K., Calderoni, A., 1996. A new
Presented at the 67th Annual Technical Conference and Exhibition constitutive law for the rheological behaviour of non Newtonian
of the Society of Petroleum Engineers, Washington. fluids. Proceedings of the XIIe Congrès International de Rheology,
Evans, I.D., 1992. On the nature of the yield stress. J. Rheol. 36, Quebec-Canada.
1313–1316. Maglione, R., Guarneri, A., Ferrari, G., 1999a. Rheologic and
Fordham, E.J., Bittleston, S.H., Tehrani, M.A., 1991. Viscoplastic flow in hydraulic parameter integration improves drilling operations. Oil
centered annuli, pipes and slots. Ind. Eng. Chem. Res. 29, 517–524. Gas J. 97, 44–48.
Gallino, G., Maglione, R., 1996. L'applicazione del Modello di Maglione, R., Robotti, G., Romagnoli, R., 1999b. Drilling hydraulic
Herschel and Bulkley alle Attuali Teorie per l'Ottimizzazione della optimization: how managing it. GEAM Bull. 96, 23–30.
Pulizia del Foro. Proceedings of the Fourth National Conference Maglione, R., Robotti, G., Romagnoli, R., 2000. In-situ rheological
on Applied Rheology, Vico Equense—Italy. characterization of drilling mud. SPE J. 5, 377–386.
Gallino, G., Guarneri, A., Poli, G., Xiao, L., 1996. Scleroglucan Merlo, A., Maglione, R., Piatti, C., 1995. An innovative model for
biopolymer enhances WBM performances. Paper SPE 36426, drilling fluid hydraulics. Paper SPE 29259 Presented at the Asian-
Presented at the SPE Annual Technical Conference and Exhibition, Pacific Oil and Gas Conf., Kuala Lumpur, Malaysia.
Denver-USA. Miano, F., Rabaioli, M.R., 1994. Rheological scaling of montmor-
Govier, G.W., Aziz, K., 1972. The Flow of Complex Mixtures in illonite suspensions: the effect of electrolytes and polyelectrolytes.
Pipes. Krieger, Malabar, FL. Colloids Surf., A Physicochem. Eng. Asp. 84, 229–237.
Graves, W.G., Collins, R.E., 1978. A new rheological model for non Mihalakis, A., Makri, P., Kelessidis, V.C., Christidis, G., Foscolos, A.,
Newtonian fluids. Paper SPE, vol. 7654. Papanikolaou, K., 2004. Improving rheological and filtration
Gray, H.C.H., Darley, G.R., 1980. Composition and Properties of Oil- properties of drilling muds with addition of Greek lignite.
Well Drilling Fluids. Gulf Pub. Co., Houston. Proceedings of the 7th National Congress on Mechanics, Chania,
Grinchik, I.P., Kim, A.K., 1974. Axial flow of a non-linear viscoplastic Greece.
fluid through cylindrical pipes. J. Eng. Phys. 23, 1039–1041. Mnatsakanov, A.V., Litvinov, A.I., Zadvornykh, V.N., 1991. Hydro-
Gucuyener, I.H., 1983. A rheological model for drilling fluids and dynamics of the drilling in deep, thick, abnormal pressure
cementing slurries. Paper SPE 11487 Presented at the Middle East reservoirs. Paper SPE/IADC 21919 Presented at the Drilling
Oil Technical Conference, Manama-Bahrain. Conference, Amsterdam.
Guillot, D., 1990. Rheology of well cement slurries. In: Nelson, E.B. Nguyen, Q.D., Boger, D.V., 1987. Measuring the flow properties of
(Ed.), Well Cementing. Schlumberger Educational Services, yield stress liquids. Annu. Rev. Fluid Mech. 24, 47–88.
Houston. Ohen, H.A., Blick, E.F., 1990. Golden search method for determina-
Hanks, R.W., 1979. The axial flow of yield — pseudoplastic fluids in a tion of parameters in Robertson–Stiff non-Newtonian fluid model.
concentric annulus. Ind. Eng. Chem. Process Des. Dev. 18, 488–493. J. Pet. Sci. Eng. 4, 309–316.
Hanks, R.W., 1989. Couette viscometry of Casson fluids. J. Rheol. 27, Press, W.H., Flannery, B.P., Teukolsky, S.A., Vetterling, W.T., 1992.
1–6. Numerical Recipes in Fortran, 2nd edition. Cambridge University
Hartnett, J.P., Hu, R.Y.Z., 1989. Technical note: the yield stress—an Press, Cambridge.
engineering reality. J. Rheol. 33, 671–679. Robertson, R.E., Stiff Jr., H.A., 1976. An improved mathematical
Helland, I., 1988. On the structure of partial least squares regression. model for relating shear stress to shear rate in drilling fluids and
Commun. Stat., Simul. Comput. 17, 581–607. cement slurries. SPE J. 16, 31–36.
Hemphil, T., Campos, W., Tehrani, M.A., 1993. Yield power law Schurz, J., 1992. A yield value in a true solution? J. Rheol. 36,
model accurately predicts mud rheology. Oil Gas J. 91, 45–50. 1319–1321.
Herschel, W.H., Bulkley, R., 1926. Konsistenzmessungen von Shulman, Z.P., 1968. On Phenomenological Generalization of
Gummi-Benzollosungen. Kolloid-Z. 39, 291–300. Viscoplastic Rheostable Disperse System Flow Curves, vol. 10.
Kelessidis, V.C., Mihalakis, A., Tsamantaki, C., 2005. Rheology and Teplo-Massoperenos, Minsk.
rheological parameter determination of bentonite–water and Turian, R.M., Ma, T.W., Hsu, F.L.G., Sung, D.J., 1997. Characteriza-
bentonite–lignite–water mixtures at low and high temperatures. tion, settling and rheology of concentrated fine particulate mineral
Proceedings of the 7th World Congress of Chem. Engr., Glasgow. slurries. Powder Technol. 93, 219–233.
Krieger, I.M., 1968. Shear rate in the Couette viscometer. Trans. Soc. Walker, R.E., Mayes, T.M., 1975. Design of muds for carrying
Rheol. 12, 5–11. capacity. Paper SPE 4975, J. Petr. Techn. 27, 893-900.
Maglione, R., 1999. The drilling well as viscometer: the route towards Yeow, Y.L., Ko, W.C., Tang, P.P.P., 2000. Solving the inverse problem
new drilling frontiers. Proceedings of the Southern Europe of Couette viscometry by Tikhonov regularization. J. Rheol. 44,
Conference in Rheology, Sangineto, Italy. 1335–1351.

You might also like