You are on page 1of 26

This article was downloaded by: [Clemson University]

On: 28 March 2011


Access details: Access Details: [subscription number 932494102]
Publisher Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-
41 Mortimer Street, London W1T 3JH, UK

Journal of Hydraulic Research


Publication details, including instructions for authors and subscription information:
http://www.informaworld.com/smpp/title~content=t916282780

Steady, laminar, flow of concentrated mud suspensions in open channel


Philippe Coussota
a
Ingenieur du CREF, CEMAGREF, Division Protection contre les Erosions, Domaine Universitaire, St.-
Marlin-d'Hères, France

Online publication date: 26 November 2010

To cite this Article Coussot, Philippe(1994) 'Steady, laminar, flow of concentrated mud suspensions in open channel',
Journal of Hydraulic Research, 32: 4, 535 — 559
To link to this Article: DOI: 10.1080/00221686.1994.9640151
URL: http://dx.doi.org/10.1080/00221686.1994.9640151

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

This article may be used for research, teaching and private study purposes. Any substantial or
systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or
distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses
should be independently verified with primary sources. The publisher shall not be liable for any loss,
actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly
or indirectly in connection with or arising out of the use of this material.
Steady, laminar, flow of concentrated mud suspensions
in open channel
Ecoulements a surface libre permanents et laminaires
de suspensions boueuses concentrees

PHILIPPE COUSSOT
Ingenieur du CREF,
CEMAGREF, Division Protection contre les Erosions,
Domaine Universitaire, BP 76,
38402 St.-Marlin-d'Hires, France
Downloaded By: [Clemson University] At: 22:33 28 March 2011

ABSTRACT
Flows of mud, in the form of a large amount of more or less natural fine particles (less than 100 u_m) suspended
in water, are often encountered in industry and nature (sewage sludge, submarine landslides, mountain mud-
flows, coal slurries, drilling muds, etc). Data concerning these flows are often empirical. We aim here to
describe laminar, free surface flows of such materials.
On the basis of the majority of Theological results concerning concentrated mud suspensions the constitutive
equation of such fluids can generally be assumed to follow a Herschel & Bulkley model. Consequently, in the
case of uniform flow on an infinitely wide, inclined plane or in a semi-cylindrical channel, the velocity distri­
bution within the fluid can be computed exactly. For uniform flow in open channels with other cross-section
types we propose to determine the discharge equation in the form of a relation between two characteristic non-
dimensional parameters and aspect ratios. In the case of a gradually varying but steady flow on an infinitely
wide, inclined plane we assume that shear stress at the wall, at a specific point, is identical to the value of the
uniform flow for a comparable discharge and height. Using this hypothesis we demonstrate that the possible
flow properties are quite similar to those met in usual free surface hydraulics (super-critical and subcritical
regimes, hydraulic jump, roll waves, etc).
We then present the results of uniform flows in an open channel with two types of cross-sections: rectangular
and trapezoidal ( a = 45°). The channel slope varies mainly in the range [2; 40%], and the discharge is within
the range [0.01; 8 1/s]. The theory appears able to predict relatively well experimental results concerning uni­
form flows along with roll waves occurrence. In a rectangular or trapezoidal channel the assumption of the
infinitely wide open channel is valid as long as the aspect ratio is less than 0.1. For each channel type we pro­
pose empirical expressions for the mean wall shear stress also valid for a higher aspect ratio.

RESUME
Dans l'industrie et dans la nature on rencontre de nombreux ecoulements de melanges "boueux" (boues
residuaires, glissements sous-marins, laves torrentielles, charbons liquides, boues de forage, etc.). Les donnees
concernant ces ecoulements sont souvent empiriques. Le but de cet article est de fournir des outils theoriques
pour les decrire.
Les suspensions boueuses concentrees ont une loi de comportement qui suit en general un modele du type
Herschel & Bulkley. Le profil des vitesses dans une section en travers peut etre determine exactement dans le
cas d'un ecoulement uniforme dans un canal demi-cylindrique ou sur un plan incline. En ce qui concerne les
ecoulements uniformes dans des canaux de forme quelconque on propose de determiner la loi d'ecoulement
sous la forme d'une relation entre deux nombres adimensionnels caracteristiques et des parametres de forme.
Pour decrire les ecoulements graduellement varies sur un plan incline on suppose que la contrainte a la paroi
en un point quelconque est egale a la contrainte a la paroi de l'ecoulement uniforme ayant meme debit et meme
hauteur de fluide locale. Avec cette hypothese on demontre que les caracteristiques de ces ecoulements sont
tout a fait similaires a celles rencontrees en hydraulique a surface libre classique (regimes torrentiel et fluvial,
ressaut hydraulique, vagues deferlantes,...).

Revision received March 31, 1994. Open for discussion till February 28, 1995.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 535


On presente ensuite les resultats d'experiences d'ecoulements en regime permanent uniforme dans des canaux
de deux types: rectangulaire et trapezoidal (a = 45°), avec une pente principalement comprise dans l'intervalle
[2; 40 %] et un debit compris entre 0,01 et 8 1/s. La theorie s'avere capable de predire correctement les resultats
experimentaux obtenus dans le cas d'ecoulements uniformes ainsi que les instabilites (vagues deferlantes)
observees dans certains cas. L'hypothese simplificatrice du plan infini est valide tant que le parametre de forme
est inferieure a 0,1 pour un canal rectangulaire ou trapezoidal. Pour chaque type de section on propose des
expressions empiriques de la contrainte a la paroi valables aussi pour des parametres de forme superieurs.

Introduction
Previous rheological studies concerning concentrated mud suspensions have shown that these
materials are non-Newtonian, viscoplastic fluids [1-7]. Indeed they exhibit a yield stress below
which they are almost rigid and above which they can flow. Generally, for a given material, this
yield stress increases with solid concentration. Because of their high apparent viscosity, natural
flows of concentrated mud suspensions are laminar.
Free surface flows of mud are often encountered in nature, but many industrial applications also
involve free surface flows of yield stress fluids (foodstuffs, paints, fresh concrete). However
Downloaded By: [Clemson University] At: 22:33 28 March 2011

research has mainly concentrated on the problem of determining yield stress fluid properties [8-10],
theoretical developments of constitutive relationships [11-12], and experimental studies of mud
rheology [1-5]. Elementary theoretical studies (steady uniform flow) have been made of free sur­
face flow of Bingham fluids [13-14] or other simple, viscoplastic materials [10, 15] on an infinitely
wide, inclined plane. The work of Kozicki & Tiu 16 provides a general approach for uniform flows
of non-Newtonian fluids in closed conduits or open channels but their results have still to be vali­
dated experimentally. Indeed very little complete experimental research work related to free surface
flow of yield stress fluids has been done. As far as we know, no systematic comparison of theoreti­
cal predictions using fluid behaviour knowledge with experimental observations has been made
even for uniform flows. Experimental work essentially concerns the determination of the transition
zone between laminar and turbulent flow [17-20], or peculiar instability phenomena [21]. How­
ever, either in industrial or natural fields, it would be fundamental to know in what conditions one
can use existing rheological models to predict free surface flows quantitatively. In this paper, we
intend to provide a theoretical framework for the description of laminar free surface flows of muds
and to compare this theory with experiments in simple cases.
In our theoretical section we first discuss the choice of an adequate general constitutive equation
capable of describing the behaviour of most natural, concentrated mud suspensions. Then we exam­
ine the uniform steady flow of a mud suspension in an open channel and we discuss the characteris­
tics of gradually varying flows. For our experiments, presented in the second section, we used
natural mud mixtures with a yield stress in the range [13; 53 Pa]. From our theoretical analysis, it
appears that these materials have such characteristics that our experiments can be considered as
being similar to natural muddy debris flows (with a scale factor between 1/20 and 1/10). Since for
our comparison between theory and experiments the first and very important step is the determina­
tion of rheological parameters, rheometrical results along with experimental procedure and cautions
are extensively presented. Then we compare experimental results of uniform open channel (rectan­
gular or trapezoidal) flows with theoretical predictions using rheometrical results.

536 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


1 Theory
1.1 Preliminaries
1.1.1 C h o i c e of t h e r h e o l o g i c a l e q u a t i o n
Here we consider only "concentrated" mud suspensions, i.e., materials with such a solid concentra­
tion that they exhibit a yield stress due to a continuous network of double-layer (surrounding clay
particles) interactions through the whole fluid. The concentration threshold above which such a
phenomenon occurs is highly dependent on material type [11]. Though some pure clay-water
mixtures exhibit thixotropy [11, 22], this phenomenon seems to be minor for most natural cation-
saturated mud suspensions [23]. For such materials the most simple but also most commonly used
model is the Bingham one. Considering most published results, it appears clearly that, when the
flow curve is determined in a fairly large, shear rate range, a Bingham model is inadequate. Indeed
mud mixtures are generally shear-thinning. Various time-independent models (DDK, Casson,
Herschel & Bulkley) can be proposed [15] to describe such a behaviour type. We shall retain the
Herschel-Bulkley [24] model since it appears to be in keeping with the rheometrical measurements
in a very large, shear rate range. In simple shear, this model expresses as follows:
Downloaded By: [Clemson University] At: 22:33 28 March 2011

x = t(xc + K\y\") w h e n y * 0 (e = 4-1 (la)


V \j\J

|x| <xc when y = 0 (lb)

where x is the shear stress, y the shear rate, and xc, K, and n are positive fluid parameters. Normal
stress differences appear unsignificant 23].

1.1.2 S i m i l a r i t y
We consider the gravity, laminar, flow of a Herschel & Bulkley fluid in an open channel. One can
easily deduce from flow equation that the three non-dimensional numbers governing this flow are:

x w
■ JgL (cos i) c KK VJ

where V and L are respectively reference velocity and length, g the gravity, p the fluid density and i
the channel slope. In order to ensure similarity between two fluids flows (k = 1,2) we should have:

^2 R P21 rr z , P 2 1 U + n/2)
n = n, = n2; —2 = JX; xc2 = X cl -A.; K2 = K.-X (3)
v
i Pi Pi

where pk, Kk, nk, and xck (k = 1,2) are respectively the density and the three Herschel & Bulkley
parameters of fluid k, and with X = L/Lt (here Lk and Vk axe reference length and velocity of flow k).
Thus the similarity conditions impose the use of a second fluid with smaller values of K and xc.
Since for concentrated mud suspensions these two parameters usually vary widely with solid con­
centration [23], it appears relatively easy to follow a "global" similarity for example by using the
same fluid type but at a different concentration and without coarse particles.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 537


1.1.3 R e g i m e
We shall consider flows which are slow enough for turbulence not to appear. No criterion for the
transition from laminar to turbulent regime exists for fluids following a Herschel & Bulkley model.
At the very least we may have an idea of the transition zone from the work of Naik [19] who made
Hanks criterion [25] (for pipe flow) suitable for free surface flows of Bingham fluids:

R„= ^ ( l - 1 . 5 P + 0.5[33) (4)

with

4 j ^ ( 4 * ^ a n d P H.
b
K~' ' 2
TS<2 / . o v33 48000
K ' ' K' (1-P)

where K' and x'c are the Bingham plastic viscosity and yield stress of the fluid, and RH the hydrau­
lic radius. Though this approach is not adequate for our Herschel-Bulkley fluids it may be used to
Downloaded By: [Clemson University] At: 22:33 28 March 2011

determine approximately how far we are from the transition zone. To do so it is necessary to fit a
Bingham model to the flow curve.

1.2 Steady uniform flow


1.2.1 F l o w on an i n c l i n e d p l a n e
We consider here the uniform laminar flow of an uncompressible Herschel & Bulkley fluid on an
infinitely wide inclined plane with a slope / relative to the horizontal. We use the frame of coordi­
nates (0,x,y,z) such that y is perpendicular to the plane and x directed downwards. Using the mass
equation and the symmetry of the problem regarding each (y,z) plane, we can deduce that the only
non zero velocity component is the component along jc-axis («), which additionally only depends on
v. This velocity distribution corresponds to a simple shear. Then we deduce that the only non zero
components of the stress tensor are the shear stress (t) and the isotropic pressure (p) which only
depend on y. With the condition of atmospheric pressure (p0) at the free surface, the flow equation
may easily be integrated as:

T = pg(sini) (h-y);p= pg(cosi) (h-y)+p0 (5)

Here we have y = du/dy . Then, assuming the fluid flows downwards (e = 1) and using (1) and (5)
in conjonction with the condition of no wall slip (w(0)=0), we obtain the velocity distribution within
the fluid, which, in non-dimensional variables, may also be expressed as follows:

V(Y) = ( 1 - ( l - y ) 1 + '") w h e n K < l (6a)

V(Y) = 1 when 1 < Y< -^— ( 6b )


&— 1
where

w m+\ y „ pgh(sini)
V = —-u (v), Y = — and G = *-£—i -
il + m)
ay0 ya *c

538 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


with
x
c , (pgsin;
m= - 4 — and a = ^ —
pg(sini) v k

The equations (6) are relevant only when G > 1, i.e. when there is a flow. With these non-dimen­
sional variables there exists an unsheared zone, which is often referred to as "a plug", beginning at
Y= 1 and whose thickness is V{G - 1). The form of velocity distribution for different values of n is
shown in Figure 1. For a given fluid, from one flow to another, in this non-dimensional representa­
tion, only the plug height varies. It is noteworthy that when m is large it would be possible to
approximate this distribution by a straight line with a slope of (1+m) and a thicker, non-sheared
region. When m is smaller than 1 it would be possible to approximate the velocity distribution by a
straight line starting from zero and going to the base of the plug. All other things being equal, the
fluid shear near the wall grows more intense as n decreases.
Downloaded By: [Clemson University] At: 22:33 28 March 2011

-i <- V(Y)
1
Fig. 1. Velocity distribution over the cross-section for a uniform flow on an inclined plane, for different n
values.

These velocity distribution types are quite different from those which can be obtained for water. For
example, for water, if we do not take into account surface tension effects, a flow will occur as soon
as the slope is positive, whatever the fluid depth, whereas for a concentrated mud suspension there
is no flow as long as the fluid depth does not exceed a given value.
The discharge by unit of length through a vertical cross-section may be written in the following non
dimensional form:

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 539


*? = 7—TTT7 ^G2(G-l)°"+l)((m+l)G+l) (7)
(m + 1) (m + 2)
with

Hb =
ikv)
where U is the mean flow velocity. In the case when in = 3, which is of great practical interest (Cf
Section II), we propose the following approximated formula for G as a function of Hb:

G = 1 + 1.93 (ff„)"<0,9) (8)

Equation (8) is an excellent approximation in the (G-l) range [0.1; 10] (CfFigure 2), which corres­
ponds to most laboratory or natural flows.
Downloaded By: [Clemson University] At: 22:33 28 March 2011

1 0*2 10'1 10° 1 01 1 02


(G-1)
Fig. 2 Fit of the empirical formula (3) on the implicit equation (2).

1.2.2 F l o w in a s e m i - c y l i n d r i c a l c h a n n e l
Here we only consider the case when the semi-cylindrical channel is filled with fluid. Then the
exact solution of mass and momentum equations for velocity distribution can be found by assuming
that the only non zero velocity component is the component along the channel axis («). In a cylin­
drical coordinate frame (x,r,Q) with the same x axis as previously, the mass and momentum equa­
tions can then be integrated as:

540 JOURNAL DE RECHERCHES HYDRAUL1QUES, VOL. 32, 1994, NO. 4


p = pg(cosi')r(cos9) + P o ; x=-pg(sin()r (9)

where x is the shear stress due to simple shear along each radius, r the distance between the current
point and the channel axis and 6 is the angle between the radius joining this current point and a ver­
tical plane. Using (1) and (9) with now |y| = du/dr, we deduce that the non dimensional form of
the velocity distribution is still given by (6) with now:

y = 2"(«+l) Y = R-r = R._Jl^ and G = P8*(sin0


ar0(l+m) r0 pg(sinj) 2xc

The non-dimensional discharge equation is given by:

1 ^,-3 , ^ , s ( m + 1) G -^_ 2 (G_1)2_2(G-12 (10)


Hf = ^-[G (G-l)
m+ 1 m+3 m+2 .
Downloaded By: [Clemson University] At: 22:33 28 March 2011

with here

The discharge value given by equation (10) is half the value corresponding to a gravity-driven flow
in a closed cylindrical conduit.

1.2.3 U n i f o r m f l o w in an o p e n c h a n n e l of a n y c r o s s - s e c t i o n
When the channel cross-section has no particular property of symmetry the flow equation can no
longer be solved analytically. Numerical or empirical solutions must be found. However one can
notice that, for both an infinitely wide plane and a semi-cylindrical channel, the non-dimensional
discharge equation takes the form of a relation between Hb and G (where G is the ratio of mean wall
shear stress (uniform in these cases) to yield stress and where Hb is written using the maximum
flow depth). As a consequence, for an open channel of any cross-section, we can expect to find an
equation linking Hb and G in the following form:

Hb = / ( G , a „ a 2 , . . . ) OU

where G and Hb would correspond to the above definitions, and the coefficients (a,, a2, ■■■) would
depend on the geometrical characteristics of the wet section. We must recall that these coefficients
are constant for the infinitely wide plane, and yet for any other channel cross-section they will
depend on some aspect parameters since the form of the wet cross-section generally depends on
flow depth.

1.2.4 F l o w s t a b i l i t y
The phenomenon of roll waves, commonly observed with water [26-27], was also observed with
free surface flow of mud suspensions [21]. In order to establish under which conditions of slope and
discharge our flows are stable we shall use the general and simple approach - valid for any type of
fluid -that has been proposed by Trowbridge [28], His method consists in studying the linear
stability of the mass and momentum equations applied to a portion of fluid in the situation of
uniform flow.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 541


He assumes that 8, equal to

fp (u.n) 2d a / p (U) 2S (where S is the channel section and n the outer unit normal vector)
s
can be approximated by 1. His result involves the mean wall shear stress xp which depends on the
flow depth h and the mean velocity U of the corresponding theoretical uniform flow. The general
condition for flow instability is then:

^UMilj (12)

Hjorth [29] considered that Trowbridge analysis did not take into account that for viscous fluid
flows the growth of the unstable mode is related to the shear at the surface, and he therefore claimed
that this method was unable to predict flow instability for yield stress fluids. Hjorth proposed an
analysis based on a bi-viscous model in two dimensions which led him to conclude that yield stress
fluids are always stable to free surface oscillations. Unfortunately, to simplify the problem, He was
Downloaded By: [Clemson University] At: 22:33 28 March 2011

forced to consider plug thickness as a constant. This hypothesis is obviously an approximation but
its consequences on his reasoning are enormous. Indeed we think that if this hypothesis were satis­
fied, instability would never develop, as the fluid would be forced to flow with a completely con­
stant and rigid plug above it. In these conditions it is not surprising that, following Hjorth's
reasonings, unlike various field and laboratory observations, there seems to be no way for any
instability to develop. Finally we consider that the Trowbridge calculation constitutes a very simple
and good approximation of the reality.
To express (12) in a more explicit way, we shall make an assumption similar to the one for a
gradually varying flow in hydraulics. When flow depth and velocity vary locally and slightly
around their theoretical steady values corresponding to uniform flow, we assume that a correct esti­
mation of the shear stress at the wall, at a specific point, is identical to value of the uniform flow of
comparable discharge and height. Then using (12) with the mean wall shear stress expression given
by the empirical expression (11) one will deduce the stability criteria. The problem that remains is
the validity of the assumption concerning the velocity distribution. In the case of flow on an
infinitely wide plane, this last hypothesis is probably valid since the maximum value of 8 for
uniform flow on an inclined plane is 10/9. Differentiating (7) we deduce:

^-2§+(m+l)
2+m
with

«-'4
Then we get the following criterion for the development of an instability starting from steady
uniform flow:

F i (m+l)(G)2-mG-l (14)

(m+ \)2(G)2 + mG+ 1

542 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


We may note that this result is quite different from the criterion obtained for water flow. For a
turbulent water flow on an inclined plane and using the Manning-Strickler friction formula, the
instability criterion is F > 2.25, whereas our present result for mud suspensions predicts that no
stable flow may be obtained for F values greater than l/(m+l), whose value is generally close to
0.25.

2 Gradually varying, steady, laminar flow in an open channel


We consider here the case of a steady laminar non-uniform flow in an open channel. We assume the
flow varies gradually along the Jt-axis. We shall only develop in detail the case of the infinitely
wide inclined plane. For the general case, assumptions similar to those described in the following
can be made and qualitative results would probably be similar, but analytical solutions may hardly
be found.

2.1 Flow profile equation


Downloaded By: [Clemson University] At: 22:33 28 March 2011

Because of the symmetry of the problem, stress and velocity do not depend on the z-coordinate and
velocity has a z-component equal to zero. Thus the velocity components and isotropic pressure
depend on x and y. In this case the exact analytical solution of the mass and momentum equations
seems very complex in general. We shall circumvent this problem by applying the momentum
equation in integral form to a fluid portion in conjonction with the following simplifying hypo­
theses:
* (a) Because the flow varies gradually along the *-axis the free surface is not that far from a plane
parallel to the wall and we can consider that the pressure field is of hydrostatic type as for a
uniform flow (see section II) on a plane with a slope ('.
* (b) In each cross-section the velocity distribution (along the axis perpendicular to the plane) may
be approximated by the velocity distribution of the uniform flow with the same discharge and
local depth occurring on this slope with an appropriate tangential component of body force.
This appropriate body force component will be written (siny')g.

These hypotheses are quite similar to those used in water hydraulics on smooth slopes and lead to
neglecting the effects of the vertical velocity component. Consider a fluid portion D limited by two
surfaces S, and S2 situated in (y,z) planes and by two parallel and equal surfaces Z and £' situated in
(x,y) planes. The length and width of this portion are respectively dx and L. The free surface is
noted S0 and the surface of contact with the solid plane is S3. Using hypotheses (a) and (b), we find
that equations (5) are valid for Sk (with flow depth hk and sloped) (k = 1,2), where jk is the slope for
which a uniform flow would be obtained with a unit length discharge equal to q and a normal depth
equal to hk. Then, taking into account the no-slip condition at the wall and the free surface condi­
tion, the momentum equation in projection on the x-axis may be expressed as follows:

r j p d v " ] b ( s i n i ) ] + r j ( x . Z . i i ) d d - [ p 2 S 2 - / > , £ , ] = [p(8 2 (I/ 2 ) 2 5 2 -8,([/,) 2 5 1 )] (15)

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 543


where pk (k = 1,2) is the mean pressure exerted on 5*:

K
Pk = p g ( s i n i ) - (16)

Uk is the mean fluid velocity through Sh and 5k(k= 1,2) is defined by:

Jp(u.n) 2 dc
8* = *■* — (17)
P(Uk)%
The second term of the equation (15) is the resistance force exerted on the fluid by the rigid plane.
It arises from the shear rate at the wall. Hypotheses (a) and (ft) led us to neglect second order terms
in the velocity and stress expressions, and thus, when dx tends towards zero, we (dropping indices 1
and 2) obtain the equation giving the variation of flow depth ft as a function of x:

dft _ (sin/) + (sin/(ft, q))


Downloaded By: [Clemson University] At: 22:33 28 March 2011

(C0S )+
' iftdft(6((/)/l)
sin j expressed as a function of U and ft may be found from equation (7) now using j instead of i.
We may note that the equation (18) is similar in form to the equation giving the evolution of water
depth for gradually varying flow. However, at first sight it is not clear whether the flow profiles are
of the same type, since we do not know the various possible values of the denominator and numer­
ator of expression (18) as a function of ft and q.

2.2 Flow profile types


From the results of the previous paragraph along with those of Appendix I, we deduce that:

dft" -h-h.
sign .dx. = sign h-h (19)

where hc is the critical depth and ft„ the uniform flow depth (Cf definitions in Appendix I). This
result is quite similar to what is obtained for open-channel water flows. Additionally the various
flow profiles which may be obtained depending on the relative values of ft, ft,, and hc are formally
quite similar to those encountered in open-channel hydraulics. We shall not reproduce these
different flow profile forms which are presented, for example, in [30] in the case of water flow.
Obviously flow profile calculation using equation (18) is a good approximation of reality as long as
flow depth is not too close to the critical depth (since in this latter zone the flow is no longer
gradually varying). We also can speak of super-critical and sub-critical regimes depending on
whether the flow depth is above or below the critical depth.

2.3 Rapidly varying flow


In a manner similar to hydraulics a kind of "hydraulic jump" will form as soon as a sub-critical
regime is imposed upstream while a super-critical regime is imposed downstream. The equation of
this "hydraulic jump" may be found by applying the momentum equation in integral form to a fluid
portion including the jump. Then we can use the usual hydraulic hypotheses which consist in
assuming that in a zone close enough to the jump the flow is negligibly perturbated by this jump. If
we also neglect gravity and wall resistance we obtain:

544 JOURNAL DE RECHERCHES HYDRAUL1QUES, VOL. 32, 1994, NO. 4


-pg (cost) L^ + pg (cost) L^ = p(82(f/2)252-8l(t/l)25l) (20)

It is then possible to obtain the strict relation giving ht as a function of h2 from this equation by
using equation (17) along with our assumption concerning the velocity distribution over each cross-
section (see above). But it is also possible to assume that the velocity distribution is uniform and
then to obtain the less precise but more practiceful expression:

ji + SF; - 1 (21)
ht - h2
2 -
with
U2
F2= , . .. •
Jgh2(cosi)
Downloaded By: [Clemson University] At: 22:33 28 March 2011

3 Experiments

3.1 Rheology
3.1.1 M a t e r i a l
We used natural clay-water mixtures at different solid concentrations. The clay was taken from a
landslide (Sinard, Isere department, France). The solid particle diameter was less than 40 (i.m. The
non-treated material was first dried at ambient temperature, then mixed again with water at a fixed
solid volumic concentration in a concrete mixer. Afterwards other fluids were obtained by adding
water or dry clay. The mixture pH did not vary during experiments.

3.1.2 A p p a r a t u s
For rheometrical tests fluid samples were taken from the recirculating mixtures during the open
channel flow experiments and then tested with a Rotovisco RV20 (HAAKE) rheometer equipped
with parallel plates (diameter 5 cm; gap of 3 mm). With a rheometer of this type, only strong time-
effects, with characteristic times higher than about 1 s, might be detected, but we did not observe
any such effect. Here we shall neglect other possible slight thixotropic properties.

3.1.3 E x p e r i m e n t a l p r o b l e m s
The experimental precautions required for such fluids are numerous and we used all these tech­
niques developed in [1, 11, 23, 31]. Tool surfaces must be rough so that wall slip may not occur.
One must check that neither sedimentation nor fracture occurs. This was not the case here because
the solid concentration was neither too low nor too high. After these fundamental precautions, with
such materials the main experimental problem originates in edge effects at the free surface along
the periphery of the sample. As noted in [31], with parallel plate geometry, the rheometrical data
were very sensitive to the way one cleans material all around the sample after squeezing it between
tools and before the beginning of the test. From comparison of results obtained with different
rheometer geometries [11] it appears that, to be representative of the behaviour of the material,

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 545


measurements must be made after cleaning. Furthermore, after having cleaned up, during shear, the
free surface appeared curved towards the interior so that the sheared volume did not correspond to
the theoretical one (using the tool dimensions). After measuring the depth of this curvature (about
1 mm) and considering that the curved free surface tends to fix a very small fluid volume within the
sample, we estimated that the real sheared volume corresponded in general to a cylindrical fluid
portion with a radius equal to 2.35 cm. Though we think this correction provides a good approxi­
mation to reality we believe this rheometrical problem is here the main source of error when
comparing theoretical calculations using rheometrical data and fluid flow experiments in various
conditions.
Furthermore this edge effect appeared to vary during the test: at a given rotation velocity the
measured torque changed slightly (few percents) after a relatively long time (30 seconds) but this is
not due to thixotropic effects since this result was not inverted after a rest. In fact this is due to a
slight edge effect change. This effect was also especially sensitive when changing rotation velocity
during one test from a high level to a very low level. Then the torque decreased slightly after about
15 seconds certainly because the shear in the peripherial zone changed progressively. Indeed, close
Downloaded By: [Clemson University] At: 22:33 28 March 2011

to the curved peripherial free surface, the way the material is strained is also dependent on shear
rate intensity. Additionally it changes with time. In this case we recorded the data corresponding to
the higher torque value that we thought to correspond to the situation for which the edge effect is
the same as for higher rotation velocities.

3.1.4 P r o c e d u r e
The experimental methods for flow curve determination presented in [23] were used. Essentially it
involved imposing a succession of different rotation speeds, in varying sequences. We thus
obtained the flow curve in the form of torque vs peripherial shear rate. To simply take into account
the shear rate heterogeneity within the sample [32], we made corrections to rheological parameters
after having fitted a Herschel & Bulkley model (Cf equation (1)) (assuming systematically n = 1/3
in order to unify results) on the rheogram obtained without corrections. Practically this leads to
simply multiplying by 10/9 the K value which is first obtained (without corrections) and keeping
the same xc value. For the fitting we used the method described in [23]. We also demonstrated in
[23] that this model provides a "yield stress" parameter which is very close to the "true yield
stress", corresponding to the limit for which a clear change in material response to imposed shear
stress is observed (from an essentially elastic flow to an essentially viscous flow). Their yield stress
and high apparent viscosity constitute the main difference of such materials with water.

3.1.5 R e s u l t s
The following empirical law can be proposed to describe the increase of our fluid yield stress with
solid concentration:

Tc = 0.11 exp(23C„) (22)

where Cv is the solid volumic concentration. The apparent viscosity of our materials ranged from 10
to 50 Pa.s for a shear rate of Is 1 , while water viscosity is about 0.001 Pa.s at 25°C. The formula
(22) is formally similar to the result commonly reported in literature [3-4] even when a Bingham
model was used. Additionally K is approximately equal to one third xc (in U.S.I.). In Figure 3,
typical flow curves corresponding to different solid concentrations are shown. One can see that the
correspondence between experimental and theoretical curves is very good as long as the shear rate

546 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


is neither too high nor too low. One may also note slight discrepancies (of 10% maximum) between
the two types of curves at their extremities. Nevertheless the Herschel & Bulkley model provides a
very good approximation to reality in a large shear rate range.
Downloaded By: [Clemson University] At: 22:33 28 March 2011

Fig. 3. Water-Sinard clay mixtures for different solid concentrations: flow curve and fit of an Herschel &
Bulkley model with n = 1/3.

When a few samples of the same material were taken at different times and places during open-
channel tests the differences in the corresponding rheological measurements were of the same order
as the fluctuations of results obtained by repeating these measurements on the same sample. Thus
we can conclude that the mixtures flowing in the recirculating system were homogeneous and con­
stant during one day as regards to the accuracy of our analysis. We also made rheometrical tests at
different temperatures between 10 and 30°C and no significant differences in results were observed.
This result, in agreement with those of [23, 33], led us to not check the temperature (probably close
to the ambient temperature of between 10 and 30°C) during next tests. Finally, considering the
various sources of error in flow curve determination we estimate the maximum relative error (in
terms of shear stress) committed in using the fitted rheological model is about 20 %.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 547


3.2 Open channel experiments
3.2.1 E x p e r i m e n t a l procedure
3.2.1.1 INSTALLATION
The recirculating system was composed of a wide rectangular channel whose width could be varied
between 0 and 60 cm and whose length was 8 meters. Two additional uniform flow tests were done
using very steep (above 100%), short planes. The material was recirculated by means of a dia­
phragm pump so that a discharge of up to 8 1/s could be supplied to the upstream hopper at the top
of the channel. We could regulate the flow by changing the opening of the valve at the bottom of
the upstream hopper. The material flowing in the channel arrived in the downstream hopper from
which it was taken again by the pump. The channel slope could be varied in the range [0; 65%].
We made experiments with channel walls (bottom and sides) which were either rough (expanded
metal (diamond-shaped ; side lengths: 1/0.6 cm) with an equivalent roughness of 0.6 cm) or smooth
(plywood).
Downloaded By: [Clemson University] At: 22:33 28 March 2011

3.2.1.2 MEASUREMENTS AND ERRORS


Discharge control was achieved using an additional retention system installed in the channel just
after the hopper exit. It consisted in a dam with a bottom valve. Keeping the dam almost full and
regulating its exit discharge, we avoided most discharge rate fluctuations originating in hopper exit
or due to surges from the pump. Then the obtained discharge range was [0.01; 8 1/s].
Discharge measurements were made by measuring the time necessary to fill up a calibrated volume
(10 liters) placed at the downstream exit of the channel and receiving the whole material flowing in
the channel. This measurement was done many times during steady flow in order to check dis­
charge stability and also diminish measurement errors. We estimate the error on these measure­
ments was about 10%. The slope was chosen arbitrarily before every test and measured directly
with a clinometer. The error originating from this method of measurement is about 5%. In com­
parison the errors for other geometrical parameters are negligible.
To measure flow depth at any point of the channel the first method consisted in plunging a small
graduated ruler vertically into the flowing fluid (thickness: 0.5 mm; width: 1 cm). Because of its
small dimensions and the fluidity of the flowing materials this ruler did not significantly disturb
flow. A second method consisted in measuring the distance between the constant channel side wall
height and the free surface of the fluid (a needle was lowered until it touched the free surface).
These two methods provided very close results so that we concluded that very small errors were
made on these measurements. Additionally for all our tests the maximum difference between flow
depth in the middle of the flow and flow depth near the channel sides was 1 mm. We consider that
flow depth measurement error due to this effect and slight global space and time fluctuations is
about 5 %.
To determine the influence of measurement errors on the quality of the comparison between
experiments and theory, we can consider the effect of experimental errors on uniform flow depth
calculation. This can be done by differentiating the flow discharge equation (7). Neglecting density
and gravity fluctuations and assuming that the maximum relative errors on i, xc, K, and U estimated
above take place at the same time, we find that the maximum error on hn calculation is about 25 %.
If we add this value to the relative error committed on hn measurement we find the maximum
uncertainty in the comparison between theory and experiments is about 30 %.

548 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


3.2.1.3 METHODOLOGY
For each material we imposed various discharges at various slopes and width (or basis for the
trapezoidal channel). Each steady flow was clearly gradually varying close to the channel upper and
lower extremities. Additionally under some particular conditions a hydraulic jump was observed.
These characteristics will be presented and analyzed in a subsequent paper. Here we only report
measurements of the corresponding uniform flow depths. These depths corresponded to the flow
depth which did not vary significantly along a distance equal to at least 40 times this depth.

3.2.1.4 PERTURBING EFFECTS


No settling effect was apparent during channel flow experiments. To avoid any slow segregation
effect we regularly raked the channel to remix all the material. Since a reference line drawn at the
free surface deformed continuously we are sure that neither slip along channel sides nor vertical
fracture occurred. Additionally no significant difference between uniform flow depths in the
smooth and rough channels was observed. This proves that no significant bottom wall slip
occurred.
Downloaded By: [Clemson University] At: 22:33 28 March 2011

3.2.2 R e s u l t s
All our results are presented in Table 1. On relatively gentle slopes the flow was stable. In this case
it was possible to measure the uniform flow depth far enough from channel edges. For these flows
no turbulence was visible at the free surface. Indeed a reference line drawn at the surface deformed
continuously without perturbations. This result may be extended to the whole flow. Let us consider
the extreme following case of our less viscous fluid (xc = 13.7, K = 4.5) flowing at a discharge of
5.27 1/s in a rectangular channel whose width is 17 cm. If we fit a Bingham model on its rheogram
in the shear rate range [10; 100 s^1], we obtain its Bingham rheological parameters: 23.7 Pa and
0.108 Pa.s. The critical number for Rb is then 5596 according to criterion (4) whereas here Rb is
equal to 1108. This means that, except in some cases (like this one) for which we could be close but
still before turbulence transition, all our flows were quite laminar.
Starting from a stable flow, keeping a constant discharge and increasing the slope, we observed
systematically that the flow became unstable above a certain slope which was a function of the
discharge. The phenomenon that we could observe was quite similar to the roll waves which can be
usually observed in water flows [27] under some specific conditions of slope and discharge. In the
case of mudflows it has been already observed in a laboratory [21] and in the field. Let us remark
that, for our mud suspensions, even within the rapid rolls, the flow was laminar. When no external
disturbance was applied to the flow, small waves appeared at the free surface at some distance
downstream. They flowed down growing progressively to become roll waves occupying the whole
depth. These roll waves had a larger depth and a higher velocity than the corresponding (with the
same mean discharge) uniform flow. They were separated by zones of almost no flow. The distance
between each of these rolls seemed to fluctuate around a mean value depending on the discharge,
slope and fluid characteristics. In these cases, when we created a small perturbation upstream it also
degenerated rapidly into a specific roll wave. We did not make a systematic study of the flow
characteristics of the unstable situations but only recorded the cases for which instability appeared.
A flow was considered to be unstable when, without additional action, a clear irregular discharge
was observed at least at the downstream channel exit, otherwise it was considered as stable. It is
obvious that this method may lead to some error because in some cases the channel was not long
enough for instability to develop clearly.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 549


Table 1. Material and flow characteristics for our channelized mud flows tests.
Rectangular channel Trapezoidal
Stable Unstable channel
(L=0,6 m) (o=45°)
Material L Q <WL Q Material B Q i hne/B
W 0/s) (%) (m)
xc=13.7 Pa ,.w...
0.6 0.42 22.0 0.018 tc=13.7 Pa 0.18 0.55 20.0 0.081
K=4.5 0.6 0.38 20.0 0.020 K=4.5 0.18 0.59 7- 0.092
Cv=20.5% 0.17 0.55 8.0 0.182 Cv=20.5% 0.18 0.62 • 0.097
0.17 5.27 12.0 0.182 0.18 0.59 • 0.119
0.17 1.24 5.0 0.347 0.18 0.55 0.124
0.17 4.65 6.0 0.347 0.18 1.65 • 0.184
0.17 1.19 4.0 0.424 0.18 3.76 0.238
0.17 4.74 4.0 0.535 0.18 0.80 • 0.281
0.17 1.34 3.0 0.594 0.05 2.06 2 • 0.440
0.17 1.62 3.0 0.641 0.05 2.01 20. 0.480
xc=14.4 Pa 0.17 3.45 27.5 0.088 3.2 14 0.18 3.79 3.0 0.497
K=5.3 0.17 4.35 23.0 0.106 2 15 0.05 2.06 •o 0.580
Cv=21.1% 0.17 4.76 18.0 0.106 0.05 2.04 0.0 0.760
0.17 0.55 15.0 0.112 0.05 1.93 .0 1.500
0.17 4.55 15.0 0.129 0.05 1.75 _ 2 . 0 3.000
0.17 5.26 19.0 0.135 0.13 1.01 0.146
xc=16 Pa T.
Downloaded By: [Clemson University] At: 22:33 28 March 2011

0.17 0.73 12.0 0.141 K=4 0.13 3.06 13.5 0.162


0.17 1.49 12.0 0.153 Cv=20.9% 0.13 2.05 10. 0.208
0.17 1.45 10.0 0.176 0.13 2.04 »• 0.300
0.17 1.82 10.0 0.194 0.13 1.35 y 0.462
0.17 2.22 6.0 0.335 xc=20.4 Pa 0.17 1.82 6.5 0.286
0.17 2.70 6.0 0.353 K=4.8 0.17 4.00 6.5 0.303
xc=15 Pa 0.17 5.00 16.5 0.141 Cv=24% 0.12 0.42 9.5 0.325
K=5.5 0.17 4.76 13.5 0.165 0.12 5.90 9. 0.350
Cv=21.4% 0.17 4.35 10.5 0.206 0.17 0.64 5.0 0.429
0.17 3.13 7.5 0.329 0.12 1.20 7- 0.433
xc=16 Pa 0.2 0.63 7.0 0.090 0.12 2.27 7.5 0.450
K=4 0.2 0.86 11.0 0.135 0.12 3.70 7.5 0.467
Cv=20.9% 0.2 2.80 11.0 0.140 0.17 2.00 5.0 0.469
xc=16 Pa 0.6 1.37 20.0 0.020 0.09 1.69 9.5 0.533
K=4.5 0.6 2.04 13.0 0.030 0.09 5.90 • 0.578
Cv=21.5% 0.6 3.85 13.0 0.035 0.12 1.59 .0 0.583
0.6 3.12 9.0 0.048 0.09 1.26 • 0.622
0.16 1.73 14.5 0.135 0.12 3.33 .0 0.633
0.16 0.52 10.0 0.190 0.06 0.24 10.0 0.646
0.16 0.07 5.0 0.380 0.12 4.55 .0 0.658
xc=16.8 Pa 0.6 0.27 25.0 0.018 1.3 25 0.06 0.66 10.0 0.677
K=5.7 0.6 0.26 10.5 0.037 1.7 40 0.09 5.90 8.5 0.689
Cv=21.4% 0.6 0.50 10.5 0.040 0.06 5.26 10.0 0.754
0.6 1.20 10.5 0.045 0.09 1.54 7.0 0.756
0.6 0.67 5.5 0.067 0.06 3.45 9.0 0.800
0.6 1.40 5.5 0.073 0.09 4.34 7.0 0.856
xc=17.4 Pa 0.6 1.37 20.0 0.020 3.3 14 0.06 5.90 9.0 0.877
K=5.2 0.6 2.04 13.0 0.030 2.8 15 0.09 1.31 5.0 0.889
Cv=21.8% 0.6 3.85 13.0 0.035 2.8 16 0.09 2.08 •0 0.956
0.6 1.73 14.5 0.037 2.8 17 0.06 2.85 .0 0.985
0.6 3.12 9.0 0.048 2.8 18 0.06 2.27 ■5 1.231

Tc=21.6 Pa 0.6 0.91 19.0 0.030 2 19 0.06 4.00 6.5 1.308


K=7.5 0.6 0.69 11.5 0.037 1.2 21.5 0.06 1.16 5.0 1.492
Cv=23.8% 0.6 0.55 13.0 0.043 xc=21.6 Pa 0.03 0.08 7.0 2.467
0.1 0.30 6.0 1.070 K=7.5 0.03 1.04 7.0 2.833
t c =26.4 Pa 0.6 0.91 19.0 0.030 Cv=23.8% 0.03 0.23 6.0 3.000
K=9.3 0.6 0.55 13.0 0.043
Cv=24.4% 0.6 0.04 4.35 26.5 1.122
Xc=30 Pa 0.6 1.42 192 0.013 2.5 20 xc=28.2 Pa 0.04 5.88 26.5 1.220
K=12.3 0.6 2.00 239 0.013 2.5 22.5 K=10.5 0.04 2.56 22.5 1.244
Cv=24.4% 0.6 1.42 21.0 0.042 1.5 26 Cv=24.4% 0.04 5.56 26.5 1.244
0.6 2.22 21.0 0.045 1.5 29.5 0.04 3.85 22.5 1.268
0.6 2.22 16.0 0.057 1 33 0.04 5.00 22.5 1.341
0.6 2.50 11.0 0.073 1.5 33 0.04 6.25 22.5 1.366
I
550 JOURNAL RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4
0.6 1.50 6.0 0.115 0.5 36 0.04 3.03 16.0 1.732
0.6 2.75 5.5 0.127 1.5 36 0.04 3.45 16.0 1.780
0.6 3.33 5.5 0.130 0.6 40 0.04 4.00 16.0 1.829
0.14 2.22 21.0 0.241 0.04 2.50 11.5 2.293
0.14 2.00 11.0 0.462 0.04 1.43 7.5 2.561
Tc=32.4 Pa 0.6 0.40 31.0 0.032 Tc=37.2 Pa 0.04 0.00 32.0 0.488
K=12.8 0.6 0.44 10.5 0.072 K=10.5 0.04 0.11 27.0 0.659
Cv=24.5% 0.6 1.00 10.5 0.077 Cv=25.1% 0.04 0.08 24.0 0.927
0.6 2.00 10.5 0.080 0.04 0.91 28.0 1.049
0.6 3.33 10.5 0.085 0.04 0.17 23.0 1.122
t c =37.2 Pa 0.6 0.63 13.5 0.063 0.04 0.83 24.0 1.195
K=13.3 0.6 0.63 13.5 0.063 0.04 0.40 20.0 1.341
Cv=25.3% 0.04 0.08 17.5 1.366
Tc=39 Pa 0.18 2.10 150 0.081 4.1 13 0.04 0.06 13.0 1.463
K=15 0.18 3.33 91.0 0.092 3 14 0.04 0.06 16.0 1.488
Cv=25.4% 0.18 1.29 25.0 0.157 2.6 17.5 0.04 0.05 15.0 1.537
0.18 3.33 30.0 0.157 2.2 20 0.04 0.07 13.5 1.707
Downloaded By: [Clemson University] At: 22:33 28 March 2011

0.18 3.33 25.0 0.173 0.04 0.04 11.5 1.951


0.18 1.48 20.0 0.200
0.18 2.00 21.0 0.205 Tc=48 Pa 0.19 0.16 31.5 0.142
0.18 2.67 21.0 0.216 K=16.8 0.19 1.11 26.5 0.205
0.18 4.00 18.5 0.243 Cv=27.1% 0.19 0.44 21.0 0.226
0.18 1.82 15.5 0.254 0.04 0.12 36.0 0.902
0.18 2.35 15.5 0.259 0.04 0.17 30.0 1.098
0.18 3.08 16.0 0.265 0.04 0.05 21.5 1.341
0.18 2.86 15.5 0.276 0.04 0.15 21.5 1.561
0.18 4.00 16.5 0.276 0.04 0.01 13.0 1.829
0.18 3.33 15.5 0.292 Tc=52.8 Pa 0.05 0.75 22.0 1.120
0.18 3.77 11.0 0.411 K=17.8 0.05 1.00 22.0 1.180
0.18 3.33 10.0 0.535 Cv=26.5% 0.05 2.60 22.0 1.320
0.18 4.00 8.0 0.643 0.05 0.54 17.0 1.440
0.08 2.50 22.5 0.753 0.05 3.33 22.0 1.480
0.08 0.83 17.0 0.824 0.05 2.50 17.5 1.620
0.08 2.50 17.0 1.035 0.05 2.82 14.0 1.940
0.08 0.50 12.0 1.047 0.05 0.81 12.0 1.980
Tc=40.8 Pa 0.17 0.02 15.0 0.206 0.05 2.76 10.0 2.640
K=15.2 0.17 0.12 15.0 0.241 0.05 0.94 7.0 3.100
Cv=25.9% 0.17 1.39 15.0 0.294 0.05 1.79 7.0 3.400
0.17 1.25 15.0 0.312 0.05 1.92 7.0 3.500
0.17 0.14 11.0 0.329 0.05 0.73 4.0 4.200
0.17 0.34 11.0 0.359
0.17 1.35 11.0 0.418
0.17 0.48 11.0 0.376
0.17 0.05 8.0 0.453 7.7 13 t c =18.6 Pa
0.17 0.15 8.0 0.488 4.5 14 K=4.2 D
Tc=45.6 Pa 0.6 1.54 10.6 0.083 4.3 16 Cv=22.9% 1
K=16.7 0.6 0.15 5.7 0.110 6.3 18
Cv=26.6% 3.3 18
Tc=45.6 Pa 0.6 0.26 20.5 0.052
K=18 0.6 0.10 10.0 0.088 5 10.5 xc=20.4 Pa
Cv=27.4% 0.6 0.01 6.5 0.100 4.8 11.5 K=4.8
xc=48 Pa 0.6 1.54 10.6 0.083 4.8 12 Cv=24%
K=12 0.6 0.15 5.7 0.110 6.3 13
Cv=26.6% 5.3 i*
xc=48 Pa 0.6 0.80 35.0 0.042
K=16.8 0.6 0.48 31.5 0.045 2 19
Cv=27.1% 0.6 0.28 23.5 0.050 1.2 21.5
0.6 0.20 13.1 0.072
xc=52.8 Pa 0.6 2.00 20.5 0.065
K=17.8
Cv=26.5%

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4


3.3 Comparison with theory
In the following we shall systematically distinguish the experimental flow depth hne and the
theoretical one hn (using particular rheological model and discharge equation). We shall also
distinguish the case for which the infinitely wide plane assumption appears valid (h,JL<0A) and
the general case. We shall essentially present diagrams giving h,Jhne as a function of aspect ratio
because this procedure allows to compare different theoretical predictions.

3.3.1 U n i f o r m f l o w on an i n f i n i t e l y w i d e plane
3.3.1.1 UNSTABLE FLOWS
In Figure 4 we reported results concerning flows in a rectangular channel (with hJL<0.\) and the
theoretical limit curve (equation (14)) in a (F,G) diagram. In each case the non-dimensional param­
eters were calculated using the normal depth computed from discharge equation (7). There is over­
all good agreement between theory and experiments since no unstable flow occurred below the
theoretical limit. However it is not possible to state that the theory is sound since many stable flow
Downloaded By: [Clemson University] At: 22:33 28 March 2011

points are situated above the theoretical limit. We think this discrepancy is certainly not due to
approximations made in the theoretical analysis of [28]. This is more likely to be due to uncertainty
in experimentally determining flow instability, because of the limited channel length. In order to
progress in this way additional careful and systematic experiments should be done along with a
theoretical approach similar to [34].

10'
Rectangular channel
F (hn/L<0.1)

IO-1H

+ Stable flow
n Roll waves
— Theoretical limit (eq. 14)
10 , 1 1

1 2 G
Fig. 4 Stable and unstable flows and comparison with theoretical predictions.

552 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


3.3.1.2 STABLE FLOW
As long as the channel width is more than ten times the flow depth, the fluctuations of the theoreti­
cal results around the experimental ones are inferior to the maximum error calculated above (30%)
(see Figure 5) and the mean ratio h,Jh„ is close to 1. When the aspect ratio increases above 0.1 the
theoretical flow depth progressively departs from the experimental one. Since here theory neglects
side-wall resistance, it predicts a lower flow depth than observed. For trapezoidal channel flows Fig­
ure 6 proves clearly that the infinitely wide assumption is not valid for aspect ratios greater than 0.1.

Rectangular channel
1,8"
1,6"

1,4"
D
Downloaded By: [Clemson University] At: 22:33 28 March 2011

<D 1,2"
C
D Iff "Bo
g n m rr F='r,i i-nfl
J
l,0 " to n n

0,8" n ^i**W
0,6- n^^b
0,4- - Herschel & Bulkley model

0,2- - Discharge equation (7)


(infinitely wide channel hypothesis)
o.o-
1 0' 10"1 10'
hne/L
Fig. 5 Uniform flow in rectangular channel: comparison of experimental flow depth with the theoretical one
calculated using the infinitely wide channel hypothesis.

Considering the very widespread use of the Bingham model for describing clay water mixture flow
curves [3^1,19,23,35], it is worth examining the error committed when using this model to predict
our uniform flow depth. Since our materials clearly exhibit shear-thinning properties the main
problem is the choice of the shear rate range for fitting this model on data. Considering that, for our
tests, wall shear rates are mainly within [10; 100 s _l ], we first fitted a Bingham model in this range
(see such a typical fitting in Figure 3). The non-dimensional discharge equation is now:

Hb = \G~\G- 1) (2G + 1) (23)

written with the Bingham rheological parameters (m = 1). For aspect ratios less than 0.1 the results
are quite acceptable: except for a particularly slow flow, the maximum error is 30%. This agree­
ment is not surprising because our experimental discharge range is convenient. Comparing stop­
page fluid depths the predictions of this model would be in complete disagreement with reality.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 553


Furthermore, to demonstrate the importance of the shear rate range, we carried out the same opera­
tion as before with a Bingham model fitted to the range 1; 10 s-1]. The results obtained are much
worse: the error is up to 60%. In any case the Bingham model is acceptable for practical applica­
tions (in industry or nature) if fitted to the range corresponding to the shear rate range which is
important for the applications (generally the level of wall shear rate should be considered).

- Herschel & Bulkley model


- Discharge equation (7)
(infinitely wide channel hypothesis)

O
Downloaded By: [Clemson University] At: 22:33 28 March 2011

c
SZ

0,4-

0,2-
Trapezoidal channel
o.o ■ i i 11

1 0 -1 10u 10
hne/B
Fig. 6. Uniform flow in trapezoidal channel: comparison of experimental flow depth with the theoretical one
calculated using the infinitely wide channel hypothesis.

3.3.2 U n i f o r m f l o w i n a r e c t a n g u l a r o r t r a p e z o i d a l channel
3.3.2.1 RECTANGULAR CHANNEL
* Kozicki & Tiu approach.
Here we wish to determine a flow equation capable of predicting reality for a large range of aspect
ratios (at least up to 1). For a rectangular or a trapezoidal channel no such analytical expression can
be found directly from momentum and mass equations. However to solve this problem a theoretical
approach was proposed by Kozicki & Tiu [16]. It consists in assuming that the Rabinovitsch-
Mooney equation, demonstrated to be exact for flow in semi-cylindrical or infinitely wide channels,
is also formally valid for mean variables (mean wall shear stress and mean velocity) for any fluid
flowing steadily in a channel of any cross-section. Then, in this equation, one uses values of the two
parameters (a and b) determined (analytically) for simple (Newtonian) fluids. For a fluid following
a Herschel & Bulkley model flowing in an open channel without wall slip, the Kozicki & Tiu
approach leads to the following implicit equation for the mean wall shear stress xp:

554 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


-rb/a + 3 _ -rb/a + 3 ^ T (Xb/a +2
— Xb/a + 2
)
2U = x;r
b + 3a b + 2a

3T c 2 (T,* / " + l -T* / a + l ) T- 1 (T* /a -T* /fl ) ( 24)


b+a b

For a rectangular channel the values of a and b as a function of the channel section are given in
[16]. With this quite complex model the error is larger than 30% for aspect ratios greater than 0.4.
Thus this approach is not very interesting in our case. This is not really surprising because parame­
ters a and b of the Rabinovitsch-Mooney equation should depend on the form of the velocity distri­
bution. Indeed, when there is no particular channel symmetry this distribution is highly dependent
on fluid behaviour.
Downloaded By: [Clemson University] At: 22:33 28 March 2011

* Empirical wall stress formula.


There seems to exist a simple linear equation relating hn and hne for the aspect ratio between 0.1 and
1. However we prefer to use another approach which consists in searching an empirical expression
for the mean wall shear stress. Such an expression will be particularly useful for theoretical or
numerical analysis of gradually varying or unsteady flows. As suggested in Section I we search a
discharge equation in the form of a relation between Hb and G, where G is equal to the ratio of x^ to
xc. In a (G, Hb0-9) diagram we remark that the experimental points seem to be gathered around two
different straight lines depending on their aspect ratio (greater or less than 0.1). Then we propose
the following empirical expression:

MI (25)
Tp = xc (1 + a (Hbyl0-9)) with a = 1.93 - 0.43 arctang <1
L

For a low enough aspect ratio this gives the approximate discharge equation (8) valid for an
infinitely wide channel. Equation (25) is able to predict all our experimental results within an
acceptable range of error (30%), considering the different possible experimental errors.

3.3.2.2 T r a p e z o i d a l c h a n n e l ( b a s i s B , e d g e s l o p e : 4 5 ° )
A similar analysis leads us to propose the following empirical expression for the mean wall shear
stress:

x„= xc(l+a(Hby(0'9)) with a = 1.93 - 0 . 6 arctang U, 4h Y" (26)


B ) . B<A

This equation is able to predict (but overestimates slightly) all experimental results to within 35%.
In fact the obtained dispersion around mean value is larger than with the discharge equation (7).
More clearly than in the case of the rectangular channel, a relation between experimental and theo­
retical normal depth would give better results but would be less useful.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 555


3.3.2.3 COMMENTS
These empirical formulae are only valid for this type of material (Herchel-Bulkley model with
n = 1/3) and within our experimental ranges of slope, discharge and aspect ratio. With these restric­
tions, equations (25) and (26) can be used to compute the normal depth of any flow in rectangular
or trapezoidal channels. To do so one has to equate these expressions for wall shear stress to the
wall shear stress arising from gravity action (pg^?w(sin /)).

Conclusion
We note that the theoretical prediction of open channel flow using rheometrical measurements was
obtained in our case to within 30%. This apparently bad result is essentially due to rheometrical
measurement uncertainties which are inherent to these tests and especially important for yield stress
fluids [8]. Even if done with many precautions we think that, generally, it is hard to obtain rheo­
metrical results to within 10%. This stresses the importance of accurate rheological studies for a
reliable prediction of open channel flows. This demonstrates again the great difference between
Downloaded By: [Clemson University] At: 22:33 28 March 2011

mud flows prediction and usual hydraulics techniques for which, since one deals with flows at
sufficiently high Reynolds numbers, the influence of the exact fluid behaviour is negligible.
These results provide interesting perspectives for applications in nature or industry since it appears
possible to predict any free surface flow on a wide plane as soon as we know the discharge, the
slope and the fluid parameters (determined with a rheometer). Furthermore, our qualitative results
obtained in the case of gradually varying flows on an inclined plane could be extrapolated to the
case of flow in a channel of any cross-section. The corresponding calculations could be done in a
similar way using an empirical expression for wall shear stress like those determined in our particu­
lar cross-section forms. On the other hand it should also be possible to determine fluid parameters
from experiments in a wide channel using the method presented in [36] and which consists in meas­
uring normal flow depth for different discharge values.

Appendix I
Sign of(dy/dx) in the case of a gradually varying flow on an inclined plane
With the assumptions of the Section 3.1, in the case of a gradually varying flow on an infinitely
wide inclined plane, we can compute the exact expression for 8:

a'v' 0 ( i + m} .
(4 +3m)
8(U)2Lh = J(u.n) 2 do= L h-
s
1+m (2 + m) (3 + 2m)7o

where

PS (sin;) \ K J

A. 1 Normal depth and sign of the numerator of (18)


We will call normal depth the flow depth hn corresponding to a uniform flow for a discharge q and
a slope i. This value may obviously be found by writing dh/dx = 0 in (18). Differentiating (7) we
deduce:

556 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


— (sin/) = ^ = (w+l)(l-5)
2 (A.2)
£2-2£+(m+l)
J. + m
with ^ = y\/h. The expression (A.2) is always negative since x is smaller than 1. Thus (sin f)
decreases when h increases and the sign of [(sin i)-(siny)] is the same as the sign of (h - h„).

A.2 Critical depth and sign of denominator of (18)


If the denominator of (18) can tend to zero the flow profile slope will tend to infinity. We will call
critical depth a flow depth (hc) for which such a phenomenon occurs. Using (A.l) and (A.2) we can
deduce:

l 1 1 (2m + I)
" A^ + B^ + C£+D "
- -±(8(UJh)=I(h)=l-
ghdh g .1 +m\K .(1-U 2 ,
-2% ¥ (m + 1)
L2 + m s
Downloaded By: [Clemson University] At: 22:33 28 March 2011

with:

8m + 14m + 4 ■; B 10m 3 +48m 2 + 68m+28.


(m + 2 ) 2 ( 3 + 2m) (m + 2 ) 2 ( 3 + 2m)

2m + 16m2 + 32m + 18
; D = - ( m + 1) (A.3)
(m + 2) (3 +2m)

It can easily be demonstrated that t, decreases when h increases, all other things being equal. We
can also show that when h tends towards 0, h, tends towards 1 and when h tends towards infinity, ^
tends towards 0. Then it is possible to demonstrate by using (A.2) that when h tends towards
infinity 1(h) tends towards 0 whereas when h tends towards 0,1(h) tends towards minus infinity.
Consequently there exists at least one critical depth. Additionally we can demonstrate that if
n < 3.45 (m > 0.29), 6J(h)IAh is always negative which shows that in this case there exists only one
critical depth. In the other cases theoretically there can be three critical depths. Obviously an accu­
rate analysis of flow stability should be carried out for flow conditions corresponding to these dif­
ferent critical depths. It is very likely that two of them correspond to an unstable situation. Since
such a study is complex and the case n > 3.45 seems to be of little practical interest we shall not
consider this case in detail here and assume that n < 3.45 in the text.
The sign of the denominator of (18) is finally the same as for (h - hc). Practically speaking, as a first
approach to determining this critical depth, we can assume 8 is constant and equal to 1. Then we
directly obtain the usual expression:

K = (A.4)
gL (cost)

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 557


Main notations

B = bottom width for trapezoidal channel


cv = solid volumic concentration
8 = gravity modulus
V = shear rate
h = fluid depth
K = theoretical normal flow depth
rine = experimental normal flow depth
i = slope angle
j = fictitious slope angle providing a uniform steady flow with the same local flow discharge
and depth
K, n = Herschel & Bulkley fluid behaviour parameters
L = Rectangular channel width
n = unit outer normal vector
P = pressure
Downloaded By: [Clemson University] At: 22:33 28 March 2011

Po = atmospheric pressure
q = unit length discharge
p = density
RH = hydraulic radius
?c = yield stress
\ = wall shear stress
u = mean velocity over a cross section
u = velocity
I = non-dimensional plug level
yo = plug level

References
1. COUSSOT, P., (1992), "Rheology of debris flows -Study of concentrated dispersions and suspensions",
Ph.D. thesis, Institut National Polytechnique de Grenoble, Grenoble, France, 420p. (in french)
2. LOCAT, J., and DEMERS, D., (1988), "Viscosity, yield stress, remolded strength, and liquidity index
relationships for sensitive clays", Canadian Geotechnical Journal, No 4, 25, pp.799-806.
3. MAJOR, J.J., and PlERSON, T.C., (1992), "Debris flow rheology: experimental analysis of fine-grained
slurries", Water Resources Research, No 3, 28, pp.841-857.
4. O'BRIEN, J.S. and JULIEN, P.-Y., (1988), "Laboratory analysis of mudflow properties", Journal of
Hydraulic Engineering, No 8, 114, pp.877-887.
5. PHILLIPS, C.J., and DA VIES, T.R.H., (1991), "Determining rheological parameters of debris flow material",
Geomorphology, 4, pp. 101-110.
6. QIAN, N., and WAN, Z., (1986), "A critical review of the research on the hyperconcentrated flow in
China", International Research and Training Centre on Erosion and Sedimentation publication, China.
7. WILDEMUTH, C.R., and WILLIAMS, M.C., (1985), "A new interpretation of viscosity and yield stress in
dense slurries: coal and other irregular particles", Rheologica Acta, 24, pp.75-91.
8. N'GUYEN, Q.D., and BOGER, D.V., (1992), "Measuring the flow properties of yield stress fluids", Annual
Review of Fluid Mechanics, 24, pp.47-88.
9. DE KEE, D., TURCOTTE, G., FILDEY, K., and HARRISON, B., (1980), "New method for the determination of
yield stress", Journal of Texture Studies, 10, pp.281-288.
10. UHLHERR, P.H.T, PARK, K.H., Tiu, C , and ANDREWS, J.R.G., (1984), "Yield stress fluid behaviour on an
inclined plane", In Advances in Rheology, editor B. Mena, A. Garcia-Rejon, C. Rangel-Nagaile, Mexico
City, 2,pp.l83-190.
11. COUSSOT, P., LEONOV, A.I., and Piau, J.-M., (1992), "Rheology of concentrated dispersed systems in a
low molecular weight matrix", Journal of Non-Newtonian Fluid Mechanics, 46, pp. 179-217.

558 JOURNAL DE RECHERCHES HYDRAULIQUES, VOL. 32, 1994, NO. 4


12. MOORE, F., (1959), "The rheology of ceramic slips and bodies", Transactions of the British Ceramic
Society, 58, p.470.
13. HOWARD, C.D.D., (1963), "Flow of clay-water suspensions", Journal of the Hydraulic Division, Proceed­
ings of the American society of Civil Engineers, HY5, pp.89-97.
14. PASLAY, P.R., and SLIBAR, A., (1958), "Flow of an incompressible viscoplastic layer on an inclined
plane", Transactions of the Society of Rheology, II, pp.255-262.
15. DE KEE, D., CHHABRA, R.P., POWLEY, M.B., and ROY, S., (1990), "Flow of viscoplastic fluids on an
inclined plane: Evaluation of yield stress", Chemical Engineering Communications, 96, pp.229-239.
16. KOZICKI, W., and Tiu, C, (1986), "Parametric modeling of flow geometries in non-Newtonian flows", in
Flow dynamic and transport phenomena, Chapter 8, Encyclopedia of Fluid Mechanics, N.P. Cherimis-
sinoff Editor, Gulf Publishing Company, Houston, London, Paris, Tokyo, pp. 199-252.
17. WARD, T.J., and O'BRIEN, J.S., (1980), "Flume mechanics of mud flows", Completion report to National
Science foundation Research Initiation, Fort Collins, Colorado, U.S.A., 130p.
18. ZHANG, H., and REN, Z., (1982), "Discussion on law of resistance of hyperconcentration flow in open
channel", Scientia Sinica, A, XXV, No 12, pp.1332-1342.
19. NAIK, B., (1983), "Mechanics of mudflow treated as the flow of a Bingham fluid", Ph.D. thesis,
Washington State University, U.S.A., 164p.
20. HANKS, R.W., (1986), "Principles of slurry pipeline hydraulics", in Slurry and suspension flow properties,
Chapter 6, Encyclopedia of Fluid Mechanics, N.P. Cherimissinoff Editor, Gulf Publishing Company,
Houston, London, Paris, Tokyo, pp.213-276.
Downloaded By: [Clemson University] At: 22:33 28 March 2011

21. WANG, Z., LIN, B., ZHANG, X., (1993), "Instability of non-Newtonian open channel flow", Contributions
to non-stationary sediment transport, W. Kron, I.H.W., Universitat Karlsruhe, Germany, E, pp. 1-23,.
22. Hu, G., and FANG, Z., (1985), "Antithixotropic model for fluid with hyperconcentration of sediment",
Proceedings of the International Workshop on Flow at Hyperconcentrations of Sediment, Beijing, China.
IRTCES Publication.
23. COUSSOT, P. and PlAU, J.-M., (1993), "On the rheology of mud suspensions", to be printed in Rheologica
Acta.
24. HERSCHEL, W.H., and BULKLEY, R., (1926), "Uber die viskositat und Elastizitat von Solen", American
Society for Testing Material, 26, pp.621-633.
25. HANKS, R.W., (1963), "The laminar-turbulent transition for flow in pipes, concentric annuli, and parallel
plates", American Institute of Chemical Engineers Journal, 9(1), pp.45^18.
26. DRESSLER, R.F., (1949), "Mathematical solution of the problem of roll-waves in inclined open channels",
Communications on Pure and Applied Mathematics, 2, pp. 149-194.
27. DRESSLER, R.F., and POHLE, F.V., (1953), "Resistance effects on hydraulic instability", Communications
on Pure and Applied Mathematics, 6, pp.93-96.
28. TROWBRIDGE, J.H., (1987), "Instability of concentrated free surface flows", Journal of Geophysical
Research, 92 (C9), pp.9523-9530.
29. HJORTH, P.G., (1990), "Stability of free surface sediment flow", Journal of Geophysical research, 95
(Cll),pp.20363-20366.
30. CHOW, V.T., (1959), "Open-channel hydraulics", Mc Graw-Hill, Civil Engineering Series, New York,
Toronto, London.
31. COUSSOT, P., and PlAU, J.M., (1993), "Simple shear rheometry of concentrated dispersions and
suspensions", Les Cahiers de Rheologie, Groupe Francais de Rheologie, XII, pp.1—13. (in french)
32. PIAU, J.M., (1979), "Non-Newtonian fluids", Techniques de l'Ingenieur, A710, 16p., A711, 24p. (in
french)
33. MIGNIOT, C, (1989), "Tassement et rheologie des vases", La Houille Blanche, N«l, pp.11-29, and N»2,
pp.95-111. (in french)
34. JULIEN, P.Y., and HARTLEY, D.M., (1986), "Formation of roll waves in laminar sheet flow", Journal of
Hydraulic Research, 24, No 1, pp.5-17.
35. JOHNSON, A.M., (1970), Physical processes in Geology, Freeman Cooper and Co, 577p.
36. ASTARITA, G., MARRUCCI, and PALUMBO, G., (1964), "Non-Newtonian gravity flow along inclined plane
surfaces", Industrial Engineering and Chemistry Fundamentals, 3, No 4, pp.333-339.

JOURNAL OF HYDRAULIC RESEARCH, VOL. 32, 1994, NO. 4 559

You might also like