Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Minerals Engineering 92 (2016) 168–175

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Granular flows in rotating drums: A rheological perspective


Indresan Govender
School of Engineering, University of KwaZulu-Natal, Durban 4041, South Africa

a r t i c l e i n f o a b s t r a c t

Article history: A review of granular flow in rotating drums, with a specific focus on the underlying rheology, is
Received 22 January 2016 presented. The rich coexistence of flow regimes in tumbling mills – the industrial application of rotating
Revised 14 March 2016 drums – highlight the difficulty in obtaining key flow field measurements like velocity and volume
Accepted 15 March 2016
concentration distributions, with non-invasive techniques proving the most useful. The mixture of
Available online 19 March 2016
experimentally derived scaling laws underscore the difficulty in defining a suitable granular rheology
for tumbling mills. The visco-plastic rheology proposed by Jop et al. (2006) denotes a major step forward
Keywords:
in the understanding of dense granular rheology with the scalar form having some experimental corrob-
Rheology
Inertial number
oration in rotating drums. Unfortunately, the success is militated by the mixed results in subsequent
Scaling laws numerical and experimental studies. More specifically, it fails the full tensorial test with notable lack
Constitutive relations of prediction of the well-known hysteresis between flow initiation and cessation, and the expected phase
Non-invasive measurements transition to cataracting flows. Beyond the zeroth order approximation of the visco-plastic rheology, we
also explore the pragmatic approaches of depth averaged modelling that include kinetic theory based
ingredients to successfully capture the two phase transitions.
Ó 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
2. Flow regimes in rotating drums . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
3. Flow field measurements in rotating drums . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
3.1. Scaling relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
4. Towards a granular rheology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
5. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

1. Introduction Tumbling mills are reported to be <5% efficient in the


conversion of input power to useful comminution energy while
Particulate flow in rotating drums exhibit complex phenomena accounting for more than 60% of the plant’s operating costs
such as avalanching, segregation, mixing, drying, aggregation, (Wills, 1997). Current breakage and transport models in tumbling
comminution (abrasion, attrition and fracture) and convection. mills are purely empirical machine models. While such models
Industrially, the flow of granular material in rotating drums spans provide highly tweaked recipes for interpolating within the
applications of mixing, rotary kilns and comminution. The nature boundary conditions from which they were developed, the ability
of the latter application (tumbling mills) usually requires high to extrapolate beyond the window of design is limited, and argu-
frequency rotations of dense media in drums fitted with radial ably dangerous, in light of depleting ore bodies and tighter envi-
baffles to facilitate the strain rate and cataracting demands of the ronmental regulations. To remain competitive and economically
process, while mixers and kilns typically require an active flowing viable within these stringent and ever changing boundary condi-
layer to mediate the mixing or heat transfer requirements. tions, mining companies are forced to consider new comminution
devices that can potentially increase operational efficiency. Central
to this pursuit is an understanding of the mechanisms governing
E-mail address: indresan.govender@gmail.com breakage and the rheology underpinning granular flow. While

http://dx.doi.org/10.1016/j.mineng.2016.03.021
0892-6875/Ó 2016 Elsevier Ltd. All rights reserved.
I. Govender / Minerals Engineering 92 (2016) 168–175 169

the latter is not directly responsible for comminution, accurate


constitutive relations between the flow stresses and associated
strain rates provide clear indications of damage probability. Recent
advances in non-invasive measurement (Parker et al., 1997;
Govender et al., 2004) have greatly enhanced the ability to distill
scaling laws that govern granular rheology, and ultimately the con-
stitutive relations. On this premise, we expect that a description of
granular rheology is more attainable in the near horizon than a
fundamentally-based micro-mechanical model of fracture, and
therefore, more likely to make a greater impact on comminution
practices in the short term.
This review explores the flow regimes within rotating drums,
the experimental efforts employed in measuring key rheological
ingredients (velocity, volume concentration and flowing layer
depth) of these flow regimes, the associated measurement-based
scaling laws that necessarily bound the constitutive choices to
physical reality, and the recent rheological models that have subse-
quently emerged. We acknowledge the powerful role that numer- Fig. 1. Granular flow in a clockwise rotating drum according to Midi (2004). The
ical modelling techniques, like the Discrete Element Method (DEM) arrows below the top of the surface flowing layer indicate velocities whose profile is
delineated by the dotted line. The linear velocity profile in the surface flowing layer
(Cundall and Strack, 1979; Dury et al., 1998), have to offer; how- is extrapolated along the dashed line to find the intersection with the y-axis.
ever, we exclude detailed exposition in lieu of the fact that these
are well established and validated micro-scale descriptions for pre-
dicting bulk flow behaviour. In fact, DEM data is often repackaged
into continuum (meso-scale) descriptions for the purpose of test- occurring along a line similar to the y-axis shown in Fig. 1
ing granular flow models. (Nakagawa et al., 1993; Yamane et al., 1998; Ding et al., 2001;
Orpe and Khakhar, 2001, 2007; Midi, 2004). The choice of this line,
2. Flow regimes in rotating drums which is essentially along the central region of the bed, is moti-
vated by the homogeneity, maximum thickness of the flowing
Granular flows in rotating drums are often described by a flow- layer and exponential tail, unidirectionality of the flow, and the
ing free surface layer over a densely packed rising en-masse that is slow variation of the velocity profile across the layer interfaces.
considered static relative to the rotating drum. Fig. 1 is a simple Many experimental investigations of rotating drum flows have
illustration (not to scale) of the typical flow regimes studied pri- exploited imaging modalities across the electromagnetic spectrum,
marily in the physics literature and is based on the schematic given ranging from transparent end window photographic filming in the
in Midi (2004). visible spectrum (Rogovin and Herbst, 1989; Santomaso et al.,
Starting at the top of the surface flowing layer (often referred to 2003) to low wavelength gamma ray techniques, Parker et al.
as the free surface layer) and moving into the bed along the nega- (1997). Particle image velocimetry (PIV) has provided useful mea-
tive y-direction, Fig. 1, the velocity in the surface flowing layer surements of the ensemble-averaged streamwise velocity profiles
decreases essentially linearly with depth until very close to the in the fluidised layer of slowly rotating drum flows, Jain et al.
bottom of the flow where the decrease starts to slow down expo- (2002). Nakagawa et al. (1993) used magnetic resonance imaging
nentially with further increases in depth. The exponential tail is a (MRI) of nearly spherical mustard seeds in a smoothly lined hori-
solid-like regime that is characterised by dense quasi-static flows zontal cylinder to measure the velocity and free surface for speeds
in which the deformations are very slow and the particles interact operating in the rolling mode; see Fig. 2c. Morrell (1992) employed
by frictional contact (Roux and Combe, 2002). The linear region is a streak photography of coloured tracers through a transparent end
liquid-like regime that is also densely packed but still able to flow window of a pilot scale tumbling mill to constitute the velocity
like a liquid with particles interacting by both friction and collision profile of steel balls. The resulting linear velocity profile formed
(Pouliquen and Chevoir, 2002; Midi, 2004; Forterre and Pouliquen, the key ingredients to his well-known power draught model.
2008). Henein et al. (1983) and Mellmann (2001) classified granu- Orpe and Khakhar (2001) also employed streakline photography
2 to the flow of mono-sized grains (steel balls, glass beads and sand)
lar flows in rotating drums by the Froude number F r ¼ xg R; see
in slowly rotating drums and successfully measured the free sur-
Fig. 2, where x is the angular speed in radians per second, R face. Positron Emission Particle Tracking (PEPT) has been very suc-
denotes the internal radius of the drum and g is the usual acceler- cessful in studying granular flows in rotating drums that span most
 
ation due to gravity in m=s2 . Cascading (Fig. 2d) and cataracting of the Froude regimes shown in Fig. 2. Parker et al. (1997) used
(Fig. 2e) flows exhibit an additional gas-like regime that is charac- PEPT in smoothly lined drums operated in the rolling-to-
terised by very rapid and dilute flows in which the particles inter- minimally cascading regime to measure a non-linear surface layer
act mainly by collision (Goldhirsch, 2003). Below the quasi-static and an underlying bed (the rising en-masse region) that deviated
regime shown in Fig. 1, the flow is assumed static relative to the from solid body motion due to considerable slip at the drum wall.
rotating drum, i.e. the grains are assumed to be moving with the Ding et al. (2001) overcame the slip problems experienced by
same angular velocity as the drum, and is often compared to a solid Parker et al. (1997) with the use of sand paper (lined along the
plug moving with the rotating drum – the so-called plug flow. inner azimuthal wall) to produce measurements consistent with
Nakagawa (1994) and Nakagawa et al. (1997). The data also
3. Flow field measurements in rotating drums successfully validated their continuum model based on the thin
layer approximation.
Flow field measurements relating to granular rheology are pre- Govender (2005) used bi-planar X-ray imaging to track the 3D
sented herein. The complicated influence of boundary conditions motion of a representative plastic bead moving within a 142 mm
to the flow in rotating drums, especially near the highest and low- diameter experimental tumbling mill containing 6 mm
est points of the bed, has resulted in most reported measurements diameter plastic beads and operated in the cataracting flow regime,
170 I. Govender / Minerals Engineering 92 (2016) 168–175

Fig. 2. Six categories of rotating drum flows in alphabetical order of increasing drum rotational speed or wall friction.

Fig. 3. Illustration of the biplanar X-ray imaging technique (centre image) used to track the 3D motion of a representative particle (6 mm diameter plastic bead with a thin
layer of silver lacquer painted onto the surface) moving within the bulk of the flowing granular material (6 mm plastic beads). The 142 mm diameter perspex tumbling mill
was filled to 40% and rotated at 78.6 rpm. The resulting time-averaged velocity per unit volume of the flow field (right and left images) clearly shows all Froude regimes being
accessed by the tracer (after Govender (2005)).

Fig. 3(centre). Despite the high accuracy trajectory fields reported influence of the baffle geometry – the leading face angle of the baf-
by the authors (Govender et al., 2004), the constraints imposed fles in this case – on the amount of cataracting induced, and high-
by diagnostic X-rays and medical protocols for continuous expo- lights the rich coexistence of Froude regimes that are encountered
sure limited the study to very low density materials and insuffi- in industrial granular flow systems.
cient data for computing key rheological parameters like the Morrison (2012) used PEPT to study granular flows in tumbling
volume fraction distribution ð/Þ. Notwithstanding this limitation, mills. After constituting the residence time fractional distribution
the time-averaged velocity per unit volume of the flow was very (RTFD) in a similar manner to Wildmann et al. (2000), the resulting
accurately calculated from the trajectory ðx; y; z; tÞ of a single tracer distribution was used to measure and numerically model the free
under steady, fully developed flow conditions and the ergodic surface (dashed white line) and equilibrium surface (solid white
assumption (Wildmann et al., 2000). Fig. 3 also clearly shows the line) shown in Fig. 4. The cascading layer (intermediate liquid-
I. Govender / Minerals Engineering 92 (2016) 168–175 171

Emergent from the various measurements of rotating drum


flows are a mixture of scaling laws, not all in agreement with cur-
rent models. For example, the recent visco-plastic rheology of
dense granular flow (Jop et al., 2006) suggests that the depth-
averaged velocity hv i / h
3=2
. Parker et al. (1997) used PEPT mea-
surements in rolling mode flows to conclude hv i / h for the size
0

ratios D=d ¼ ð33; 48; 90Þ. Midi (2004) performed a wide range of
measurements spanning the size ratio range D=d 2 ½5; 2500 to
obtain a scaling of hv i / h. In order to check the existence of con-
stitutive laws Rajchenbach (2000) used two drums (10 cm and
20 cm in diameter) partially filled with metal spheres (1:5 mm or
3 mm) to obtain the same scaling relation as Midi (2004). He
obtained a universal dimensionless value of 0:4 when the shear
Fig. 4. Illustrating the free surface (dashed white line) and equilibrium surface qffiffiffiffiffiffiffiffiffi
d
(solid white line) of a 5 mm glass bead mixture in a tumbling mill operated in the rate was normalised by g sin h
, where h denotes the average flux
cataracting flow regime. The residence time fractional distribution was calculated
angle at the given flow rate. Interestingly, by using the average flux
from PEPT data of a representative 5 mm glass bead tracer and formed the basis for
angles over all flow rates investigated, hhi ¼ 42 , the universal
measuring and numerically modelling the free surface and equilibrium surface qffiffi
(after Morrison (2012)). shear rate becomes c_  0:3 gd. Bonamy et al. (2002) also found a
linear scaling that facilitated closure of the depth averaged equa-
like regime) is separated from the rising en-masse layer (static flow
regime) by a surface of zero velocity termed the equilibrium sur- tions for the size ratio D=d ¼ 150. An attractive consequence of
face1 in the minerals engineering literature (Powell et al., 2003). the linear scaling hv i / h is a constant shear rate
h i
The shoulder denotes the region at the top of the flow where the c_ ¼ hvhi  constant . The attractiveness relates to the fact that most
granules lose contact with the inner drum wall and either cascade constitutive relations for the shear stress s are based on c_ , and
down the free surface layer or get projected into free fall. hence simple constitutive relations can exist. The work of Orpe
and Khakhar (2001) indirectly implies hv i / h along the midsec-
3

tion of the flowing layer for D=d 2 ½20; 800: For flow angles
3.1. Scaling relations
b0  bs  1, where b0 and bs are respectively the repose angles at
midpoint of free surface and static angle of repose, and the exper-
Well formulated scaling laws hold the key to unravelling the
imentally observed linear relation b0  bs / x, their predicted
governing mechanisms of multidirectional flows in rotating drum. h i1=2
sinðb0 bs Þ
In the context of granular flows, the key research questions relate shear rate at midsection c_ 0 ¼ g Md cos bs
(M is an adjustable
to the constitutive choices for the stress tensor. With respect to the h i1=2  
conservation laws that underpin the continuum framework, the parameter) reduces to @h@hv i ¼ Mdgcos
x
bs
. Flux balance Q r ¼ Q f
mathematically allowed phase space of constitutive choices are vi
along the midsection then yields x ¼ 2hh R2
, which when back-
nearly infinite. Robust scaling laws bound the theoretical develop- @hv i
pffiffiffiffiffiffiffiffiffiffi
substituted into c_ 0 gives @h / hhv i. Subsequent integration, then
ment of these constitutive choices to a narrow slice of physical
reality – the physics, if you will. Pragmatically, this is the most yields the desired scaling law. Felix et al. (2007) performed exper-
expedient route to realising new theory of granular rheology. Off iments across a wide range of size ratios, D=d 2 ½47; 7400, and
found hv i / h , where the exponent m 2 ½0:48; 5:2 decreases
m
course, scaling laws can remain forever elusive in the absence of
fundamental flow field measurements like the velocity and volume (almost linearly) with increasing D=d. Variations in m were also
concentration distributions. Combining these measurements observed by Ancey et al. (1999) in the frictional–collisional regime
according to the rules of dimensional analysis, as demanded by of flows down rough, inclined channels: for high inclinations of the
the Buckingham Pi theorem (Bertrand, 1878), is usually the first channel 1 < m < 2 while for gentle slopes m ’ 0, suggesting that
step taken towards uncovering the scaling law. the mean velocity of the flow is constant and hence independent
Most scaling relations reported in the literature are based on of the flowing layer depth. We note here the similarity with the
measurements along the central region of the flowing layer, similar scaling of Parker et al. (1997) at low angles. Pignatel et al. (2012)
to the y-axis shown in Fig. 1. The following assumptions are usu- also found that m varies with size ratio for D=d 2 ½30; 7400. Using
"   # 0:44
ally employed when formulating scaling relations: hv ih
a best fitting power law h
d
¼ 2:86 pffiffiffiffi that was obtained
d dg

(i) No cascading or cataracting. using their entire data set, it is easily shown that m  1:27.
(ii) An essentially constant bed repose. In addition to the lack of a clear scaling law between the flow
(iii) Bed depth ðhÞ is small compared to the drum radius ðRÞ so rate and flowing layer depth, all reported scaling relations are
2
that h  R2 . obtained indirectly as follows: For slowly rotating drums with no
(iv) Rising region exhibits plug flow such that the flux per unit cascading or cataracting motion, the flux per unit length of the
 h i  
2
length of drum Q r ¼ x2 R2  h  x2 R2 . flowing layer Q f ¼ hv ih, where v is the component of the velocity
tangent to the free surface. By assuming that the rising region is
(v) Flux per unit length of drum in the flowing layer is given by
2
Q f ¼ hhv i, where hv i is the depth-averaged, stream-wise plug flow and that the square of the depth h  R2 (R is the drum
velocity in the flowing layer. radius), the flux in the rising region is easily shown to be
 
(vi) Flow rate is influenced by the geometrical parameters D=d Q ¼ x2 R2 . Continuity then allows balance between the flux terms
r 
and W=d, where W is the drums length, D its diameter and Q f ¼ Q r and hv ih / x follows naturally. Subsequent measured
d the particle diameter. correlations between h and x then facilitate scaling laws between
hv i and h. It is easy to speculate why high Froude regimes (cascad-
1
The equilibrium surface projects axially along the length of the drum. Fig. 4 only ing and cataracting) are excluded from nearly all studies:
shows the front view of the surface.
172 I. Govender / Minerals Engineering 92 (2016) 168–175

Table 1
Scaling laws for the flowing layer in rotating drums.

Author Scaling: hv i versus h jc_ j estimates Geometric ratio

hv i / h 0 ¼ ð33; 48; 90Þ


Parker et al. (1997) 0 D
d
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
hv i / h 2 ½20; 800
Orpe and Khakhar (2001) 3 gx D
Md cosðhÞ d
qffiffi
Midi (2004) hv i / h c_  0:5 g D
d
2 ½5; 2500
d
qffiffiffiffiffiffiffiffiffiffiffiffi qffiffi
Rajchenbach (2000) hv i / h D
¼ ð33; 66; 133Þ; hhi ¼ 42
c_  0:4 g sinhhi
d
 0:3 gd d
qffiffi
Bonamy et al. (2002) hv i / h c_ ¼ 34s1  0:6 gd
D
d
¼ 150

Jop et al. (2006) hv i / h


3=2 Depends on Q  W
d
2 ½19; 570
hv i / h
m
Felix et al. (2007) Variable D
2 ½47; 7400
"  0:44 #
d
Pignatel et al. (2012)
v ih
Variable D
2 ½30; 7400
¼ 2:86 hp ) hv i / h
1:27 d
h
d
ffiffiffiffi
d dg

(a) The S-shaped profile of the free surface leads to an ill- Repose scaling laws that capture this hysteresis are thus key to
defined repose angle. realising a granular rheology for rotating drum flows.
(b) The plug flow assumption in the rising region is likely to Notwithstanding the significant progress in the last decade, the
break in favour of richly sheared layers. fact that various scaling relations exists suggest that the underly-
(c) Flux balance between the rising and flowing layer (below ing mechanisms governing granular flows in rotating drums are
the free surface) is violated in the presence of cataracting not yet fully understood. The problem is further compounded by
flows. the limited measurements in the cascading and cataracting Froude
(d) Simple correlations between Q f and drum rotation rate x regimes where the key rheological measures (angle of repose, flow-
may not exist, thereby invalidating the formulation of cur- ing layer depth and depth-averaged velocity) become ambiguous
rent scaling laws between flow depth and depth-averaged or inaccessible to most measurement schemes.
velocity.
4. Towards a granular rheology
Notwithstanding the limitation to low Froude regime measure-
ments, big differences between the various published scaling rela- Beyond the approaches that capture the qualitative features of
tions exist, clearly highlighting the need for further experimental rotating drum (granular) flows in the physics (Rajchenbach,
work and modelling of rotating drum flows. Table 1 summarises 1990; Zik et al., 1994; Elperin and Vikhansky, 1998; Yamane et
the various scaling relations obtained for rotating drum flows. al., 1998; Puri and Hayakawa, 1999; Orpe and Khakhar, 2001;
The dynamic angle of repose ðhÞ in rotating drums (see Fig. 1) Taberlet et al., 2006) and engineering (Hogg and Feurstenau,
have also had mixed scaling relations. At the very low end of the 1972; Harris et al., 1985; Morrell, 1992) literature, the absence of
continuous flow regime, Rajchenbach (1990) found ðh  h0 Þ / x2 a clear rheological description has limited their value. Fortunately
(h0 is the static angle of repose) that differed from the progress towards a constitutive law for dense granular flow has
ðh  h0 Þ / x1:43 obtained by Tang and Bak (1988). Dury et al. emerged in the last decade. Using dimensional analysis of discrete
(1998) used MRI measurements of mustard seeds, end-window simulation data, da Cruz et al. (2005) showed that the shear stress
measurements (by looking through acrylic end caps) and Discrete ðsÞ verifies a friction law with the resulting friction coefficient ðlÞ
Element (DE) simulations to show a linear dependence ðh / xÞ being dependent on a single dimensionless parameter, the
consistent with the continuous flow regime. Yamane et al. (1998) Inertial number ðIÞ:
also found a linear relationship using MRI measurements. In a
study involving peas and rice with size ratio D=d ¼ 22, s jc_ jd
¼ lðIÞ with I ¼ pffiffiffiffiffiffiffiffiffiffiffiffi ; ð1Þ
Khosropour et al. (2000) found that the dynamic angle of repose P P=qm
did not scale quadratically with drum rotation rate but produced
where jc_ j is the norm of the shear rate and ðPÞ the confining pres-
ðh  h0 Þn / x, with n ¼ 1:9 and 2:6 for peas and rice respectively.
sure. By prescribing the inertial number within a homogeneous
The difference, they argued, was due to the size ratio being three
state, two fundamental dimensionless quantities, the volume con-
times smaller than that used by Rajchenbach (1990). Sepulveda
centration ð/Þ and effective friction coefficient ðlÞ, were found to
et al. (2005) reported a linear scaling between the angular span
vary linearly with inertial number:
of the granular bed and angular speed of the drum. The micro-
scopic effects of rough particles was shown to increase the angle /ðIÞ ¼ /max  aI; ð2Þ
of repose by 10 over smooth particles across a wide range of
rotation speeds. Many other repose measurements exist in the lðIÞ ¼ lmin þ bI ð3Þ
literature, but none seem to deal with flows above the rolling
Froude regime where the characteristic S-shaped free surface and where /max ; /min ; a; b are constants. Interestingly, the inertial num-
cataracting streams render the usual definition of the repose angle ber is also the square of the Savage number or Coulomb number
ambiguous. The apparent deficiency in the definition clearly (Savage, 1984; Ancey et al., 1999).
warrants a fresh perspective on repose angles. Finally, intimately Midi (2004) used data from different flow geometries to provide
connected to the angle of repose is the hysteresis between starting a physical interpretation of the inertial number in the dense flow
 
ðhstart Þ and stopping hstop angles, i.e. hstop < hstart , associated with regime ðI > 0:01Þ: the ratio of the inertial time of rearrangement
steady, continuous flows (Dury et al., 1998; Forterre and (confinement time which relates to the confining pressure)
pffiffiffiffiffiffiffiffiffiffiffiffi
Pouliquen, 2008). An exacerbating factor relates to the influence tmic ¼ d qm =p to the time of strain (shear rate or deformation
of finite-sized effects and/or boundary effects on hstop and hstart . time) 1=c_ between contiguous flowing layers of granules. Consis-
I. Govender / Minerals Engineering 92 (2016) 168–175 173

tent with the dimensional analysis of da Cruz et al. (2005) the


solids concentration / and dimensional stress Ps ¼ l must be
slaved to the inertial number such that the resulting granular flow
is completely determined by lðIÞ and /ðIÞ.
Jop et al. (2005) and Pouliquen et al. (2006) showed that Eq. (1)
is compatible with their experimentally determined basal friction
law (Pouliquen, 1999; Pouliquen and Forterre, 2002) for the choice
ðlmax  lmin ÞI
lðIÞ ¼ lmin þ ; ð4Þ
I0 þ I
where I0 ¼ 0:279 is a constant determined by Jop et al. (2005) and
(lmin and lmax ) are respectively constants relating to the minimum
and maximum friction coefficients that bound steady flow. The cor-
responding functional form of /ðIÞ was observed to be linear and fit-
ted as
/ðIÞ ¼ /max þ ð/min  /max ÞI ð5Þ
with typical constants given by /max ¼ 0:6 and /min ¼ 0:4. Jop et al.
(2005) then included this simple rheology, Eqs. (1) and (4), into the
momentum balance equations of surface flows on heaps (inclined at
angle h to the horizontal) with side walls to obtain Eq. (6).
Fig. 5. Variation of the effective friction coefficient lðIÞ with inertial number as
x proposed by Jop et al. (2005) for typical rotating drum data.
tanðhÞ ¼ lðIðxÞÞ þ lw ; ð6Þ
W
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where lðIðxÞÞ is given by Eq. (4) with IðxÞ ¼ jc_ ðxÞjd= /gx cos h now
dependent on depth ðxÞ below the free surface of the heap, tanðhÞ they interpreted as the free fall time scale with no drag. The results
characterises the total effective friction coefficient that accounts showed disagreement with the expected shape of the lðIÞ-
for sidewall and internal frictional effects and W is the width of rheology which they attributed to significant wall friction and a
the heap. Noting the similarity to rotating drums with h represent- logarithmic decay of the stream-wise velocity that does not satisfy
ing the repose of the bed as shown in Fig. 1, W the length of the the constant shear rate assumption of the lðIÞ-rheology unless
drum, and x the depth of the flowing layer, Fig. 5 illustrates the non-local extensions are considered, Kamrin and Kovall (2012).
lðIÞ-rheology described by Eq. (6) using measured (via PEPT2) Consistent with the assumptions governing the tensorial for-
angles that bound steady flow (hmin and hmax ) in tumbling mill exper- mulation, an appropriate procedure for testing the constitutive
iments with I0 chosen similar to that proposed by Jop et al. (2005). sxz jc_ d j
The 3D generalisation of the friction law proposed by Jop et al. relation describing the rheology entails plotting c_ d P against
xz

(2006) can be summarised in terms of the deviatoric stress tensor  pffiffiffiffiffiffiffiffiffiffiffiffi


the inertial number I ¼ jc_ jd P=qm across the entire granular
sij and deviatoric strain rate tensor c_ dij under the assumptions that: flow field and checking whether it follows the form of Eq. (6).
By invoking the univocal relation for the volume concentration
(i) small variations of the volume fraction ð/Þ can be neglected as stipulated by the dimensional analysis of Midi (2004), assump-
(implying that the flow is incompressible), and tion (i) can be relaxed without affecting the consistency of the lðIÞ-
(ii) sij and c_ dij are collinear. rheology. It is useful to note that other univocal relations have had
equal success in describing an effective frictional rheology. In par-
ticular, Hatano’s (2007) dilatancy law: / ¼ /0  aIb where
P jc_ jd a ¼ 0:11; b ¼ 0:56 and /0 ¼ 0:6, has been employed by Lee and
sij ¼ lðIÞ d c_ dij with I ¼ qffiffiffiffiffi ; ð7Þ
jc_ j P
Huang (2012) to described a three dimensional flow down an
qm
inclined plane that exhibited additional kinetic stresses.
where sij ¼ rij þ Pdij is the deviatoric part of the stress tensor, sxz jc_ d j
  qffiffiffiffiffiffiffiffiffiffiffiffi Testing the resulting form of c_ d P against the inertial number
c_ ij ¼ 12 @ i vj þ @ j vi the strain rate tensor, c_ d ¼ 12 c_ ij c_ ij the norm  pffiffiffiffiffiffiffiffiffiffiffiffi
xz

I ¼ jc_ jd= P=qm , is now borne entirely by assumption (ii).


of the deviatoric part c_ dij ¼ c_ ij  12 c_ kk dij ; P ¼ 12 rkk the local pressure
The collinearity assumption – assumption (ii) – is obvious for
and vi the components of the velocity.
the unidirectional flows encountered down inclined planes and
Flow image analysis studies performed by Chou and Lee (2009)
heaps; however, it breaks down for multidirectional flows, like
and Orpe and Khakhar (2007) along the midsection of rotating
the rotating drum (Cortet et al., 2009). In light of multidirectional
drum flows spanning rolling-to-cataracting successfully recovered
qffiffiffiffiffiffiffiffiffiffiffi flow within rotating drums (Cortet et al., 2009) expanded the test
the lðIÞ-rheology for I ¼ c_ g cosðhÞ d
2 ð0; 1Þ with the free surface requirements of the rheology to include:
slope tanðhÞ ¼ lðIÞ. Pignatel et al. (2012) also used flow image
analysis on rolling mode flows in rotating drums to characterise (a) jPsj ¼ lðIÞ,
the lðIÞ-rheology in the presence of lateral wall friction. Difficulties (b) sij and c_ dij have identical principal directions.
in measuring the pressure and parameters relating to the effective
lðIÞ-rheology led them to approximate the total friction coefficient According to Cortet et al. (2009) condition (b) is naturally satis-
by the slope to the free surface at midsection and the inertial num-
fied in unidirectional flows (like the inclined plane and heap flow)
ber by I ¼ c_ t ff , where tff is a derived characteristic time scale which
so verifying Eq. (7) reduces to verifying condition (a) only. In mul-
tidirectional flows, like that encountered in rotating drum flows,
2
Repose angles correspond to the average repose of the S-shaped free surface as Cortet et al. (2009) calculated the angular difference ðdhÞ between
reported in Govender et al. (2014).
174 I. Govender / Minerals Engineering 92 (2016) 168–175

Fig. 6. Effective friction coefficient as a function of inertial number for three drum rotation rates: X ¼ 12 rpm (green points), X ¼ 6 rpm (red points) and X ¼ 2 rpm (blue
points). (a) Illustrates the ratio for the entire flow field with (b) showing the same plot for a logarithmic inertial number scale. (c) Shows the ratio along the central region of
the drum and (d) is the equivalent plot on a linear-log scale. Figures adapted from Cortet et al. (2009). (For interpretation of the references to colour in this figure legend, the
reader is referred to the web version of this article.)

the principal directions of sij and c_ dij to test the collinearity


demanded by condition (b). He found that ðdhÞ ranges from 10
for inertial numbers 5 104 < I < 5 102 , corresponding to
the ‘‘liquid” surface flows, and up to 35 in the deeper part of the
flow, corresponding to smaller inertial numbers
5 106 < I < 5 104 . He therefore concluded that the consti-
tutive relations proposed in Eq. (7) fails the collinearity test over
the whole range of inertial number. This conclusion is further sup-
ported by the important dispersion that results from plotting
sxz jc_ d j  pffiffiffiffiffiffiffiffiffiffiffiffi
c_ d P
against the inertial number I ¼ jc_ jd= P=qm ; see Fig. 6.
xz

Even when the analysis was restricted to data taken along the cen-
ter of the drum, an unexpected kink in the curve, Fig. 6d, shows
that the proposed rheology fails even along the (assumed) well-
behaved central region. Fig. 7. Variation of the effective friction coefficient with inertial number. The lines
Interestingly, the norm of the stress ratio jPsj ¼ lðIÞ follows the are model predictions for three different effective restitution coefficients while the
frictional rheology described by Eq. (6) for the inertial range open circles correspond to the data from Midi (2004). Figure adapted from Lee and
Huang (2012).
I > 104 with a hysteresis-type jump just below I ¼ 104 . The hys-
teresis effect is well-known Forterre and Pouliquen (2008) for continuity at the transition to the inertial flow regime (second
rotating drum flows and highlights a clear limitation of the Jop- turning point in Fig. 7) warrants further study.
rheology to capture the phase transition from solid-like flow to
liquid-like flow. Lee and Huang (2012) addressed this limitation
by employing rate-independent and rate-dependent components 5. Conclusions
for static and kinetic contributions of the shear stress respectively
within a depth-averaged formulation consistent with the Saint– We have reviewed granular flow in rotating drum systems with
Venant hydrodynamic description. Granular kinetic theory (Lun a specific focus on models and measurement that relate to the
et al., 1984; Jenkins and Richman, 1985; Lun and Savage, 1987) underlying rheology. Scaling laws have been identified as a key sig-
was used for the kinetic parts while the static components con- nature of the stress–strain relations governing granular flow and
formed to the friction law of Jop et al. (2006) and the dilatancy rheology. Unfortunately, existing measurement-derived scaling
law of Hatano (2007). For a range of effective restitution values relations appear to contradict current model predictions of the
as suggested by Jenkins and Zhang (2002), the hysteresis at flow same and each other in some cases. Consequently, an appropriate
initiation (first turning point in Fig. 7) and the transition from scaling law for describing typical comminution boundary condi-
liquid-like flow to gas-like flow (second turning point in Fig. 7) tions is not obvious from the literature. We assert that research
was successfully captured. The authors acknowledge that the dis- into experimentally-derived scaling laws for typical comminution
I. Govender / Minerals Engineering 92 (2016) 168–175 175

boundaries conditions are currently possible with available mea- Jenkins, J., Zhang, C., 2002. Kinetic theory for identical, frictional, nearly elastic
spheres. Phys. Fluids 14 (3), 1228–1235.
surement tools like PEPT, PIV and X-rays – this assertion consti-
Jenkins, J.T., Richman, W., 1985. Grads 13-moment system for a dense gas of
tutes our future work. inelastic spheres. Arch. Rat. Mech. Anal. 87, 355–377.
A rich co-existence of flow regimes are qualitatively measured Jop, P., Forterre, Y., Pouliquen, O., 2005. Crucial role of side walls for granular surface
in rotating drums, especially when realistic conditions (like tum- flows: consequences for the rheology. J. Fluid Mech. 541, 167–192.
Jop, P., Forterre, Y., Pouliquen, O., 2006. A constitutive law for dense granular flows.
bling mills) are considered. No current models satisfactorily cap- Nature 441 (8), 727–730.
tures all the observed flow regimes. While the new rheology of Kamrin, K., Kovall, G., 2012. Nonlocal constitutive relation for steady granular flow.
Jop et al. (2006) is a major step forward in describing dry granular Phys. Rev. Lett. 108 (17), 178301–178305.
Khosropour, R., Valachovic, E., Lincoln, B., 2000. Flow and pattern formation in a
flows, it does not cope well with the very low and very high inertial binary mixture of rotating granular materials. Phys. Rev. E 62 (1), 807–812.
numbers. In particular, it fails to capture the hysteresis between Lee, C.H., Huang, C.J., 2012. Kinetic theory based model of dense granular flows
the three flow regimes that delineate granular flows. This poses a down inclined planes. Phys. Fluids 24, 073303.
Lun, C.K.K., Savage, S.B., 1987. A simple kinetic theory for granular flow of rough,
severe limitation when dealing with systems like tumbling mills inelastic spherical particles. Trans. ASME E: J. Appl. Mech. 54, 47–53.
that can span all three regimes when cataracting conditions pre- Lun, C.K.K., Savage, S.B., Jeffrey, D.J., Chepurniy, N., 1984. Kinetic theories for
vail. Despite these limitations, the 3D tensorial formulation of granular flow – inelastic particles in Couette-flow and slightly inelastic particles
in a general flowfield. J. Fluid Mech. 140, 223–256.
Jop et al. (2006) – which is akin to well known visco-plastic rheolo- Mellmann, J., 2001. The transverse motion of solids in rotating cylinders: forms of
gies like the classical Bingham or Herschell–Bulkley fluids – has motion and transition behavior. Powder Technol. 18 (3), 251–270.
greatly facilitated the ability to describe the various flow regimes. Midi, G.D.R., 2004. On dense granular flows. Eur. Phys. J. E 14, 341–365.
Morrell, S., 1992. Prediction of grinding mill power. Trans. Inst. Min. Metall. (Sect. C:
In this regard, the model of Lee and Huang (2012) is particularly
Miner. Process. Extr. Metall.) 101, C25–C32.
interesting. Morrison, A.J., 2012. Using Positron Emission Particle Tracking (PEPT) to Investigate
Going forward, a key research investment into carefully delin- the Motion of Granular Media in a Laboratory-Scale Tumbling Mill. Master’s
eated scaling laws relevant to current comminution practices is thesis, Dept. of Physics, University of Cape Town.
Nakagawa, M., 1994. Axial segregation of granular flows in a horizontal rotating
warranted. We hypothesise that such efforts – which we note as cylinder. Chem. Eng. Sci. 49 (15), 2540–2544.
glaringly missing in the current literature – will ultimately yield Nakagawa, M., Altobelli, S.A., Caprihan, A., Fukushima, E., 1997. NMR measurement
the fundamentally sound engineering rules for safe machine scale and approximate derivation of the velocity depth profile of granular flow in a
rotating, partially filled, horizontal cylinder. Powders Grains 52 (23), 447–450.
up, and indeed for new comminution technologies. Nakagawa, M., Altobelli, S.A., Caprihan, A., Fukushima, E., Jeong, E.K., 1993. Non-
invasive measurements of granular flows by magnetic resonance imaging. Exp.
References Fluids 16 (1), 54–60.
Orpe, A.V., Khakhar, D.V.V., 2001. Scaling relations for granular flow in quasi-two-
dimensional rotating cylinders. Phys. Rev. E 64 (3), 1–13.
Ancey, C., Coussot, P., Evesque, P., 1999. A theoretical framework for very
Orpe, A.V., Khakhar, D.V.V., 2007. Rheology of surface granular flows. J. Fluid Mech.
concentrated granular suspensions in a steady simple shear flow. J. Rheol. 43,
571, 1–32.
1673–1699.
Parker, D.J., Dijkstra, A.E., Martin, T.W., Seville, J.P.K., 1997. Positron emission
Bertrand, J., 1878. Sur l’homogénéité dans les formules de physique. C.R. 86 (15),
particle tracking studies of spherical particle motion in rotating drums. Chem.
916–920.
Eng. Sci. 52 (13), 2011–2022.
Bonamy, D., Daviaud, F., Laurent, L., 2002. Experimental sudy of granular surface
Pignatel, F., Asselin, C., Krieger, L., Christov, I.C., Ottino, J.M., Lueptow, R.M., 2012.
flows via a fast camera: a continuous description. Phys. Fluids 14 (5), 1666–
Parameters and scalings for dry and immersed granular flowing layers in
1674.
rotating tumblers. Phys. Rev. E 86 (1), 001304–001312.
Chou, H.-T., Lee, C.-F., 2009. Cross-sectional and axial flow characteristics of dry
Pouliquen, O., 1999. Scaling laws in granular flows down rough inclined plane. Phys.
granular material in rotating drums. Granul. Matter 11, 13–32.
Fluids 11 (3), 542–548.
Cortet, P.-P., Bonamy, D., Daviaud, F., Dauchot, O., Dubrulle, B., Renouf, M., 2009.
Pouliquen, O., Cassar, C., Jop, P., Forterre, Y., Nicolas, T., 2006. Flow of dense granular
Relevance of visco-plastic theory in a multi-directional inhomogeneous
material: towards a simple constitutive laws. J. Stat. Mech. 2006, 7–20.
granular flow. Eur. Phys. Lett. 88, 14001.
Pouliquen, O., Chevoir, F., 2002. Dense flows of dry granular material. C.R. Phys. 3,
Cundall, P.A., Strack, O.D.L., 1979. A discrete numerical model for granular
163–175.
assemblies. Geotechnique 29, 47–65.
Pouliquen, O., Forterre, Y., 2002. Friction law for dense granular flows: application
da Cruz, F., Emam, M., Prochnow, M., Roux, J.N., Chevoir, F., 2005. Rheo-physics of
to the motion of a mass down a rough inclined plane. J. Fluid Mech. 453, 133–
dense granular materials. Phys. Rev. E 72, 021309.
151.
Ding, Y.L., Seville, J.P.K., Forster, R., Parker, D.J., 2001. Solids motion in rolling mode
Powell, M.S., McBride, A.T., Govender, I., 2003. Application of DEM outputs to
rotating drums operated at low to medium rotational speeds. Chem. Eng. Sci. 56
refining applied sag mills mode. In: Proceedings International Mineral
(5), 1769–1780.
Processing Congress (IMPC) XXII 1, pp. 325–334.
Dury, C.M., Ristow, G.H., Moss, J.L., Nakagawa, M., 1998. Boundary effects on the
Puri, S., Hayakawa, H., 1999. Dynamical behaviour of rotated granular mixtures.
angle of repose in rotating cylinders. Phys. Rev. E 57, 4491–4497.
Physica A 270 (1–2), 115–124.
Elperin, T., Vikhansky, A., 1998. Granular flow in a rotating cylindrical drum.
Rajchenbach, J., 1990. Flow in powders: from discrete avalanches to continuous
Europhys. Lett. 42 (6), 619–623.
regim. Phys. Rev. Lett. 65 (18), 2221–2225.
Felix, G., Falk, V., D’Ortano, U., 2007. Granular flows in a rotating drum: the scaling
Rajchenbach, J., 2000. Granular flows. Adv. Phys. 49 (2), 229–256.
law between velocity and thickness of the flow. Eur. Phys. J. E 22, 25–31.
Rogovin, Z., Herbst, J.A., 1989. Charge motion in a semi-autogeneous grinding mill.
Forterre, Y., Pouliquen, O., 2008. Flows of dense granular media. Annu. Rev. Fluid
Miner. Metall. Process. 6, 18–23.
Mech. 40, 1–24.
Roux, J.N., Combe, G., 2002. Quasistatic rheology and the origins of strain. C.R. Phys.
Goldhirsch, I., 2003. Rapid granular flows. Annu. Rev. Fluid Mech. 35, 267–293.
3 (2), 131–140.
Govender, I., 2005. X-Ray Motion Analysis of Charge Particles in A Laboratory Mill.
Santomaso, A.C., Ding, Y.L., Lickiss, J.R., York, D.W., 2003. Investigation of the
Ph.D. thesis, University of Cape Town, Cape Town.
granular behaviour in a rotating drum operated over a wide range of rotational
Govender, I., McBride, A.T., Powell, M.S., 2004. Improved experimental tracking
speed. Trans. Inst. Chem. Eng. 81, 936–945.
techniques for validating discrete element method simulations of tumbling
Savage, S.B., 1984. The mechanics of rapid flows. Adv. Appl. Mech. 24, 289–366.
mills. Exp. Mech. 6, 593–607.
Sepulveda, N., Krstulovic, G., Rica, S., 2005. Scaling laws in granular continuous
Govender, I., Pathmathas, T., Richter, M.C., De Klerk, D.N., October 2014. Power
avalanches in a rotating drum. Physica A 356, 178–183.
dissipation modelling in tumbling mills using positron emission particle
Taberlet, N., Richard, P., Hinch, E.J., 2006. S shape of a granular pile in a rotating
tracking. In: Yianatos, J. (Ed.), XXVII International Mineral Processing
drum. Phys. Rev. E 73 (5), 3–6.
Congress. No. 705.
Tang, C., Bak, P., 1988. Critical exponents and scaling relations for self-organised
Harris, C.C., Schnock, E.M., Arbiter, N., 1985. Grinding mill power consumption.
critical phenomena. Phys. Rev. Lett. 60 (23), 2347–2350.
Miner. Process. Technol. Rev. 1, 297–345.
Wildmann, R., Huntley, J., Hansen, J.-P., Parker, D., Allen, D., 2000. Single-particle
Hatano, T., 2007. Power law friction in closely packed granular materials. Phys. Rev.
motion in three dimensional vibrofluidised beds. Phys. Rev. E 62, 3826–3835.
E 75 (6), 060301(R).
Wills, A., 1997. Mineral Processing Technology: An Introduction to the Practical
Henein, H., Brimacombe, J.K., Watkinson, A.P., 1983. Experimental studies of
Aspects of Ore Treatment and Mineral Recovery. Oxford.
transverse bed motion in rotary kilns. Met. Trans. B 14B, 191–205.
Yamane, K., Nakagawa, M., Altobeli, S.A., Tanaka, T., Tsuji, Y., 1998. Steady
Hogg, R., Feurstenau, D., 1972. Power relationships for tumbling mills. SME-AIME
particulate flows in a horizontal rotating cylinder. Phys. Fluids 10, 1419–1427.
Trans. 252, 418–423.
Zik, O., Levine, D., Lipson, S.G., Shtrikman, S., Stavans, J., 1994. Rotational induced
Jain, N., Ottino, J.M., Lueptow, R.M., 2002. An experimental study of the flowing
segregation of granular materials. Phys. Rev. Lett. 73 (5), 644–649.
granular layer in a rotating tumbler. Phys. Fluids 14 (2), 572–582.

You might also like