Sei sulla pagina 1di 22

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

www.elsevier.com/locate/petrol

Optimal determination of rheological parameters for


HerschelBulkley 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 bentonitelignite as additives exhibit non-Newtonian rheological behavior which can
be described well by the three parameter HerschelBulkley 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; HerschelBulkley; Pressure drop; Velocity profile; Penetration rates

1. Introduction
In oil-well drilling, bentonite is added in drilling
fluids for viscosity control, to aid the transfer of cuttings
from the bottom of the well to the surface, and for
filtration control to prevent filtration of drilling fluids
into the pores of productive formations. It is long known
Corresponding author.
E-mail address: kelesidi@mred.tuc.gr (V.C. Kelessidis).
0920-4105/$ - see front matter 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.petrol.2006.06.004

that above around 120 C and in conditions of high


salinity, bentonite slurries begin to thicken catastrophically (Gray and Darley, 1980; Bleler, 1990; ElwardBerry and Darby, 1992). Attempts to describe and to
predict the gelling tendencies of bentonite suspensions
have not yet resulted in a concise method which could
predict rheological and filtration properties, given the
amount of added bentonite and its physical characteristics. The flocculation of bentonite suspensions at high
temperatures could be resolved with the addition of

204

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

thinners to reduce the rheology of the mixture but many


thinners degrade over the same temperature range. A
thinner with high thermal stability is lignite (Clark,
1994; Briscoe et al., 1994; Miano and Rabaioli, 1994)
and recent evidence (Mihalakis et al., 2004; Kelessidis
et al., 2005) demonstrated the stabilizing effect of Greek
lignite in terms of rheological and filtration control of
bentonite slurries. Their measurements also showed that
the three parameter HerschelBulkley model describes
well the rheology of these bentonitelignite water
suspensions.
Various rheological models have been proposed to
describe the rheological behavior of bentonite mixtures,
particularly for drilling applications. The two parameter
Bingham plastic model (Bingham, 1922) or the power
law model (Govier and Aziz, 1972; Bourgoyne et al.,
1991) are used most often because of their simplicity
and the fair agreement of predictions with the
rheograms. The power law model, although useful as a
first correction to Newtonian behavior, it may lead to
substantial errors if the fluid exhibits yield stress. Other
two parameter models like the Casson model (Casson,
1959; Hanks, 1989) or the PrandlEyring model
(Govier and Aziz, 1972) have not found wide
acceptance. Three constant parameter models have
been proposed by Herschel and Bulkley (1926), by
Graves and Collins (1978), by Gucuyener (1983) and by
Robertson and Stiff (1976). More complex four
parameters models (Shulman, 1968; Mnatsakanov et
al., 1991) or even five parameter models (Maglione et
al., 1996) have also been proposed. Detailed description
of the various rheological models proposed and
derivation of the appropriate flow equations have been
given by Bird et al. (1982) and by Maglione and
Romagnoli (1999).
The more complex rheological models are deemed
more accurate in predicting the behavior of drilling fluids
than the two parameter models that are widely accepted
at present. However, there is not wide acceptance and
wide application of the more complex models because of
the difficulty in finding analytical solutions for the
differential equations of motion and because of the
complexity of the calculations for the derivation of the
appropriate hydraulic parameters such as Reynolds
number, flow velocity profiles, circular and annular
pressure drops and criteria for transition from laminar to
turbulent flow. Simulation of rotational viscometer data
of non-Newtonian fluids appears to be better when a
larger number of rheological parameters is used but in
this case, the hydraulic parameters can be obtained only
by numerical methods for most of the more complex
rheological models. As of today, a compromise between

the accuracy in the calculations and the simplicity of the


use is required and the best way to achieve this is with the
use of the HerschelBulkley rheological model. The
three parameter HerschelBulkley model has not been
used widely until very recently, although it was not only
proposed almost at the same time as the Bingham plastic
model but it also describes most drilling fluid rheological
data much better (Fordham et al., 1991; Hemphil et al.,
1993; Maglione and Ferrario, 1996; Kelessidis et al.,
2005). The reason for the nonfrequent use is that
derivation of the model's three parameters is complex
(Nguyen and Boger, 1987; Hemphil et al., 1993).
Furthermore, analytical solutions for laminar flow in
pipe and annuli are not possible, requiring either
graphical or trial-and-error solutions (Hanks, 1979;
Govier and Aziz, 1972; Fordham et al., 1991). The
advent of personal computers and their online use in the
field, however, made trial-and-error solutions trivial
tasks, hence, more and more investigators opt to use
HerschelBulkley rheological models in fluid mechanics computations of drilling fluids (Maglione et
al., 1999a; Maglione et al., 2000; Becker et al., 2003). A
search in the Society of Petroleum Engineers electronic
library of scientific articles, covering the period of 1975
2003, resulted in 319 articles having as keywords power
law, 131 articles with keywords Bingham, 51 articles
with keywords HerschelBulkley, and 16 articles with
keywords Casson.
Viscometric data reduction procedures applicable to
various rheological models have been proposed by
many investigators, addressing also some of the inherent
problems associated with data reduction (Krieger, 1968;
Darby, 1985; Borgia and Spera, 1990; Yeow et al.,
2000). The standard procedure for the estimation of the
three rheological parameters for HerschelBulkley
liquids, with rheological equation,
s sy Kn

where , y are the shear stress and the yield stress


respectively, K, n are the fluid consistency and fluid
behavior indices respectively and is the shear rate, is
through non-linear regression of the viscometric data
from concentric cylinder geometry. This is normally
done using a numerical package, minimizing the sum of
error squares and judging the goodness of fit through the
value of the correlation coefficient Rc2 from the
linearized form of Eq. (1), as in Eq. (2),
lnssy lnK nln

However, non-linear fit to various data in this


laboratory with a numerical package sometimes has

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

given the best fit (highest correlation coefficient Rc2) with


negative values for the yield stress, which is meaningless. In these situations, the condition y > 0 must be
imposed to get meaningful results affecting thus the
optimum determination of all three parameters, while
there is also a possibility of non-unique solutions.
Moreover, there is concern over the use of the correlation
coefficient as an indicator of the goodness of fit because
it should be used only for linear functions (Helland,
1988; Ohen and Blick, 1990). In addition, consideration
should be given to differences of rheological parameters
and their effect on subsequent computations, when
choosing data with similar correlation coefficients (Rc2),
like for example data with Rc2 = 0.95 versus data with
Rc2 = 0.96, or between Rc2 = 0.98 and Rc2 = 0.99.
Many times, especially when curve fitting non-linear
functions, the sum of square errors (Q2) is utilized for
the goodness of fit, which is defined as,
X
Q2
yi y i 2
3
i

where yi, i are the measured and the predicted


quantities. Another method to determine the goodness
of fit of a particular rheological model to rheometric data
is the method based on the best index value (BIV). It is
defined as the ratio between the sum of the squares of
the deviation of the predicted value from the average
value y and the sum of the squares of the deviation of
the measured value from the average value (Maglione
and Romagnoli, 1999).
X
y i y 2
i
BIV X
4
yi
y 2
i

The closer the value of BIV to one, the better is the


capacity of the analytical equation (or the rheological
model) to approximate viscometer data. Values of BIV
larger than one indicate tendency for over-prediction
with the model while values lower than one indicate
tendency for under-prediction.
The final choice of the particular parameters can
always be questioned when the choice could be, for
example, between data fits with Rc2 = 0.995 or Rc2 = 0.998,
between Q2 = 0.9 or Q2 = 1.1, and between BIV = 0.99 or
BIV = 1.01, as it will be demonstrated later. Most of the
time, two or even more rheological parameter sets can
approximate well and in a similar way rotational
viscometer data, but in turn they could have large effects
on the end results. For example, parameters like Reynolds
number, velocity profile, pressure drop for the circular and

205

the annular sections in a drilling circuit can all be largely


influenced by this choice and can have desirable or
undesirable carry-over effects. For instance, a variation
in the velocity profile and the Reynolds number can
affect the cuttings carrying capacity of the drilling fluid.
So, the same drilling fluid can be estimated to perform
better if a rheological parameter set is preferred to
another, even though both may have very similar values
of Rc2, Q2 or BIV.
As most flows during drilling oil and gas wells are
laminar, for which analytical solutions exist for the flow
of HerschelBulkley fluids in pipes and annuli, although
not explicit, effort should be made to determine the
impact of the use of different sets of rheological parameters, derived with different methodologies, on the
main parameters of interest, pressure drop and velocity
profiles, thus providing a more robust indicator for the
proper choice of the appropriate rheological parameters.
Prior work (Maglione and Romagnoli, 1998; Maglione,
1999) has suggested that the variation of the flow
behavior index of the HerschelBulkley model could be
the most important factor because it can influence all
hydraulic parameters of the drilling hydraulic circuit,
from flow regimes, to velocity profiles, to pressure drop
and to rates of penetration. It was shown that the lower
the flow behavior index is, the lower the pressure drop is,
in the circular and the annular sections. As the maximum
pressure available is normally constant, the gained
pressure drop (due to a lower n) can be used to increase
the flow rate thus providing more hydraulic power at the
drilling bit or allowing the drilling of longer sections.
The authors concluded, however, that more studies were
required to better define and resolve this problem.
The scope of the present work is to propose a different
and optimal methodology to determine the three
HerschelBulkley rheological parameters of drilling
fluids from concentric cylinder viscometric data, avoiding the observed pitfalls of current non-linear regression
techniques, which may give meaningless negative yield
stress values. Furthermore, the paper aims to demonstrate that the particular choice of the rheological
parameters, among similarly performing data fit curves,
can greatly affect the determination of pressure drop and
velocity profiles of drilling fluids flowing in drilling
hydraulic circuits (pipe and annuli) and to present field
cases of these implications. An evaluation is also
attempted about the conditions for preferring a particular
rheological parameter set, whether it is the best fit of a
particular rheological model to rheometric data or also
the impact of the specific choice of rheological parameter
values on pressure drop and velocity profile estimation
for flow of drilling fluids during drilling.

206

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

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 (Mihalakis 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 AlZahrani (1997).
The experimental data of Mihalakis et al. (2004) and
Kelessidis et al. (2005) were taken with waterbentonite
suspensions at 6.42% w/v, either hydrated for 24 h at
room temperature or aged statically in an aging cell for
16 h at 177 C. Various types of lignites from different
places in Greece were added at 0.5% w/v and 3% w/v in
the suspension in a similar fashion (hydrated or aged
thermally). The bentonite used was a commercial
product used in oil-well drilling (Zenith) provided by
S&B Industrial Minerals S.A. All samples were prepared
following American Petroleum Institute procedures
(2000). The samples were agitated vigorously for
5 min before testing and measurements were made
with a continuously varying rotational speed rotating
viscometer (Grace, M3500) at two sample temperatures,
25 C and 65 C and 12 speeds: 3, 6, 10, 20, 30, 60, 100,
200, 300, 400, 500 and 600 rpm. The data is reproduced
in Appendix B (Table B1).
The data of Merlo et al. (1995), reproduced in Table
B2, were derived with drilling fluid from field
operations during drilling circulation tests at various
sections of the well, taking fluid samples from the outlet
of the drilling circuit. The rheometric data were derived
with a HuxleyBertran high pressure high temperature
rotational viscometer for samples S13, S14, S15, S16
and S17, while data for sample S18 were derived with a
Fann VG 35 six speed rotational viscometer (Merlo et
al., 1995). Al-Zahrani (1997) reports that the rheological
data of Blick (1992), samples S19 to S21 reproduced in
Table B3, were taken with a rotary viscometer for
different suspensions prepared by adding various
quantities of Wyoming bentonite in water.
3. Rheological parameter estimation for
HerschelBulkley drilling fluids
3.1. Current methodology
Some of the rheometric data, given in Table B1, are
presented in Figs. 1 and 2. The three rheological
HerschelBulkley parameters were derived according to

Fig. 1. Rheograms of 6.42% w/v bentonite suspensions in water,


hydrated (sample S1) and thermally aged at 177 C for 16 h (sample
S2). The rheological measurements were taken at 25 C.

standard methodology, using non-linear regression with


an appropriate numerical package. These values are
given in Table 1 together with the correlation coefficient, the sum of errors squared and the best index value,
as defined above. From the rheograms presented in Figs.
1 and 2 above, as well as from the values of the
rheological parameters listed in Table 1, it is evident that
the rheological behavior of these suspensions is typical
of yield-pseudoplastic fluids and modeling their rheological behavior as HerschelBulkley fluids is more
than appropriate.
There exist, however, situations where application of
non-linear regression to rheometric data gives as optimal
set of the three rheological parameters, a set with
negative values of the yield stress, which of course is
meaningless. Reported data showing this behavior is
presented in Table 2 for data of Kelessidis et al. (2005),
of Merlo et al. (1995) and of Blick (1992). There is the
possibility of imposing the condition y > 0 when
deriving the rheological parameters with non-linear
regression which, however, leads to non-optimal solutions, as will be demonstrated later.
It is evident that the use of a numerical package to
apply non-linear regression to rheometric data in order
to derive HerschelBulkley parameters is not always
the optimal procedure since the method can lead to
inappropriate negative values for the yield stress. For
yield-pseudoplastic fluids, as most bentonite suspensions are, a yield value exists either as an engineering
reality or as an inherent fluid property, although there
is considerable dispute about this issue (Barnes and
Walters, 1985; Chen, 1986; Hartnett and Hu, 1989;
Evans, 1992; Schurz, 1992; De Kee and Chan Man
Fong, 1993; Barnes, 1999). The yield value of the
fluid should be derived first using an appropriate
technique, followed by the estimation of the other two
parameters with non-linear regression applied to the

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

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 y is assumed, either taken from the original
rheogram or from a semi-log plot, with y the yintercept, as shown in Fig. 3. In this semi-log plot of
[ log()], 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 ( 0) it should cross
the -axis at the value of the yield stress, a procedure
followed by some investigators (Benna et al., 1999;
Barnes, 1999). This method, however, is not deemed
very accurate because measurements at very low shear
rates may suffer from wall slip effects (Guillot, 1990;
Barnes, 1995).
Other investigators have proposed to measure
separately y by other means, for example utilizing the
vane method (Nguyen and Boger, 1987; Alderman et
al., 1991; Barnes and Nguyen, 2001) which, although it
leads to better determination for the yield stress, it is
time consuming and not applicable for drilling
operations.

Fig. 2. Rheograms of bentonitelignitewater suspensions (samples


S7 and S12) both thermally aged at 177 C for 16 h. The rheological
measurements were taken at 25 C.

linearized form of HerschelBulkley rheological


equation (Eq. (2)).
This could be accomplished through a trial-and-error
procedure, where various values of y are assumed and
the original ( ) rheological data is transformed to [ln
( y) ln ] data in order to establish the best linear
curve, determined from the correlation coefficient and
the sum of error squares, giving as y-intercept, (ln K),
while the slope of the curve will be the flow behavior
index, (n) (Hemphil et al., 1993; Turian et al., 1997).
With this approach, however, there is no guarantee that
the optimum values have been derived since the
determination of (K,n) depends on the correct estimation
of y. Turian et al. (1997) found the procedure very
tedious and the results very sensitive to the assumed
value of the yield stress. They further stated that a major
weakness of the HerschelBulkley model has been the
inability to unambiguously establish the model parameters since different sets of these values can provide
equivalent fits of the experimental data, exactly the kind
of behavior described in this work. Furthermore, a set of
HerschelBulkley rheological parameters derived with
a correlation coefficient of 0.998 could be different than

3.2. New methodology


A new methodology is proposed, amenable to
computer implementation, which determines the three
rheological HerschelBulkley parameters avoiding
potential errors of deriving negative yield stress values.
The methodology is based on an initial optimal
determination of y, using a near optimum form of the
Golden Section search, sometimes called the Fibonacci
search, followed by linear regression of the linearized
form of the HerschelBulkley rheological equation. The
technique has been proposed by Ohen and Blick (1990)
for Robertson and Stiff fluids (Robertson and Stiff, 1976)

Table 1
HerschelBulkley rheological parameters of bentonitewater and bentonitelignitewater suspensions (data of Kelessidis et al., 2005)
Sample

y (Pa)

K (Pa sn)

R2c

Q2 (Pa2)

BIV

S1
S2
S3
S4
S5
S6
S7
S8
S9
S10

8.4748
11.3025
0.0788
0.0000
0.6751
6.3938
2.4095
1.1843
3.4701
0.3793

3.4010
5.9115
2.3861
0.2050
0.0732
0.4498
0.1251
0.1265
0.0313
0.0567

0.2556
0.2645
0.3407
0.4930
0.7001
0.5001
0.7012
0.6436
0.8045
0.6196

0.9876
0.9885
0.9340
0.9450
0.9937
0.9951
0.9975
0.9965
0.9666
0.9983

3.8520
12.6820
44.8843
4.1802
0.7718
1.2885
0.9188
0.5671
3.3209
0.0395

0.9980
0.9883
0.9343
0.6399
0.9927
0.9945
0.9980
0.9961
0.9687
0.9982

208

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

Table 2
HerschelBulkley rheological parameters of bentonitewater and bentonitelignitewater suspensions with negative y, derived from nonlinear
regression
Sample

y (Pa)

K (Pa sn)

R2c

Q2 (Pa2)

BIV

Source

S11
S12
S14
S15
S17
S19
S20
S21

0.2685
0.2880
0.0932
1.1650
0.6213
0.0906
3.7262
6.0980

0.2486
0.2210
1.9580
2.3990
1.1400
1.1371
5.4440
9.2885

0.5312
0.5841
0.3488
0.3158
0.3704
0.4393
0.3222
0.3208

0.9957
0.9963
0.9990
0.9988
0.9994
0.9964
0.9982
0.9988

0.5402
0.7890
0.2595
0.2850
0.0739
1.3320
2.4130
4.7160

0.9848
0.9960
0.9991
0.9981
0.9995
0.9958
0.9985
0.9996

Kelessidis et al. (2005)


Kelessidis et al. (2005)
Merlo et al. (1995)
Merlo et al. (1995)
Merlo et al. (1995)
Blick (1992)
Blick (1992)
Blick (1992)

which follow rheological Eq. (5) below, with A, B, C the


RobertsonStiff rheological parameters, and it has been
extended to HerschelBulkley fluids in this work.
s A CB

The aim is to find the value of the yield stress, y,


which minimizes the error of the difference between the
predicted and the measured shear stress values, in other
words, to bracket a minimum for a given function f (x)
in the interval (a, c). A minimum is known to be
bracketed only when there is a triplet of points a < b < c
(or c < b < a) such that f (b) is less than both f (a) and f
(c) (Press et al., 1992). The function is then evaluated at
an intermediate point, x, which is chosen, either between
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),
otherwise, if f (b) > f (x), the triplet becomes (b, x, c).
The process continues until the bracketing interval is
tolerably small. The strategy for choosing point b (or x),
given (a, c) leads to the Golden Section search, which
states that the optimal bracketing interval a, b, c has its

Fig. 3. Graphical determination of yield stress, y, taken as the yintercept from a semi-log plot of the original rheogram.

middle point b a fractional distance from one end, for


example a, of a value equal to the golden ratio of
0.61803 and a fractional distance from the other end of
0.38197 (Fig. 4). At each stage of the search of the
bracketing interval of the minimum, the next point to be
tried is the point x which is a fractional distance 0.61803
from one end and a distance of 0.38197 from the other
end (Press et al., 1992).
To estimate the three rheological parameters y, K, n,
a combination of iteration, to determine the yield stress
using the Golden Section search method, and a least
squares fitting of the linearized HerschelBulkley equation, is used (Ohen and Blick, 1990). This is accomplished
by first picking an initial interval of search (L, U), with L
the lower limit and U the upper limit, defined as,
U sy0 tol

L sy0 tol

where tol is the half width of the search interval and y0


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 y0. Although the value of y0 can be any
non-zero value and the procedure will converge, y0 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 , three
data sets are defined, one at the lowest shear stress, min
min, one at the highest shear stress, max max, while
the third set is based on the geometric mean shear stress

Fig. 4. Schematic explaining the Golden Section ratio.

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

from the first two points, that is, ()2 = (min *max), with

derived by interpolation. A system of three equations


with three unknowns is thus derived as,
0
smin sy0 K0 nmin

0
smax sy0 K0 nmax

s sy0 K0 n0

10

This system can be solved to get the initial value of


y0. Combination of Eqs. (8)(10) gives,
0
0
nmin
smax smin nmax
n0
n
smax s
max 0

n0 nmin

0
0
s
2n0 nmin
nmax

n0
0
0
n
max 2 0
2n0 nmin
nmax

12

The triplet of points (L, y0, U) is thus established


following the above procedure. The new points, in the
search of the minimum, are evaluated using the golden
ratio of 0.61803 as,
sy1 L 0:61803U L

13

for the point between (L, y0), while for the point
between (y0, U) the new value becomes,
sy2 U 0:61803U L

quick convergence. The search interval varies according


to the conditions,
R2c1 < R2c2 YU sy1

15

R2c1 > R2c2 YL sy2

16

R2c1 R2c2 YU sy1 ; L sy2

17

The whole procedure can be performed with a


numerical package.

11

with 0 < n0 < 1. Eq. (11) can be solved easily numerically


to get n0 and it is then used to derive the initial value for
yield stress, y0,
sy0

209

14

The functional relationship for optimization is


provided by the correlation coefficient Rc2 of Eq. (2)
for the chosen value of y. Schematically, the relationship Rc2 y is shown in Fig. 5 for successive
approximations of y. It is a unimodal function implying

Fig. 5. Variation of correlation coefficient with assumed values of yield


stress (the case represents real data).

3.3. Computational results and comparison with


experimental data
This methodology has been applied to several data
sets, given in Tables B1B3 in Appendix B, and has
proved to work extremely well. In Table 3, the
rheological parameters that have been calculated using
both procedures, non-linear regression with a numerical
package and with the Golden Section technique are
presented for the cases where non-linear regression
provides meaningful results, in other words, positive
yield stress values. Rheograms of some of the samples
are shown in Figs. 6 and 7 together with the curves
derived from non-linear regression (NL) and for Golden
Section search (GS).
The results show minor differences among the two
methods both in terms of all three rheological
parameters and the two of the three indices of correlation
(Rc2, Q2) while the third index (BIV) does not really
reflect the similarities between the estimated parameters.
This close agreement demonstrates the ability of the new
proposed scheme to properly determine the rheological
parameters with similar success as the application of
nonlinear regression using numerical packages, for the
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 parameters obtained by the two methodologies raises the
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) 203224

Table 3
Comparison of rheological parameters derived from non-linear regression (NL) and by Golden Section (GS)
Sample and method
S7
NL
GS
S8
NL
GS
S9
NL
GS
S10
NL
GS
S13
NL
GS
S16
NL
GS
S18
NL
GS

y (Pa)

K (Pa sn)

R2c

Q2

BIV

Source

2.4095
2.4141

0.1251
0.1369

0.7012
0.6842

0.9975
0.9910

0.9188
1.3124

0.9980
0.9401

Kelessidis et al. (2005)

1.1843
1.3012

0.1265
0.1058

0.6436
0.6680

0.9965
0.9967

0.5671
0.6013

0.9961
0.9855

Kelessidis et al. (2005)

3.4701
2.8973

0.0313
0.1566

0.8045
0.5661

0.9666
0.9800

3.3209
4.8284

0.9687
0.8285

Kelessidis et al. (2005)

0.3793
0.2847

0.0567
0.0839

0.6196
0.5625

0.9983
0.9960

0.0395
0.0670

0.9982
0.9615

Kelessidis et al. (2005)

1.7020
0.0000

1.2063
1.9940

0.4352
0.3704

0.9971
0.9952

1.0847
1.8105

0.9959
1.0246

Merlo et al. (1995)

0.1747
0.0379

0.9448
1.0200

0.4097
0.3993

0.9990
0.9995

0.1563
0.1636

0.9989
0.9956

Merlo et al. (1995)

2.6750
1.6813

0.2492
0.6496

0.6607
0.5173

0.9982
0.9950

0.7375
5.1244

0.9977
0.8035

Merlo et al. (1995)

Positive yield stress values.

choice is among values giving very similar data fit.


Furthermore it raises questions about the implications
and the real impact on the operational parameters of
interest, pressure drop and velocity profiles.
The new methodology has been applied also to
rheometric data which give negative yield stress values
when applying non-linear regression, with the results
shown in Table 4. The listed rheological parameters are
derived using three different techniques: application of
non-linear regression (NL, numerical-no penalty),
application of non-linear regression with imposition of
y > 0 (NLP, numerical with penalty) and application of
the Golden Section search (GS). Imposing y > 0 may

sometimes result in non-optimum solutions giving


lower correlation coefficients, much higher sum of
squared errors and values of best index value significantly different than one, compared to cases of nonlinear regression but without this imposition and to
cases of Golden Section search methodology (samples
S11, S12, S15). For some of the cases though, NLP
method may result in as good predictions as the Golden
Section search (samples S14, S17, S19, S20 and S21).
The new proposed scheme provides good solutions, as
evidenced by the sum of squared errors, the correlation

Fig. 6. Comparison of rheograms of original data with results from


rheological models for bentonitelignitewater suspensions.

Fig. 7. Comparison of rheograms of original data with results from


rheological models for drilling mud.

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

211

Table 4
Comparison of rheological parameters derived from numerical package and by proposed scheme, non-optimal solutions
Sample and method
S11
NL
NLP
GS
S12
NL
NLP
GS
S14
NL
NLP
GS
S15
NL
NLP
GS
S17
NL
NLP
GS
S19
NL
NLP
GS
S20
NL
NLP
GS
S21
NL
NLP
GS

y (Pa)

K (Pa sn)

R2c

0.2685
0.0000
0.0718

0.2486
0.2326
0.1462

0.5312
0.5166
0.6068

0.9957
0.9590
0.9942

0.2880
0.0001
0.3976

0.2210
1.0493
0.0940

0.5841
0.3218
0.7036

0.09323
0.0000
0.0000

1.9580
1.8900
1.9050

1.1650
0.0076
0.0000

Q2

BIV

Source

0.5402
5.1540
0.7865

0.9948
0.9149
1.0263

Kelessidis et al. (2005)

0.9963
0.9160
0.9942

0.7890
34.15
1.5646

0.9960
0.4435
0.9881

Kelessidis et al. (2005)

0.3488
0.3530
0.3523

0.9990
0.9990
0.9990

0.2595
0.2627
0.2608

0.9991
0.9938
0.9987

Merlo et al. (1995)

2.3990
1.7260
1.7330

0.3158
0.3813
0.3561

0.9988
0.9959
0.9981

0.2850
0.9695
0.4583

0.9981
1.5152
0.9832

Merlo et al. (1995)

0.6213
0.0000
0.3767

1.1400
0.8491
0.4160

0.3704
0.4079
0.4083

0.9994
0.9987
0.9987

0.0739
0.1619
0.1619

0.9995
0.9196
0.7949

Merlo et al. (1995)

0.0906
0.0000
1.4701

1.1371
1.1170
0.6234

0.4393
0.4414
0.5203

0.9964
0.9964
0.9927

1.3320
1.3380
2.2244

0.9958
0.9930
1.0398

Blick (1992)

3.7262
0.0000
0.0000

5.4440
3.6872
3.5776

0.3222
0.3683
0.3739

0.9982
0.9974
0.9975

2.4130
3.3610
3.7578

0.9985
0.9890
1.0204

Blick (1992)

6.0980
0.0003
0.0000

9.2885
6.3864
6.1803

0.3208
0.3650
0.3712

0.9988
0.9980
0.9987

4.7160
7.2000
8.4888

0.9996
0.9900
1.0262

Blick (1992)

(NL) is the application of non-linear regression, (NLP) is the application of non-linear regression with the imposition of y > 0 and (GS) is the
application of the new technique.

coefficients and the best index values. The suitability of


the proposed scheme is further demonstrated in Figs. 8
12, where the rheograms for some of the samples of

Table 4 are given for the various approaches. Similar


graphs produced for the rest of the samples have further
demonstrated the suitability of the proposed technique.

Fig. 8. Comparison of rheograms of original data with results from


rheological models for bentonitelignitewater suspensions (sample
S11).

Fig. 9. Comparison of rheograms of original data with results from


rheological models for bentonitelignitewater suspensions (sample
S12).

212

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

Fig. 10. Comparison of rheograms of original data with results from


rheological models for drilling mud (sample S15).

From the data reported in Table 4 it is evident that


the values of the flow behavior index, determined
with the three different methods, do not differ
significantly except for sample S12. The main
difference is the yield point which can affect both
cuttings carrying capacity of drilling fluids and
pressure drop. The results of Tables 3 and 4 show
that when the fluid shows a positive yield point,
when determined by non-linear regression, the GS
method determines quite different rheological parameters with the larger difference observed on the
value of the flow behavior index. When the fluid
shows negative value of the yield point, when derived
by non-linear regression, the only difference derived
when using the GS method is on the yield point
itself.
3.4. Discussion
The results presented above have indicated that
optimal determination of the rheological parameters of
drilling fluids described by the HerschelBulkley

Fig. 11. Comparison of rheograms of original data with results from


rheological models for drilling mud (sample S19).

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


rheological models for drilling mud (sample S21).

rheological model is not an easy task. There will be


cases where application of non-linear regression with
available numerical packages can give estimates of the
three rheological parameters with high degree of
confidence. Caution, however, should be exercised
when these techniques give meaningless negative
yield stress values. Imposition of the condition that
yield stress is positive may lead to significant differences of the estimated rheological parameters and the
choice of an inappropriate rheological model, other than
the HerschelBulkley model which might have been
derived if determination of the rheological parameters is
performed following the Golden Section search methodology proposed here.
It has been demonstrated that the proposed
technique, which relies on the proper choice of the
yield stress using the Golden Section search methodology and on the minimization of the sum of squared
errors on the linearized form of the HerschelBulkley
rheological equation, can lead to meaningful and
appropriate values of all three rheological parameters
with high degree of confidence. The technique
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 correctness. 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 HerschelBulkley
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
known nor it has been investigated thoroughly in the
past.

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

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, nonlinear 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
internal diameter of the outer pipe of 0.4445 m and an
external diameter of the inner pipe of 0.127 m, an
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
outer pipe of 0.216 m and an external diameter of the
inner pipe of 0.089 m. Results have been derived for a
range of flow rates encountered in oil-well drilling
situations, but keeping the flow laminar, which is the
normal condition encountered in drilling applications,
particularly for flow in annulus. The appropriate flow
equations for HerschelBulkley fluids are summarized
in Appendix A, both for pipe flow and for concentric
annulus, utilizing slot flow solutions for the latter, a
common approach in drilling hydraulics, particularly for
annulus diameter ratios greater than 0.3 (Bourgoyne et
al., 1991; Maglione and Romagnoli, 1999).
Computations of pressure drop-flow rate data sets
have been performed for all samples with rheological
data presented in Tables B1B3. Results, however, are
presented here for samples S12, S17 and S19, for which
negative yield values have been estimated as well as for
some of the samples giving positive yield stress values
(samples S7, S9, S10 and S18). The rheological
parameters were determined with the new methodology
applying Golden Section search (GS) and with nonlinear regression with penalty for y > 0 (NLP) and are
listed in Tables 3 and 4.
In Fig. 13, data is presented for the pressure drop
variation with the flow rate, for flow in the 0.311 m
by 0.127 m concentric annulus. For sample S12, there
is a 2 to 3 times variation between the computed
pressure drop using rheological parameters derived
with the GS method when compared with the values
computed when the rheological parameters were
obtained according to method NLP. For this sample,

Fig. 13. Pressure drop-flow rate graph for the three fluids, with
rheological parameters determined by Golden Section (GS) and by
non-linear regression with y > 0 (NLP) in a 0.311 m by 0.127 m
concentric annulus. Laminar flow computations.

the correlation coefficient for NLP is a low 0.9160


(Table 4) while for method GS is 0.9942, hence the
variation observed in Fig. 13 may be partly explained
by the bad fit of the rheogram when using method
NLP. For sample S17, the variation in the predicted
pressure drop values with the rheological parameters
derived by methods GS and NLP is between 1.5 to 2.5
times. For this case, however, the correlation coefficient for the estimation of the rheological parameters
is a high 0.9987 for both methods (Table 4). For
sample S19, for which the rheological parameters
were determined with GS have a correlation coefficient of 0.9927 and with NLP have a correlation
coefficient of 0.9964 (Table 4), the difference in the
pressure drop predictions is smaller than the previous
cases but it is still significant ranging from 10% at
high flow rates to 50% at low flow rates.
The computed velocity profiles for the 0.311 m by
0.127 m concentric annulus at various flow rates for
sample S19, the sample that showed the smallest
variation in pressure drop values, are shown in Fig.
14. Small variations on velocity profiles are observed
with larger variations in plug velocities, particularly at
high flow rates. Larger variations are expected and have
been determined for the other fluid samples for which
larger pressure drop variations have been observed
when using rheological parameters derived with the two
techniques.
Fig. 15 shows the results of the pressure drop
computations for the smaller annulus and similar trends
are observed as in Fig. 13. For sample S12, the variation
between the pressure drop values computed with

214

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

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

Fig. 16. Velocity profiles for fluid S12 with rheological parameters
determined by Golden Section (GS) and non-linear regression with
penalty (NLP), for the 0.216 m by 0.089 m concentric annulus, for five
flow rates.

rheological parameters obtained by methods GS and


NLP, is two to three times. For sample S17, this
variation is between 1.5 to 2 times, while for sample S19
there are no significant differences for flow rates higher
than 600 l/min. The velocity profiles for sample S12, the
sample that showed the largest variation in the estimated
pressure drop values, are shown in Fig. 16, and
differences are observed both in velocity profiles but
also in maximum (plug) velocities for the computed
values using the rheological parameters from the two
cases.

For pipes, the results for pressure drop estimation are


shown in Fig. 17 for the pipe size of 0.1264 m and Fig.
18 for the pipe size of 0.0883 m for samples S12, S17
and S19, for laminar flow. The variation observed is
similar to the presented cases for the concentric annulus,
with significant variation for samples S12 and S17 for
both pipe sizes, while the variation is very small for
sample S19 for both pipe sizes.
Similar computations, for the samples for which the
yield stress was found positive and with no significant
differences among the model parameters listed in Table
3, have revealed that the differences in pressure drop

Fig. 15. Pressure drop-flow rate graph for three fluids, with rheological
parameters determined by Golden Section (GS) and by non-linear
regression with penalty (NLP), for the 0.216 m by 0.089 m concentric
annulus. Laminar flow computations.

Fig. 17. Pressure drop-flow rate graph for three fluids, with rheological
parameters determined by Golden Section (GS) and by non-linear
regression with penalty (NLP), for the 0.1264 m pipe. Laminar flow
computations.

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

Fig. 18. Pressure drop-flow rate graph for three fluids, with rheological
parameters determined by Golden Section (GS) and by non-linear
regression with penalty (NLP), for the 0.0883 m pipe. Laminar flow
computations.

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 selection 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
to derive an index of appropriateness for estimating
rheological parameters.
4.2. Implications on drilling operations

computation are much smaller in this case while the


velocity profiles for the cases of these samples do not
vary much. Some of the results for the pressure drop
estimation are shown in Fig. 19 for samples S7, S9, S10
and S18.
The presented results demonstrate that differences in
computed pressure drop values and velocity profiles
exist when computations are performed using either the
rheological parameters obtained by GS (the most
correct one) or by NLP. The variations can be rather
large for typical drilling situations not only on pressure
drop but also on velocity profiles. This is particularly

The implications on drilling from use of different


rheological parameters but from same rheological data
set have been assessed with field data of Merlo et al.
(1995). An analysis is presented for the circulation
test performed in well A at the depths of 555 m and
2008 m with the flow geometry shown in Fig. 20.
The circulation test at 555 m depth is deemed as the
most reliable in terms of data analysis because the
drill pipe was located inside the casing with a
nominal diameter of 0.508 m (inside diameter of
0.4826 m) with a well-known geometry and because
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
by non-linear regression (NL), for the 0.311 m by 0.127 m concentric
annulus. Laminar flow computations.

Fig. 20. Geometry of well A for the reported drilling circulation test at
3200 l/min. Relevant geometrical data are given by Merlo et al.
(1995).

216

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

the temperature of the well. For purposes of


simulation, the S13 sample was considered, for
which the rheological parameters have been determined 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
29.100 Pa (Merlo et al., 1995). From the above data it
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
fluid respectively. Furthermore, the pressure drop
calculated by using the GS fluid is less than the
pressure drop calculated with the NL fluid by an
amount equal to 16.4% of the GS fluid value,
indicating that the GS fluid performs better than the
NL fluid, in terms of pressure drop. This holds even
though both fluids have the same formulation and
characteristics and the viscometer data are very well
described by both methods ( R c2 = 0.9971 and
Rc2 = 0.9952 for the NL fluid and the GS fluid
respectively). In Fig. 21, the variation of the calculated
frictional pressure loss in the drilling hydraulic circuit
is plotted against the flow rate for the NL fluid and
GS fluid, together with measured data.
When the flow velocity distribution in the annulus
(0.483 m by 0.127 m) is computed for the two fluids at a

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
rate of 3202 lpm.

flow rate of 3202 lpm, the results show similar velocity


profile in the annular section (Fig. 22) meaning that both
fluids can have the same cuttings carrying capacity.
Cuttings transport efficiencies, calculated as per Walker
and Mayes (1975) but extended for HerschelBulkley
fluids by Gallino and Maglione (1996), are estimated at
78.9% and 77.5% for the NL fluid and GS fluid
respectively.
Finally, considering the apparent viscosity of the
drilling fluid at the nozzles of the drilling bit, a,
which affects the rate of penetration among other
parameters, differences in the behavior of the NL and
GS fluids are expected which can be evaluated. The
apparent viscosity at the drilling bit can be
computed, following standard drilling hydraulics
procedures (Bourgoyne et al., 1991; Maglione and
Romagnoli, 1999), by estimating the wall shear
stress, w, and the wall shear rate, w, of the drilling
fluid at the bit nozzles, and computing the apparent
viscosity, a, as,
la

Fig. 21. Pressure drop-flow rate curves for drilling circulation test for
Well A, using NL and GS derived rheological parameters for drilling
fluid sample S13. The measured data at the one flow rate (3202 lpm) is
also shown.

sw
w

18

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
flow rates.
The rate of penetration during drilling (ROP) is
directly affected by fluid density (), fluid velocity in the

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

217

fluid, with the only difference in the approach


followed to determine the best fit rheological parameters from viscometer data. The variation of the
rheological parameters affects mostly frictional pressure drop and apparent viscosity at the bit. Hence, at
the design phase and during oil-well drilling operations, 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
annulus of 0.4826 m by 0.127 m).

drilling bit nozzles (Vn), bit nozzle diameter (dn), while


it is inversely proportional to drilling fluid apparent
viscosity at the drilling bit (a,bit) (Eckel, 1967; Beck et
al., 1995; Gallino et al., 1996). This variation is
described well by Eq. (19),
qVn dn
ROP~
la;bit

!c
19

where c is a constant, which depends on the drilled


formation. From Eq. (19) one can notice that, keeping
all other parameters constant, the rate of penetration is
inversely proportional to the apparent viscosity at the
bit. Considering for instance the case for well A
reported above (Merlo et al., 1995), if a value for the
rate of penetration of 10 m/h is assumed for the NL
fluid, the corresponding rate of penetration for the GS
fluid would be predicted around 12.6 m/h, an increase
of 26%. This particular example reveals that just by
considering and studying better the best fit curves of
viscometer data of drilling fluids, a surprising result
with respect to the rate of penetration can be derived.
This significant increase has been computed using
rheological parameters that were very close to each
other (Table 3, sample S13). The results could be
more dramatic, if larger differences occur among
rheological parameters derived using the two different
techniques.
All the above results could be considered as the
direct effect of the decrease of the flow behavior
index of the drilling fluid when determined using the
GS method and its consequences on hydraulic
parameters. It seems that the GS fluid performs better
than the NL fluid, even though they are the same

5. Conclusions
In this study it was shown that the HerschelBulkley
rheological model properly described all rheological
data of drilling fluids obtained with rotational viscometer data. It has been demonstrated that the three
rheological HerschelBulkley parameters can be
derived with a numerical package using non-linear
regression but the procedure may not always lead to
optimal solutions, because sometimes, meaningless
negative yield stress values are determined. Imposition
of the condition for positive values of the yield stress
gives non-optimal solutions. A methodology has been
proposed to alleviate this problem. It estimates the best
value for the yield stress using the Golden Section
search methodology, and then applies linear regression
to the transformed data, yielding as accurate results as
the numerical schemes in normal cases but also giving
positive values for the yield stress in situations where
numerical schemes determine negative values. The
recommended approach leads to unique solutions and
can be easily implemented.
It was further shown that pressure drop and
velocity profiles for laminar flow in pipes and
concentric annulus can be significantly affected by
proper choice of rheological parameters. The most
appropriate set should be determined utilizing not only
the statistically best fit indices but also the impact on
pressure drop and velocity profiles. Use of the
relevant flow equations for laminar flow in typical
oil-well drilling situations in concentric annulus and
pipe can aid significantly in determining the impact of
the particular choice of rheological parameters on
pressure drop, on velocity profiles and on rate of
penetration, aiding in choosing the most appropriate
rheological parameters. The computed results have
demonstrated that it is very important to make the best
simulation of rheological behavior of drilling fluids
before computing hydraulic parameters. Once the best

218

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

rheological parameters are obtained, utilization of


appropriate calculation tools for estimation of pressure
drop and velocity profiles can aid in avoiding many
problems during drilling operations and can allow
better exploitation of the properties of existing drilling
fluids.
Nomenclature
A, B, C RobertsonStiff rheological parameters
BIV
Best index value
dp/dL Pressure drop per unit length (M/L2T2)
dn
Bit nozzle diameter (L)
h
Gap of annulus, distance between plates in the
slot (L)
K
Fluid consistency index (M/LT2n)
L
Lower limit for the estimation of yield stress
n
Fluid behavior index
p
Pressure (M/LT2)
r
Radius (L)
R
pipe radius (L)
R1
radius of outer pipe of annulus (L)
R2
radius of inner pipe of annulus (L)
Rc2
Correlation coefficient
ROP
Rate of penetration (L/T)
q
Flow rate (L3/T)
Q2
Sum of error squares
tol
Half width of search interval
u
Liquid velocity (L/T)
U
Upper limit for the estimation of yield stress
Vn
Liquid velocity at the drill bit nozzle (L/T)
ya
Distance of top of inner layer from bottom
plate (L)
yb
Distance of bottom of top layer from bottom
plate (L)
yi
Measured parameter
i
Predicted parameter
y
Average value of the parameter
w
Width of the slot (L)
Greek letters

Shear rate (T 1)

Geometric mean shear rate (T 1)

pressure drop per unit length (M/L2T2)

Liquid viscosity (M/LT)

Dimensionless parameter, ratio of yield stress


to wall shear stress

Liquid density (M/L3)

Shear stress (M/LT2)

Geometric mean shear stress (M/LT2)


y
Yield stress (M/LT2)
y0
Initial estimate of yield stress (M/LT2)
o
Integer constant

Subscripts
min
minimum value
max
maximum value
o
initial value
w
wall
a
apparent value
a, bit
at the bit
Appendix A. Flow of HerschelBulkley drilling
fluids in pipes and concentric annuli
A.1. Flow in pipes
The geometry for pipe flow together with the
appropriate parameters is shown in Fig. A1. There is a
central core of the fluid which moves as a rigid plug if
the shear stress levels are smaller than the yield stress of
the fluid. Letting = dp/dL, with dp/dL the pressure
drop per unit length, and for values of the shear stress
greater than the yield stress of the fluid y, y,
balance of forces gives,
dsr
dp
r
rD
dr
dL

A1

Eq. (A1) upon integration gives,


r
C1
s D
2
r

A2

The shear stress is finite at r = 0, hence, C1 = 0. The


radius at which there is an unsheared portion of the fluid,
rp, is given by
sy

rp
D
2

A3

and the wall shear stress, w, is given by,


sw

R
D
2

A4

Fig. A1. Geometry and parameters for laminar flow in pipes.

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

219

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


HerschelBulkley fluid is given by,


du n
s sy K
dr

A5

where K, n are the fluid consistency and flow behavior


indices respectively. Hence,


du
sy K
dr

n

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

r
D
2

A6
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
wall, u = 0 at r = R, is finally given by,
1
1
u
D=2K m 1
sy
VrVR
rp
D=2

(

sy
D
R
2K
K

m1 
 )
sy m1
D
r

;
2K K

A7a

For flow in a concentric annulus, the annulus can be


approximated by a slot of gap h = R2 R1, as represented
in Fig. A2. The solution that will be derived follows the
procedure for flow of Bingham plastic fluids in a slot,
given by Bourgoyne et al. (1991).
There is a central core of the fluid which moves as a
rigid plug if the shear stress levels are smaller than the
yield stress of the fluid. Integration of the force balance
performed on a fluid element gives
s yD s0

and


sy m1
D=2Km
u up
R
; 0VrVrp
m1
D=2

A7b

with m = 1/n.
The flow rate q can be derived as,
Z
q 2k

with 0 an integration constant to be determined. Let ya,


yb the distances of the lower sheared and upper sheared
surfaces from the bottom plate respectively. At the inner
layer of the plug, ya, the shear stress a must equal
( y). It follows then from (A9),
sa sy s0 ya D

urdr

A9

A10

kn DR=2sy 1=n1
K 1=n
D=23
"
#
s2y
DR=2sy 2 2sy DR=2sy


1 2n
1 3n
1n

which gives,
ya

A8
Eq. (A8) relates pressure drop ( = dp/dL) with
flow rate (q) for flow of HerschelBulkley drilling
fluids in a pipe of radius R for laminar flow. If the
pressure drop is known, the flow rate can be directly
computed. On the other hand, if the flow rate is
known, Eq. (A8) can be solved by trial-and-error for
the estimation of pressure drop. The velocity profiles
can be computed readily from Eqs. (A7a) and (A7b)
for both cases.

sy s0
D

A11

Similarly, for the outer plug region one obtains for


the shear stress at yb,
sb sy s0 yb D

A12

In the fluid layer enclosed by the inner layer of the


plug and the bottom plate, the shear stress is,
s sy K

 n
du
s0 yD
dy

A13

220

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

The solution to differential Eq. (A13), utilizing


boundary condition u = 0 at y = 0 is finally given by,

The plug velocity, given by Eqs. (A16) and (A20) is


the same, hence,

( 

 )

K
sy s0 m1
sy s0 D m1

;
y
u

m 1D
K
K
K

hyb m1
ym1
a
A14

0VyVya

with m = 1/n. In terms of the parameter ya this equation


becomes
o
D=K n
ya m1 ya ym1 ;
u
m 1

A21

Taking the (m + 1)th root and keeping only the


positive value, it follows that,
ya hyb Z ya yb h

A22

Substituting the values of ya, yb, given above, the


unknown 0 can now be determined as,

0VyVya
A15

h
h dp
s0 D
2
2 dL

A23

Furthermore, the following holds,


The plug velocity, up, is the velocity at y = ya, given
sw s0

by,
up

 m
ym1
D
a
;
m 1 K

ya VyVyb

A16

For the fluid region enclosing the plug and the upper
plate, the shear stress is


du n
s sy K
s0 yD
dy

A17

and following similar approach as above, utilizing the


boundary condition u = 0 at y = h, the velocity is given
by,

1
u
s0 sy hDm1
DK m m 1

m1
s0 sy yD
; yb VyVh


 m

D
1
hyb m1 yyb m1 ; yb VyVh
K m 1
A19

The velocity of the plug is given when y = yb, hence,

up

 m
D hyb m1
; ya VyVyb
K
m 1

sy
dp
>
dL h=2

A25

When there is flow, then,


h
h dp
sw
s0 D
2
2 dL

A26

h sy
ya
2 D

A27

h sy

2 D

A28

The velocity profile is given by,


u

In terms of yb, Eq. (A18) becomes,

A24

The summary of the equations governing flow of


HerschelBulkley drilling fluids in a slot is given below.
There is fluid flow only if,

yb
A18

h dp
2 dL

o
D=Km n
ya m1 ya ym1
0VyVya
m 1
A29a

 m
ym1
D
a
u
m 1 K

ya VyVyb

A29b

 m


D
1
m1
m1
u
hyb yyb
; yb VyVh
K m 1
A29c

A20
with m = 1/n.

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

The flow rate per unit width of the slot, w, is given

221

If now is defined as,

by,
q

Z
udy

ya

Z
udy up

yb

Z
dy

ya

udy
and noting that for partial or full flow it must be true that
< 1, the flow rate is,

h Z h
Z h  

q
du
du

uy
y dy 0
y
dy
w
dy
dy
Z h 0 
Z0 ya 0 
du
du
y
y
dy
dy I1 I2

dy
dy
yb
0

 m
D 2wh=2m2 1nm1
n m 1
q
K
m 1m 2
A33

With some mathematical manipulation, it can be


shown that,
 m
D
ym2
a
K m 1m 2


dp 1=n 
11=n
sy
dL
1
q
1=n 11=n 2
h=2dp=dL


sy
1
1

A34
h=2dp=dL n

#
 m "
D
hyb m2 yb hyb m1
I2

K
m2
m1

 m
D
1
K m 1m 2

6
q4

 fym2
m 1hyb m2
a
A30

The flow rate can be expressed in terms of parameters


y, by replacing ya, yb and noting that 0 = h/2, so
that,

 m 
  
sy s0 m2
D
w
q

D
K
m 1m 2
 s s m2
y
0
m 1 h
D
s s  s s m1

y
0
y
0
h
m 2
D
D

1
K

The final result for the flow rate in terms of annulus


geometry parameters, noting that, wh = (R22 R12), and
h = R2 R1, is,

So finally,

m 2yb hyb m1 g

An equivalent expression has been derived by


Grinchik and Kim (1974) as referenced by Fordham et
al. (1991), where though, only the final result was given.
The flow rate also can be expressed in terms of dp/dL
and y as,

2wh=221=n

and

A32

yb

which can be written as,

I1

sy 2sy sy

s0 hD sw

A31

kR22 R21 R2 R1 11=n


2

1=n

1 dp
K dL

1=n 12=n 4

1=n 3
7
5


 1


11=n
sy
R2 R1 =2dp=dL
"
#
sy
1n 1
R R dp=dL
2

1=n

1=n 12=n 4

A35

Eq. (A35) provides the relationship between flow


rate and pressure drop for HerschelBulkley drilling
fluids flowing in laminar flow in concentric annuli,
modeled as a slot. The flow rate can be directly
determined given the pressure drop but trial-and-error
solution is required if the pressure drop is to be
determined for a given flow rate.

222

Table B1. Rheological data for bentonitewater suspensions and bentonitelignitewater 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 + L7 0.5%
A + L4 3.0%
A + L8 0.5%
A + L3 3.0%
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%
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
stress

Shear
stress

Shear
stress

Shear
stress

Shear
stress

Shear
stress

Shear
stress

Shear
stress

Shear
stress

Shear
stress

Shear
stress

Shear
stress

(1/s)

(Pa)

(Pa)

(Pa)

(Pa)

(Pa)

(Pa)

(Pa)

(Pa)

(Pa)

(Pa)

(Pa)

(Pa)

1021.38
851.15
680.92
510.69
340.46
170.23
136.18
102.14
51.07
34.05
17.02
10.21
5.11

28.50
27.75
26.83
25.00
22.58
21.00
21.00
20.08
16.67
17.50
16.00
14.67
13.20

48.42
46.83
44.58
41.58
37.83
35.67
33.25
30.83
27.42
27.83
21.90
23.75
19.90

21.92
22.75
26.08
22.42
18.58
13.58
12.00
10.50
6.92
7.42
6.67
6.25
5.10

6.92
6.17
5.58
4.75
3.67
2.17
1.83
1.33
0.58
0.50
0.08
0.00
0.00

10.25
8.92
7.83
5.92
4.75
3.42
3.08
2.92
2.00
1.67
1.25
1.00
0.50

20.75
19.25
17.92
17.08
15.17
12.33
11.67
10.75
9.00
8.83
8.08
8.17
7.70

18.17
16.67
14.75
12.75
9.92
6.50
6.17
5.75
4.00
4.17
3.33
3.25
2.80

11.75
11.08
9.75
8.58
6.58
4.25
4.08
3.75
2.58
2.42
2.08
1.83
1.60

12.83
9.75
8.58
8.00
7.33
5.58
5.42
5.08
4.00
4.17
3.67
3.67
3.25

4.50
4.17
3.58
3.08
2.42
1.75
1.50
1.42
1.08
1.00
0.67
0.58
0.50

9.33
8.58
7.92
6.83
5.58
3.42
2.92
2.42
1.50
1.33
1.00
0.58
0.50

12.17
11.00
9.83
8.33
6.83
4.00
3.42
2.75
1.58
1.42
1.17
0.92
0.10

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

Appendix B. Rheological data

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

223

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

S13

S14

S15

S16

S17

S18

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
510.66
340.44
170.22
10.21
5.11

26.6
19.4
16.9
13.3
5.6
3.6

22.0
16.9
14.8
11.8
4.6
3.1

20.4
15.8
13.8
11.2
4.1
2.6

16.4
12.3
10.2
8.2
2.6
2.0

14.3
10.7
9.2
7.2
2.0
1.5

27.1
17.4
14.8
10.2
4.1
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
765.60
510.40
340.27
238.19
153.12
119.09
85.07
51.04
17.01

23.46
21.07
17.72
14.84
12.93
10.05
8.62
8.14
5.75
4.31

46.44
42.61
37.35
32.56
28.25
23.46
21.55
18.67
14.84
10.53

78.52
72.78
63.20
55.06
47.40
40.22
36.39
32.56
25.86
17.72

References
Alderman, N.J., Meeten, G.H., Sherwood, J.D., 1991. Vane rheometry
of bentonite gels. J. Non-Newton. Fluid Mech. 39, 291310.
Al-Zahrani, S.M., 1997. A generalized rheological model for shear
thinning fluids. J. Pet. Sci. Eng. 17, 211215.
American Petroleum Institute Specifications 13I, 2000. Recommended
Practice Standard Procedure for Laboratory Testing Drilling
Fluids.
Barnes, H.A., 1995. A review of the slip (wall depletion) of polymer
solutions, emulsions and particle suspensions in viscometers: its
cause, character and cure. J. Non-Newton. Fluid Mech. 56,
221251.
Barnes, H.A., 1999. The yield stress a review or
everything flows? J. Non-Newton. Fluid Mech. 81, 133188.
Barnes, H.A., Nguyen, Q.D., 2001. Rotating vane geometry a
review. J. Non-Newton. Fluid Mech. 98, 114.
Barnes, H.A., Walters, K., 1985. The yield stress myth? Rheol. Acta
24, 323326.
Beck, F.E., Powell, J.W., Zamora, M., 1995. The effect of rheology on
rate of penetration. Paper SPE/IADC 29368 Presented at the SPE/
IADC Drilling Conference, Amsterdam.
Becker, T.E., Morgan, R.G., Chin, W.C., Griffith, J.E., 2003. Improved
rheology model and hydraulic analysis for tomorrow's wellbore

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
pH on rheological properties of purified sodium bentonite
suspensions. J. Colloid Interface Sci. 218, 442455.
Bingham, E.C., 1922. Fluidity and Plasticity. McGraw-Hill, New
York.
Bird, R.B., Dai, G.C., Yarusso, B.Y., 1982. The rheology and flow of
viscoplastic materials. Rev. Chem. Eng. 1 (1), 170.
Bleler, R., 1990. Selecting a drilling fluid. J. Pet. Technol. 42,
832834.
Blick, E.F., 1992. Non-Newtonian Fluid Mechanics Notes, Engineering Library. University of Oklahoma, Norman, OK.
Borgia, A., Spera, F.J., 1990. Error analysis for reduction noisy widegap concentric cylinder rheometric data for nonlinear fluids: theory
and applications. J. Rheol. 34 (1), 117136.
Bourgoyne, A.T., Chenevert, M.E., Millheim, K.K., Young Jr., F.S.,
1991. In: Evers, J.F., Pye, D.S. (Eds.), Applied drilling
engineering. SPE Textbook Series, vol. 2. Richardson, TX.
Briscoe, B.J., Luckham, P.F., Ren, S.R., 1994. The properties of drilling
muds at high-pressures and high-temperatures. Philos. Trans. R.
Soc. Lond. Ser. A: Math. Phys. Sci. 348, 179207.
Casson, N., 1959. Flow equation for pigment oil suspensions of the
printing ink type. In: Mills, C.C. (Ed.), Rheology of Disperse
Systems. Pergamon Press, Oxford.

224

V.C. Kelessidis et al. / Journal of Petroleum Science and Engineering 53 (2006) 203224

Chen, D., 1986. Yield stress: a time dependent property and how to
manage it. Rheol. Acta 25, 542554.
Clark, R.K., 1994. Impact of environmental regulations on drilling
fluid technology. J. Pet. Technol. 46, 804809.
Darby, R., 1985. Couette viscometer data reduction for materials with
a yield stress. J. Rheol. 29, 369378.
De Kee, D., Chan Man Fong, C.F., 1993. Letter to the Editor: a true
yield stress? J. Rheol. 37, 775776.
Eckel, J.R., 1967. Microbit studies of the effects of fluid properties and
hydraulics on drilling rate. J. Pet. Technol. 541546 (April).
Elward-Berry, J., Darby, J.B., 1992. Rheologically stable, nontoxic,
high temperature water-base drilling fluid. Paper SPE 24589
Presented at the 67th Annual Technical Conference and Exhibition
of the Society of Petroleum Engineers, Washington.
Evans, I.D., 1992. On the nature of the yield stress. J. Rheol. 36,
13131316.
Fordham, E.J., Bittleston, S.H., Tehrani, M.A., 1991. Viscoplastic flow in
centered annuli, pipes and slots. Ind. Eng. Chem. Res. 29, 517524.
Gallino, G., Maglione, R., 1996. L'applicazione del Modello di
Herschel and Bulkley alle Attuali Teorie per l'Ottimizzazione della
Pulizia del Foro. Proceedings of the Fourth National Conference
on Applied Rheology, Vico EquenseItaly.
Gallino, G., Guarneri, A., Poli, G., Xiao, L., 1996. Scleroglucan
biopolymer enhances WBM performances. Paper SPE 36426,
Presented at the SPE Annual Technical Conference and Exhibition,
Denver-USA.
Govier, G.W., Aziz, K., 1972. The Flow of Complex Mixtures in
Pipes. Krieger, Malabar, FL.
Graves, W.G., Collins, R.E., 1978. A new rheological model for non
Newtonian fluids. Paper SPE, vol. 7654.
Gray, H.C.H., Darley, G.R., 1980. Composition and Properties of OilWell Drilling Fluids. Gulf Pub. Co., Houston.
Grinchik, I.P., Kim, A.K., 1974. Axial flow of a non-linear viscoplastic
fluid through cylindrical pipes. J. Eng. Phys. 23, 10391041.
Gucuyener, I.H., 1983. A rheological model for drilling fluids and
cementing slurries. Paper SPE 11487 Presented at the Middle East
Oil Technical Conference, Manama-Bahrain.
Guillot, D., 1990. Rheology of well cement slurries. In: Nelson, E.B.
(Ed.), Well Cementing. Schlumberger Educational Services,
Houston.
Hanks, R.W., 1979. The axial flow of yield pseudoplastic fluids in a
concentric annulus. Ind. Eng. Chem. Process Des. Dev. 18, 488493.
Hanks, R.W., 1989. Couette viscometry of Casson fluids. J. Rheol. 27,
16.
Hartnett, J.P., Hu, R.Y.Z., 1989. Technical note: the yield stressan
engineering reality. J. Rheol. 33, 671679.
Helland, I., 1988. On the structure of partial least squares regression.
Commun. Stat., Simul. Comput. 17, 581607.
Hemphil, T., Campos, W., Tehrani, M.A., 1993. Yield power law
model accurately predicts mud rheology. Oil Gas J. 91, 4550.
Herschel, W.H., Bulkley, R., 1926. Konsistenzmessungen von
Gummi-Benzollosungen. Kolloid-Z. 39, 291300.
Kelessidis, V.C., Mihalakis, A., Tsamantaki, C., 2005. Rheology and
rheological parameter determination of bentonitewater and
bentonitelignitewater mixtures at low and high temperatures.
Proceedings of the 7th World Congress of Chem. Engr., Glasgow.
Krieger, I.M., 1968. Shear rate in the Couette viscometer. Trans. Soc.
Rheol. 12, 511.
Maglione, R., 1999. The drilling well as viscometer: the route towards
new drilling frontiers. Proceedings of the Southern Europe
Conference in Rheology, Sangineto, Italy.

Maglione, R., Ferrario, G., 1996. Equations determine flow states for
yield-pseudoplastic drilling fluids. Oil Gas J. 94, 6366.
Maglione, R., Robotti, G., 1996. A numerical procedure for solving a
non-linear equations systems for determining the three rheological
parameters of a drilling mud from experimental data. Proceedings
of the Fourth International Conference on Integral Methods in
Science and Engineering, Oulu, Finland.
Maglione, R., Romagnoli, R., 1998. The role of rheology optimisation
in the drilling mud design. GEAM Bull. 94, 123132.
Maglione, R., Romagnoli, R., 1999. Idraulica dei Fluidi di
Perforazione. Edizioni Cusl, Torino.
Maglione, R., Ferrario, G., Rrokaj, K., Calderoni, A., 1996. A new
constitutive law for the rheological behaviour of non Newtonian
fluids. Proceedings of the XIIe Congrs International de Rheology,
Quebec-Canada.
Maglione, R., Guarneri, A., Ferrari, G., 1999a. Rheologic and
hydraulic parameter integration improves drilling operations. Oil
Gas J. 97, 4448.
Maglione, R., Robotti, G., Romagnoli, R., 1999b. Drilling hydraulic
optimization: how managing it. GEAM Bull. 96, 2330.
Maglione, R., Robotti, G., Romagnoli, R., 2000. In-situ rheological
characterization of drilling mud. SPE J. 5, 377386.
Merlo, A., Maglione, R., Piatti, C., 1995. An innovative model for
drilling fluid hydraulics. Paper SPE 29259 Presented at the AsianPacific Oil and Gas Conf., Kuala Lumpur, Malaysia.
Miano, F., Rabaioli, M.R., 1994. Rheological scaling of montmorillonite suspensions: the effect of electrolytes and polyelectrolytes.
Colloids Surf., A Physicochem. Eng. Asp. 84, 229237.
Mihalakis, A., Makri, P., Kelessidis, V.C., Christidis, G., Foscolos, A.,
Papanikolaou, K., 2004. Improving rheological and filtration
properties of drilling muds with addition of Greek lignite.
Proceedings of the 7th National Congress on Mechanics, Chania,
Greece.
Mnatsakanov, A.V., Litvinov, A.I., Zadvornykh, V.N., 1991. Hydrodynamics of the drilling in deep, thick, abnormal pressure
reservoirs. Paper SPE/IADC 21919 Presented at the Drilling
Conference, Amsterdam.
Nguyen, Q.D., Boger, D.V., 1987. Measuring the flow properties of
yield stress liquids. Annu. Rev. Fluid Mech. 24, 4788.
Ohen, H.A., Blick, E.F., 1990. Golden search method for determination of parameters in RobertsonStiff non-Newtonian fluid model.
J. Pet. Sci. Eng. 4, 309316.
Press, W.H., Flannery, B.P., Teukolsky, S.A., Vetterling, W.T., 1992.
Numerical Recipes in Fortran, 2nd edition. Cambridge University
Press, Cambridge.
Robertson, R.E., Stiff Jr., H.A., 1976. An improved mathematical
model for relating shear stress to shear rate in drilling fluids and
cement slurries. SPE J. 16, 3136.
Schurz, J., 1992. A yield value in a true solution? J. Rheol. 36,
13191321.
Shulman, Z.P., 1968. On Phenomenological Generalization of
Viscoplastic Rheostable Disperse System Flow Curves, vol. 10.
Teplo-Massoperenos, Minsk.
Turian, R.M., Ma, T.W., Hsu, F.L.G., Sung, D.J., 1997. Characterization, settling and rheology of concentrated fine particulate mineral
slurries. Powder Technol. 93, 219233.
Walker, R.E., Mayes, T.M., 1975. Design of muds for carrying
capacity. Paper SPE 4975, J. Petr. Techn. 27, 893-900.
Yeow, Y.L., Ko, W.C., Tang, P.P.P., 2000. Solving the inverse problem
of Couette viscometry by Tikhonov regularization. J. Rheol. 44,
13351351.

Potrebbero piacerti anche