You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/286667822

Simulation of turbulent flow of a fine dispersive slurry

Article  in  Inzynieria Chemiczna i Procesowa · January 2010

CITATIONS READS

4 230

1 author:

A. Bartosik
Politechnika Świętokrzyska
32 PUBLICATIONS   157 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Project made at the University of Saskatchewan View project

Project made at Kielce University of Technology Faculty of Management and Computer Modelling View project

All content following this page was uploaded by A. Bartosik on 18 January 2017.

The user has requested enhancement of the downloaded file.


CHEMICAL AND PROCESS ENGINEERING
2010, 31, 67–80

ARTUR BARTOSIK*

SIMULATION OF TURBULENT FLOW


OF A FINE DISPERSIVE SLURRY

Kielce University of Technology, al. Tysiąclecia P.P., 7, 25-314 Kielce, Poland

Solid−liquid turbulent flow in a pipeline has been simulated for concentration of solids ranging from
0 to 45 % by volume. Water was the carrier liquid, the solid phase constituted fine particles with mean
particle diameters of tens of microns. Mathematical model comprising the time averaged momentum
equation, k-ε turbulence model with a modified turbulence damping function, Bingham rheological model
has successfully been verified. The results of simulation indicate that concentration of solids affects
strongly local velocity at the pipe wall while the pressure gradient and thickness of the viscous sublayer
increase linearly for CV in the range 0–25%, and exponentially for CV > 25%.

Symulowano turbulentny przepływ faza stała–faza ciekła w rurze w zakresie objętościowej koncen-
tracji fazy stałej 0–45%. Ciecz nośną stanowiła woda, fazę stałą – drobnodyspersyjne cząstki o średniej
średnicy kilkudziesięciu mikrometrów. Pozytywnie zweryfikowano model matematyczny, złożony
z uśrednionego po czasie równania pędu, modelu turbulencji k-ε, zmodyfikowanej funkcji tłumiącej
turbulencję przy ścianie i modelu reologicznego Binghama. Wyniki symulacji wskazują, że prędkość
lokalna przy ścianie zależy od koncentracji fazy stałej, jednostkowy spadek ciśnienia natomiast i grubość
podwarstwy lepkiej zwiększają się liniowo dla CV w zakresie 0–25% i wykładniczo dla CV > 25%.

1. INTRODUCTION

Solid–liquid turbulent flows of fine particles commonly occurring in chemical engi-


neering processes, power plants, food, and mining industry belong to main challenges of
computational fluid dynamics leaving several unsolved problems. Many approaches have
been investigated to predict their mechanism, however mixture theory models, based on
a rigorous fluid mechanics framework [1, 2] are the most general.
Mathematical modelling of solid–liquid turbulent flows is far away from knowl-
edge gathered for Newtonian flows. When developing mathematical model of a New-
tonian flow, the starting point are the Navier–Stokes equations exactly describing

____________
*
E-mail: artur.bartosik@tu.kielce.pl
68 A. BARTOSIK

laminar and turbulent flow field [3, 4]. Taking into account the time averaged Navier
–Stokes equations (Reynolds equations) one can find comprehensive library of avail-
able turbulence models appropriate for describing turbulent stress tensor, used rou-
tinely by pipeline designers throughout the world [5–7].
Rheological experiments with a solid–liquid flow with fine particles exhibit yield
stress. Thus, the slurry is not Newtonian anymore and appropriate two- or three-
parameter rheological model should be incorporated into the mathematical model. If
solid particles are sufficiently small, the relaxation time t* defined by Eq. (1), is below
the Kołmogorov [8] time scale (te), defined as the ratio of the characteristic length of
a swirl to its characteristic velocity. Hence, for the Stokes number (defined by Eq. (2))
below unity (St < 1), the particles can follow a carrier liquid motion. In such a case,
turbulent diffusion produces a uniform distribution of concentration of solids within
the pipe. Thus such fine particle slurries are of a non-settling type and the flow can be
treated as a single phase flow with increased viscosity and density. For the Stokes
number higher than the unity (St > 1), the solids will not follow a carrier liquid motion
causing that slip velocity and/or turbulence generation could appear
d P2 ρ P
t* = (1)
18 μ L

t*
St = (2)
te
In the absence of reliable turbulence data related to the immediate vicinity of the pipe
wall over a wide range of concentration of solids, it is difficult to suggest new turbulence
models for the slurry flow. However, it is possible to modify existing turbulence models
comparing calculated and measured global parameters, which are particularly relevant in
engineering application.
Wilson and Thomas [9] reported that in the case of fine dispersive slurry flow
thickening of the viscous sublayer tends to increase throughput velocity and thus pro-
motes drag reduction. Based on their analysis, Bartosik [10] formulated a new turbu-
lence damping function which together with the time averaged momentum equation,
two-equation turbulence model, and a proper rheological model (Bingham, Casson or
Herschel–Bulkley), describes a turbulent flow of fine dispersive slurry. The mathe-
matical model was used to develop database of pressure drops for various combination
of velocities, pipe diameters, and rheological parameters. The mathematical model
was successfully verified for pressure drop and velocity distribution in comprehensive
range of rheological parameters [10–13]. Comparison of results of simulation with
those of measurements showed some advantages of the Herschel–Bulkley rheological
model compared to the Bingham and the Casson models [14]. The differences were
not substantial, however. Nevertheless, it is important to examine influence of the
concentration of solids on the pressure drop, velocity distribution, and thickness of the
Simulation of turbulent flow of fine-dispersive slurry 69

viscous sublayer. The viscous sublayer and the velocity distribution are closely related
to temperature distribution, which is of special interest in food industry and pipeline
erosion.
The main aim of the paper was to examine influence of the concentration of solids
in a turbulent flow of fine dispersive slurry on the pressure drop, the velocity distribu-
tion, and thickness of the viscous sublayer in a pipeline taking into account the
mathematical model supported by the modified turbulence damping function and the
Bingham rheological model.

2. LITERATURE REVIEW

Among turbulence models of slurry flow one equation turbulence models can be
taken into account such as those of Danon et al. [16], Roco and Shook [17], Roco and
Balakrishnan [18], Wu [19], Li and Zhou [20] and Mishra et al. [21] or two equation
k-ε-Ap models of Yulin [22] or Ling et al. [23]. Danon et al. [16] built for example
one-equation k-l turbulence model using empirical turbulence length scale l. In two-
equation k-ε-Ap turbulence model of Yulin [22], k and ε are kinetic energy of turbu-
lence, and its dissipation rate, respectively, similarly as in a standard turbulence model
for a single phase flow. Ap is an algebraic equation describing solid phase. The model
has successfully been examined but only for low values of concentration of solids.
Stainsby and Chilton, [24, 25] developed a hybrid model, the apparent viscosity
being calculated by the Herschel–Bulkley rheological model at low strain rates and by
the Bingham model at high strain rates. Using the time averaged momentum equation
and low Reynolds number k-ε model of Launder and Sharma [26], they could predict
the pressure drop and velocity distribution in a fine dispersive slurry flow. They did
not introduce any changes into k-ε turbulence model declaring that the related error of
prediction of pressure drop using their mathematical model is lower than 15%. How-
ever their hybrid model was successfully examined only for low concentrations of
solids and low yield stresses (the maximum slurry density was 1105 kg/m3).
Recently, Cui and Grace [27] made a literature review on newest experiments and
modelling of solid−liquid flow with special attention to paper pulp. They noted pro-
gress in techniques of determining fibre flocculation and laminar-turbulent transition.
They noted that new models based on computational fluid dynamics (CFD) have been
published, however their potential is limited by lack of properly defined suppression
and/or amplification of turbulence and solid−liquid interactions.
As was mentioned above, in the case of limited access to reliable experimental
data on turbulence in the close vicinity of a pipe wall over a wide range of concentra-
tions of solids, it is difficult to suggest appropriate turbulence models for a slurry flow.
However, modifications of existing turbulence models are possible, based on the com-
parison of calculated and measured global parameters such as pipeline head losses and
70 A. BARTOSIK

cross-sectional time-averaged velocity profiles being very important in designing pipe-


lines, pumps impellers, and for proper calculation of heat transfer processes.

3. MATHEMATICAL MODEL

It is assumed that the transported medium is fine dispersive slurry exhibiting the
yield stress, with the mean particle diameter d50 in the range of 1–50 μm, density of
solids ρP between 1000 kg/m3 and 2650 kg/m3, concentration of solids CV from 0 to
45% by volume, and slurry constant apparent viscosity and density.
For fully developed and axially symmetrical pipe flow, the time averaged momen-
tum equation can be described as:

1 ∂ ⎡ ⎛ ∂U ⎞ ⎤ ∂p
⎢r ⎜ μ − ρ u′v′ ⎟ ⎥ = (3)
r ∂r ⎣ ⎝ ∂r ⎠ ⎦ ∂x
The component of the turbulent stress tensor in Eq. (3) was derived from the
Boussinesque hypothesis:
∂U
− ρ u′v′ = μt (4)
∂r
The turbulent viscosity μt in Eq. (4) was determined with support of the dimen-
sionless analysis [26]:
ρ k2
μt = f μ (5)
ε
where k and ε were calculated using the Launder and Sharma turbulence model [26]:
• equation for kinetic energy of turbulence k:
2 2
1⎡ ⎛ μt ⎞ ∂ k ⎤ ⎛ ∂U ⎞ ⎛ ∂ k 1/ 2 ⎞
r
⎢ ⎜ μ + ⎟ ⎥ + μ t⎜ ⎟ = ρε + 2 μ ⎜ ⎟ (6)
r ⎣⎢ ⎝ σ k ⎠ ∂ r ⎦⎥ ⎝ ∂r ⎠ ⎝ ∂r ⎠

• equation for dissipation rate of kinetic energy of turbulence ε:

2
1⎡ ⎛ μt ⎞ ∂ε ⎤ ε ⎛ ∂U ⎞
⎢r ⎜ μ + ⎟ ⎥ + C1 μt ⎜ ⎟
r ⎢⎣ ⎝ σ ε ⎠ ∂ r ⎥⎦ k ⎝ ∂r ⎠
2
(7)
ρε 2
⎛ ∂ 2U ⎞
= C2 ⎡⎣1 − 0.3exp ( − Ret2 ) ⎤⎦ − 2νμt ⎜ 2 ⎟
k ⎝ ∂r ⎠
Simulation of turbulent flow of fine-dispersive slurry 71

The turbulent Reynolds number in Eq. (7) was defined using the dimensionless
analysis:

ρ k2
Ret = (8)
με

A crucial point in the mathematical model is determining of the turbulence damp-


ing function fμ, called also the wall damping function (Eq. (5)). It is an empirical func-
tion with low values at a pipe wall, causing decrease in the turbulent viscosity and in
consequence decrease in the component of the turbulent stress tensor (Eq. (4)), which
is in accordance with results of measurements. For a fine dispersive slurry flow with
the yield stress, the wall damping function was determined based on the comparison
between predictions and measurements [10]:

⎡ ⎛ τ0 ⎞⎤
⎢ −3.4 ⎜ 1 + ⎟ ⎥
f μ = 0.09exp ⎢ ⎝ τw ⎠⎥ (9)
⎢ ⎛ Re ⎞2 ⎥
⎢ ⎜ 1+ t ⎟ ⎥
⎣⎢ ⎝ 50 ⎠ ⎦⎥

The turbulence damping function described by Eq. (9), depends on the dimen-
sionless yield stress (τ0/τw) and approaches the standard damping function, proposed
by Launder and Sharma, as the yield stress approaches zero. However, upon the in-
creasing yield stress at the same flow conditions (τw = const), the turbulence damping
function decreases the turbulent viscosity (Eq. (5)) and in consequence decreases the
turbulent shear stresses (Eq. (4)) at the pipe wall. This is consistent with the Wilson
and Thomas hypothesis [9] who reasoned that the viscous sublayer becomes thicker,
compared to a single phase flow at the same flow conditions, if a slurry with fine solid
particles is considered.
Taking into account the Bingham rheological model, which is simple, commonly
used, and quite adequate for fine dispersive slurry with the yield stress, one can calcu-
late the apparent viscosity μap from:
τ = μapγ (10)

taking into account that:


γ = 0 for τ ≤ τ0 (11)
and

τ = τ 0 + μ PLγ for τ >τ0 (12)


72 A. BARTOSIK

Taking into account Eqs. (10) and (12), the apparent viscosity for γ > 0 is:

μ PL
μap = (13)
⎛ τ0 ⎞
⎜1− ⎟
⎝ τ ⎠

Instead of the dynamic viscosity coefficient μ (Eqs. (3) and (6)−(8)), the apparent
viscosity μap is used in the mathematical model. Equation (13) has limitation that the
yield stress has to be lower than the shear stress. However, this can take place in the
close vicinity of the symmetry axis due to the fact that the shear stress varies linearly
from its maximum value at the pipe wall to zero in the symmetry axis. This means that
at some distance from the symmetry axis the shear stress can be lower than the yield
stress. Thus, it was assumed that the apparent viscosity is constant across the pipe
having the same value as at the pipe wall and the apparent viscosity in the mathemati-
cal model was calculated for the Bingham rheological model as:

μ PL
μap = (14)
⎛ τ0 ⎞
⎜1− ⎟
⎝ τw ⎠

The wall shear stress in Eq. (14), is determined from balance of forces acting on
the unit pipe length, thus it was calculated as:
dp D
τw = (15)
dx 4
The mathematical model comprises three partial differential equations (Eqs. (3),
(6) and (7)), together with the complimentary equations (Eqs. (4), (5), (8), (9)) and the
apparent viscosity (Eq. (14)). Numerical calculations were performed for known
dp/dx. The turbulence constants in the turbulence model are the same as those in the
Launder and Sharma model: C1 = 1.44, C2 = 1.92, σk = 1.0, σε = 1.3. The mathematical
model assumes non slip velocity at the pipe wall, i.e. U = 0, and k = 0, ε = 0, and axi-
ally symmetrical conditions were applied at the pipe centre, thus dU/dr = 0, dk/dr = 0
and dε/dr = 0. The mathematical model was solved by the finite difference scheme
using the elaborated computer code. Differential grid of 80 nodal points distributed on
the radius of the pipe has been used. Majority of the nodal points were localized in
close vicinity of the pipe wall. The number of nodal points was set experimentally to
ensure nodally independent computation. The mathematical model was examined for
a comprehensive range of rheological parameters defined by the Bingham, Casson,
and Herschel–Bulkley models, and for a few pipe diameters confirming satisfying
agreement of predictions with measurements [10–15].
Simulation of turbulent flow of fine-dispersive slurry 73

4. RESULTS OF NUMERICAL SIMULATION

Considering a slurry flow with fine particles, the strength of the floc structure is usu-
ally considered to be the source of yield stresses in Bingham fluids. Particle size, their
shape, and concentration of solids are extremely important since these determine the total
amount of surface available for floc formation deciding about the yield stress [28].

Fig. 1. Dependence of the yield stress on concentration of solids


in limestone slurry ([28], p. 69)

Fig. 2. Dependence of the plastic viscosity on concentration of solids


in limestone slurry ([28], p. 69)
74 A. BARTOSIK

Thomas found that the yield stress varies approximately as (CV)3/(dP)2 if the parti-
cle diameters are between 0.4 and 17 μm, [29]. Examples of such effects in industrial
slurry are shown in Figs. 1 and 2. Both figures present experimental data on the yield
stress and the plastic viscosity of limestone as functions of concentration of solids
[28]. About 67% of volume of limestone was finer than 44 μm. Thus, we can assume
that the slurry is fine dispersive. In order to perform numerical simulation of depend-
ences of concentration of solids on the pressure drop, velocity distribution, and thickness
of the viscous sublayer it was necessary to determine empirical dependence of τ0 = f(CV),
and μPL = f(CV). For this purpose, the experimental data, presented in Figs. 1 and 2, were
interpolated by dashed lines. The first experimental point (for the lowest concentration of
solids) is taken from the Slatter experiments [30], because the Shook and Roco, [28] ex-
periments were performed for moderate and high concentrations of solids. Experimental
data of Slatter [30] were performed for fine dispersive slurry flow with mean particles
diameter d50 = 28 μm.
Taking into account the mathematical model described by Eqs. (3), (6), (7), to-
gether with the complementary equations (4), (5), (8), (9), (14) and empirical relation
of τ0 = f(CV) and μPL = f(CV), presented in Figs. 1 and 2, it is possible to perform nu-
merical simulation of dp/dx = f(CV), U = f(y/R), U+ = f(R+), and δ = f(CV). Numerical
simulations were made for a fine dispersive slurry at constant temperature T = 293 K,
pipe diameter D = 0.159 m, density of solids ρP = 2500 kg/m3, and constant bulk ve-
locity of the slurry (Ub)m = 4.3 m/s. The slurry bulk velocity was chosen as 4.3 m/s to
ensure that the flow is turbulent if CV = 45%. The concentration of solids CV was var-
ied from 0% to 45%.

Fig. 3. Dependence of the pressure drop on concentration of solids


at the constant bulk velocity (Ub)m = 4.30 m/s

Figure 3 presents results of numerical simulation of the dependence of the pres-


sure gradient on the concentration of solids CV. For CV in the range 0–25%, the plot
dp/dx = f(CV) is practically linear, while for CV > 25% it is exponential. For that case,
Simulation of turbulent flow of fine-dispersive slurry 75

the Reynolds number Reap defined by Eq. (16), was 688 100 for CV = 0% (water flow),
and Reap = 7250 for CV = 45%. The laminar-turbulent transition for such fine disper-
sive slurry flow was at Reap ≈ 5000 [31], thus the flow for the data presented in Fig. 3
is turbulent
ρ m (U b )m D
Reap = (16)
μ ap

Fig. 4. Water and slurry velocity profiles at chosen concentrations


of solids and for constant bulk velocity (Ub)m = 4.30 m/s

Results of numerical simulations of velocity profiles, performed for CV in the range


of 0−45%, exhibit qualitative and quantitative differences compared to Newtonian liquid
flow at the same flow conditions (Fig. 4). Local slurry velocity at the pipe wall drasti-
cally decreases if concentration of solids increases. Thus, the shape of the velocity pro-
file is less steep. This is the effect of viscous forces which depend on the apparent vis-
cosity and is due to damping of turbulence in a fine dispersive slurry flow, [31].
Decrease of local slurry velocity at the pipe wall is compensated by increase of local
velocity in the core region as a result of constant bulk velocity equal to 4.3 m/s. Qualita-
tive and quantitative changes of slurry velocity profiles compared to Newtonian liquid
flow are pronounced on logarithmic scale presented in Fig. 5. Dimensionless distance
from the pipe wall and dimensionless velocity are defined by the equations:

τw ⎛ τ ⎞
(R − r) ρm ⎜1 − 0 ⎟
ρm ⎝ τw ⎠
R+ = (17)
μ PL
76 A. BARTOSIK

U
U+ = (18)
τw
ρm

It is seen in Fig. 5 that for all values of concentration of solids and for R+ ≤10,
U+ = R+, while for R+ > 10 the values of U+ increase upon increasing concentration
of solids. If R+ > 10 and high concentration of solids, the differences between loga-
rithmic velocity profiles of Newtonian and Bingham slurry become substantial.

Fig. 5. Logarithmic velocity profiles of water and slurry


at various concentrations of solids; (Ub)m = 4.30 m/s

Recent analysis of Bartosik [14] indicates that in a fine dispersive slurry flow, the
thickness of the viscous sublayer R+ depends on the yield stresses and can exceed 10
if the yield stress is sufficiently high, while in case of Newtonian liquid flow we ex-
pect that viscous sublayer exists for 0 < R+ ≤ 5. Such changes of velocity profiles
between the Bingham and Newtonian slurry flows in a close vicinity of the pipe wall
play a crucial role in prediction of friction.
As was mentioned above, we expect that the concentration of solids affects the
thickness of the viscous sublayer. Thus, assuming that the viscous sublayer will exist
at the maximum value of R+ = 5, which is true in the case of the Newtonian flow, and
at R+ = 10 in the case of a fine dispersive slurry flow, both distances from the pipe
wall (δ) being estimated based on numerical predictions at a constant bulk velocity
(Ub)m = 4.3 m/s. The results are presented in Fig. 6.
It is seen in Fig. 6 that for CV in the range 0–25%, the viscous sublayer at
R+ = 5 and R+ = 10 increases linearly, while for CV > 25% it is exponential.
Simulation of turbulent flow of fine-dispersive slurry 77

Fig. 6. Thicknesses of the viscous sublayer at various


concentrations of solids and for chosen R+; (Ub)m = 4.30 m/s

For CV = 45% (ρm = 1674 kg/m3), which corresponds to the Reynolds number Reap
= 7250, the thickness of the viscous sublayer δ at R+ = 10 is equal to 4 mm, while for
the Newtonian liquid flow it is 0.06 mm at R+ = 10 and 0.03 mm at R+ = 5. Such
changes of the thickness of the viscous sublayer are very important in the case of heat
transfer processes and erosion of pipelines.

5. CONCLUSIONS

The numerical simulation of influence of concentration of solids on the properties


of turbulent pipe flow for fine dispersive slurry with the yield stress allows formulat-
ing the following conclusions:
For concentration of solids CV ranging from = 0% to 25% the dependence dp/dx =
f(CV) is linear, while for CV > 25% it is exponential.
Upon increasing concentration of solids, the local slurry velocity at the pipe wall
drastically decreases making the velocity profile less steep. This is also pronounced if
logarithmic velocity profiles are considered. This is due both to effect of viscous
forces which depend on the apparent viscosity and damping of turbulence at the pipe
wall.
The concentration of solids affects strongly the thickness of the viscous sublayer.
It increases linearly for 0% < CV ≤25%, and exponentially for CV > 25%. Increase of
the thickness of the viscous sublayer causes decrease of the friction (reduction of tur-
bulent stresses tensor) and finally gives lower pressure losses.
78 A. BARTOSIK

Changes of the thickness of the viscous sublayer are important in the case of pre-
diction of pipeline characteristics, heat transfer processes, and erosion of pipelines.
Increase of the thickness of the viscous sublayer is probably due to the lift forces.
Lift forces which exist at the pipe wall in a slurry flow causes the particles to be
pushed away from the pipe wall towards the symmetry axis. Higher particles diameter
results in higher lift forces. Turbulent diffusion causes the particles with the higher
diameters to be replaced by particles with the smaller diameters which enhance the
viscosity of the slurry at the pipe wall. It results in higher viscous forces, reduces the
turbulent fluctuations which thicker viscous sublayer.

SYMBOLS

Ci – constant in the Lauder and Sharma turbulence model, i = 1, 2


CV – volume fraction of solids averaged in the cross section, %
dP – solid particle diameter, mm
d50 – weighted-average particle diameter, mm
D – inner pipe diameter, m
fμ – turbulence damping function at the pipe wall
k – kinetic energy of turbulence, m2/s2
l – turbulence length scale, m
p – static pressure, Pa
r – distance from symmetry axis, m
R – pipe radius, m
R+ – dimensionless distance from the pipe wall
Ret – turbulent Reynolds number
Reap – Reynolds number for apparent viscosity
ReL – Reynolds number for carrier liquid
St – Stokes number
te – duration of microscale swirls, s
t* – relaxation time of solid particle, s
T – temperature, K
U – velocity component in x direction, m/s
U+ – dimensionless velocity
u′, v′ – fluctuating components of velocity U and V, m/s

GREEK SYMBOLS

δ – thickness of the viscous sublayer at R+ = 5 or R+ = 10, mm


γ – strain rate, s–1
ε – rate of dissipation of kinetic energy of turbulence, m2/s3
μ – dynamic viscosity coefficient, Pa·s
μap – apparent viscosity, Pa·s
μPL – plastic viscosity in the Bingham rheological model, Pa·s
ν – kinematic viscosity coefficient, m2/s
ρ – density, kg/m3
σi – diffusion coefficients in k-ε turbulence model, i = k, ε
τ – shear stress, Pa
Simulation of turbulent flow of fine-dispersive slurry 79

τ0 – yield shear stress, Pa


τw – wall shear stress, Pa
SUBSCRIPTS

ap – apparent
b – bulk (cross-section averaged value)
i – index, i = 1, 2, …
L – liquid
m – slurry (solid–liquid mixture)
p – particle (solid phase)
t – turbulent
w – solid wall

REFERENCES

[1] SOO S.L., Multiphase Fluid Dynamics, Science Press, Beijing, 1990.
[2] BIRD R.B., STEWART W.E., LIGHTFOOT E.N., Transport Phenomena, Wiley, New York, 1960.
[3] PROSNAK W.J., Equations of Classical Fluid Dynamics (in Polish), Warsaw, 2006.
[4] SPALDING D.B., Recent Advances in Numerical Methods in Fluids, C. Taylor, K. Morgan (Eds.),
Pineridge Press, Swansea, 1980, 139.
[5] SPALDING D.B., Imperial College of Science and Technology, Rep. No. CFD/82/4, Mech. Engng.
Dept., London, 1983,.
[6] HANJALIĆ K., POPOVAC M., HADŹIABDIĆ M., Int. J. Heat Fluid Flow, 25, 2004, 1047.
[7] KOŁMOGOROV A.N., Izv. Acad. Sci. SSSR, Ser. Fiz. 6, No. 1–2, 1942, 56.
[8] WILSON K.C., THOMAS A.D., Int. J. Chem. Eng., 63, 1985, 539.
[9] BARTOSIK A., Chem. Proc. Eng., 27, 2006, 623.
[10] BARTOSIK A., Proc. 10th Int. Conf. Numerical Methods in Laminar and Turbulent Flow, C. Taylor,
J.T. Cross (Eds.), Pineridge Press, 10, 1997, 265.
[11] BARTOSIK A., HILL K., SHOOK C., Proc. 9th Int. Conf. Transport and Sedimentation of Solid
Particles, Part 1, 1997,69.
[12] BARTOSIK A., Chem. Proc. Eng., 22, 2001, 223.
[13] BARTOSIK A., 12th Int. Symp. Freight Pipelines joined with 12th Int. Conf. Transport and
Sedimentation of Solid Particles, Prague, Acad. Sci. Czech Republic, September 2004, 167.
[14] HINZE J.O., Prog. in Heat Mass Transfer, 6, 1971, 433.
[15] BARTOSIK A., Flow, Turbulence and Combustion, Springer, Berlin, 84, 2009, 277.
[16] DANON H., WOLFSHTEIN M., HETSRONI G., Int. J. Multiphase Flow, 3, 1977, 223.
[17] ROCO M.C., SHOOK C.A., Can. J. Chem. Eng., 61, 1983, 494.
[18] ROCO M.C., BALAKRISHNAN N., J. Rheology, 29, 1985, 431.
[19] WU Y., Proc. ASME Fluids Engineering Division Summer Meeting, Part 1, San Diego, July 7–11,
1996, CA, 265.
[20] LI Y., ZHOU L.X., Proc. ASME Fluids Engineering Division Summer Meeting, Part 1, San Diego,
July 7–11, 1996, CA, 311.
[21] MISHRA R., SINGH S.N., SESHADRI V., Powder Handling Proc., 10, 1998, 279.
[22] YULIN W.U., ASME Fluid Engineering Division Conference, FED, 236, 1996, 265.
[23] LING J., SKUDARNOV P.V., LIN C.X., EBADIAN M.A., Int. J. Heat Fluid Flow, 24, 2003, 389.
[24] STAINSBY R., CHILTON R.A., Proc. 2nd CFDS Int. User Conf., Pittsburgh, 1994, 259.
[25] STAINSBY R., CHILTON R.A., Proc. BHR Group, Hydrotransport-13, 1996, 21.
[26] LAUNDER B.E., SHARMA B.I., Lett. Heat Mass Transfer, No. 1, 1974, 131.
[27] CUI H., GRACE J.R., Int. J. Multiphase Flow, 33, 2007, 921.
80 A. BARTOSIK

[28] SHOOK C.A., ROCO M.C., Slurry Flow: Principles and Practice, Butterworth-Heinemann, Boston,
1991.
[29] THOMAS D.G., AIChE J., 9, 1963, 310.
[30] SLATTER P.T., Transitional and Turbulent Flow of Non-Newtonian Slurries in Pipes, PhD Thesis,
University of Cape Town, 1994.
[31] BARTOSIK A., Arch. Therm., 29, 2008, 69.

SYMULACJA TURBULENTNEGO PRZEPŁYWU


DROBNODYSPERSYJNEJ HYDROMIESZANINY

Turbulentny przepływ hydromieszaniny ciecz−cząstki stałe występuje powszechnie w przemyśle


energetycznym, wydobywczym i przetwórczym. Obecność cząstek stałych w przepływającej hydromie-
szaninie może powodować tłumienie lub wzrost turbulencji, która zależy przede wszystkim od średnicy
cząstek stałych. Gdy hydromieszanina jest drobnodyspersyjna, o średniej średnicy cząstek stałych od
kilku do kilkudziesięciu mikrometrów, pojawia się w niej zazwyczaj zjawisko tłumienia turbulencji.
Efektem tego są mniejsze jednostkowe straty ciśnienia w stosunku do ekwiwalentnego przepływu cieczy
newtonowskiej. Obecność drobnych frakcji fazy stałej powoduje również pojawienie się nienewtonow-
skich właściwości i progu płynięcia w hydromieszaninie.
Celem niniejszej pracy było zbadanie wpływu koncentracji fazy stałej w turbulentnym przepływie
drobnodyspersyjnej hydromieszaniny Binghama w przewodzie kołowym na jednostkowy spadek ciśnie-
nia, rozkład lokalnej prędkości oraz na grubość podwarstwy lepkiej z zastosowaniem modelu matema-
tycznego zawierającego funkcję uwzględniającą wzrost tłumienia turbulencji przy ścianie. Przedstawiono
wyniki symulacji numerycznej wpływu koncentracji fazy stałej na właściwości turbulentnego przepływu
hydromieszaniny, w której ciecz nośną stanowi woda, a fazę stałą drobnodyspersyjne cząstki o średniej
średnicy d50 ok. 50 μm i ρP = 2500 kg/m3, w zakresie objętościowej koncentracji fazy stałej 0–45%.
Symulację przeprowadzono za pomocą wcześniej zweryfikowanego modelu matematycznego, który
stanowią uśrednione po czasie równanie zachowania pędu, model turbulencji k-ε, funkcja tłumiąca turbu-
lencję przy ścianie oraz model reologiczny Binghama, który wykorzystano do wyznaczenia lepkości
pozornej hydromieszaniny. Model matematyczny zweryfikowano w szerokim zakresie parametrów re-
ologicznych drobnodyspersyjnej hydromieszaniny dla kilku średnic przewodu. Równania różniczkowe
modelu matematycznego rozwiązano metodą różnic skończonych z zastosowaniem ekspansyjnej siatki
różnicowej, aby zapewnić obecność większości punktów węzłowych w sąsiedztwie ściany przewodu.
Wyniki symulacji numerycznej wykazały istotne różnice jakościowe i ilościowe rozkładu prędkości
lokalnej przy ścianie w przepływie drobnodyspersyjnej hydromieszaniny w porównaniu z przepływem
newtonowskiej cieczy. Wyniki symulacji dowiodły, że ze wzrostem koncentracji fazy stałej zmniejsza się
prędkość lokalna hydromieszaniny przy ścianie, powodując, że profil prędkości jest mniej stromy niż
podczas przepływu cieczy newtonowskiej. Wzrost stężenia fazy stałej powoduje, że zarówno spadek
ciśnienia, jak i grubość podwarstwy lepkiej rosną liniowo w zakresie 0–25% i wykładniczo dla CV > 25%.
Wpływ koncentracji fazy stałej na grubość podwarstwy lepkiej w przepływie turbulentnym odgrywa
pierwszoplanową rolę w procesie wymiany ciepła i erozji przewodów. Z tych względów określenie
wpływu koncentracji fazy stałej na jednostkowy spadek ciśnienia oraz profil prędkości przy ścianie stałej
i grubość podwarstwy lepkiej w nienewtonowskiej hydromieszaninie mającej próg płynięcia należy do
istotnych wyzwań obliczeniowej dynamiki przepływu.

Received 24 November 2009

View publication stats

You might also like