Sie sind auf Seite 1von 15

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4, pp.

503517 (2014)

NUMERICAL INVESTIGATION ON FLOW GENERATED BY INVENT


MIXER IN FULL-SCALE WASTEWATER STIRRED TANK
A. A. Alleyne*, S. Xanthos*^, K. Ramalingam*, K. Temel*, H. Li* and H. S. Tang*#
*

Department of Civil Engineering, City College of New York, New York 10031, USA
NIREAS International Water Research Center, University of Cyprus, P.O. Box 20537,
1678 Nicosia, Cyprus
#
E-Mail: htang@ccny.cuny.edu (Corresponding Author)

ABSTRACT: This paper investigates behaviors of three dimensional, full-scale flows generated by the newly
developed INVENT HyperClassic mixers in a primary influent channel of a wastewater treatment plant in the New
York City. The ANSYS FLUENT, in conjunction with the standard k- turbulence closure and a multiple reference
frame model, is employed for the investigations. Numerical simulations reveal large-scale circulations on cross
sections perpendicular to the channel direction as a main feature of the flows, and such circulations are expected to
be a key mechanism to enhance their mixing and keep solid particles in suspension. Stronger turbulence mixing in
terms of turbulence kinetic energy and dissipation rate and more intensive mixing in the sense of time average and
evaluated by the Lyapunov exponent are observed in the presence of two mixers than those with a single mixer.
Additionally, a larger inlet velocity of the channel results in an increase in turbulence kinetic energy and dissipation
rate but not necessarily a growth of the Lyapunov exponent.
Keywords:

stirred tank, influent channel, INVENT HyperClassic mixer, mixing, Lyapunov exponent

designed in such a way that it can efficiently keep


solid particles in suspension and generate a flow
well mixed with the strong bottom stream it
produces. The INVENT mixer is rapidly
replacing submersible mixers around the world.
Nevertheless, so far there is little published
literature on flows and associated mixing patterns
generated by such a mixer in actual WWTPs.
Therefore, it is significant to understand the flow
generated by the mixer and thus optimize its
characteristics and performance in actual tanks
and influent channels.
Computational Fluid Dynamics (CFD) is a
powerful tool for simulation of flows within
mixing tanks in WWTPs. CFD modeling based on
the Reynolds averaged Navier-Stokes (RANS)
equations has proven to be effective in prediction
of complex flows and mixing, in particular,
velocity, turbulent kinetic energy, energy
dissipation rate, concentration of suspended
solids, etc., and it has been applied to flows in a
variety of stirred tanks, and numerical simulations
have successfully revealed key features of the
flows and reproduced experimental observations
(e.g., Alexopoulous et al., 2002; Cavadas and
Pinho, 2003; Aubin et al., 2004; Joshi et al.,
2011a; 2011b; Fot et al., 2013). CFD modeling
of flows in tanks and channels needs an
appropriate grid resolution, discretization scheme,

1. INTRODUCTION
Stirred tanks are widely used in mixing processes
in waste water treatment plants (WWTPs)
(Elnekave et al., 2006; Vaiopoulou and Gikas,
2012). The mixing processes are designed to keep
solids in suspension, both organic and inorganic
types, and their effectiveness determines the
performance and efficiency of a WWTP (Kuscu
and Sponza, 2006). In addition, operation of
mixers consumes energy, which becomes
significant when a plant operates round the clock,
and there is a strong impetus to minimize the
energy consumption (Sharma et al., 2011).
Therefore, development of a new generation of
mixers plays a significant role in optimizing
mixing efficiency and energy saving in a WWTP.
The recently introduced INVENT HyperClassic
evolution 6 hyperboloid mixer, together with the
motor assembly outside the water line and light
weight construction, is a significant entrant in the
wastewater industry that has generally been
dominated by submersible mixers (INVENT
2004; INVENT 2010). The impeller of the
hyperboloid mixer rotates near the bottom of a
tank and the motion of its eight fins produces
radially outward streams, which generate largesize circulations and promote mixing and
homogenization of the flow. This mixer is

Received: 26 Oct. 2013; Revised: 9 May 2014; Accepted: 30 Jun. 2014


503

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

turbulence model, and impeller rotating model.


All of these aspects need careful consideration
since they tend to have a momentous influence on
computational expense as well as accuracy of a
simulation (Aubin et al., 2004). Since speed and
memory of computing facilities are major
constraints, as indicated by Delgon and Meyer
(2006), frequently a CFD modeling result is a
compromise between reasonable accuracy and
reasonable computational expense. An example
of such a compromise is the adoption of the socalled multiple reference frame (MRF) in
simulation of flows within stirred tanks (e.g.,
Oshinowo et al., 2000). A flow in a mixing tank
with rotating impellers is essentially unsteady, but
MRF treats it as a steady state at a given instant.
In the MRF approach, the effects of the blade
rotation are accounted for by an extra term in the
governing equations to reflect the virtual motion
of the reference frame. The MRF approach
reduces computational cost substantially, and it
has been proven as an acceptable method for
engineering studies when compared to
experimental data (Koh et al., 2003). However,
flows in stirred tanks are complex, and their CFD
modeling remains challenging. For instance, it is
a difficult task to accurately capture turbulence
characteristics using the standard k- model, and
there could be a considerable discrepancy
between predicted and measured results of
turbulence kinetic energy and dissipation rate
(Joshi et al., 2011a; 2011b). Additionally, current
efforts in CFD modeling of flows in stirred tanks
are
usually
restricted
to
small-scale
configurations, and sparse literature is available
for flows at a scale of O(10) m or a larger size.
The objective of this paper is to predict the fullscale flow generated by the INVENT
HyperClassic mixers installed in an influent
channel at a WWTP in the New York City (NYC)
and obtain an insight of its time-average patterns
and the associated mixing processes. Although
information of flow patterns and mixing
behaviors in simplified settings has been provided
by the manufacturer (INVENT, 2004), a full-scale
flow around the INVENT mixers in actual
influent channels is not available. As indicated
above, a numerical simulation of such flow is
challenging in view of its complicated behaviors
and intensive computing involved. This paper
targets such a full-scale flow and, as a first-of-itskind attempt, makes a prediction of it through a
series of numerical tests and presents its
simulations associated with a single mixer and a
pair of mixers installed in the influent channel of
the plant.

2. METHODOLOGY
2.1 CFD model
The ANSYS FLUENT 14.1 is used for mesh
generation and three dimensional (3D) CFD
modeling in this study (ANSYS Inc., 2011). A
MRF approach is employed and steady-state
simulations are performed with a rotating
reference frame in the impeller region and a
stationary reference frame in the outer region
containing the rest of the tank and the baffles. The
impeller blades, discs (in case of the Rushton
impeller), and baffles are treated as walls with
zero thickness. The non-slip boundary condition
is applied at the walls, and symmetrical boundary
conditions are selected at the top of the tanks to
approximate free surfaces.
In general, structured meshes provide more
accurate solutions than unstructured meshes. In
view of the fact that the configurations of flow
boundaries such as the surface and the fins of the
impeller are complicated, it is difficult to generate
structured meshes of good quality. Therefore,
fully unstructured, non-uniform grids with
primarily tetrahedral body-fitted control volumes
are used in this study. Since large flow gradients
are expected to appear in the region near the
impellers, the grids are refined there to resolve the
flow structures. In the computations, the standard
k- turbulence model and its constants are used. A
second-order QUICK scheme is selected to
discretize governing equations. The discretized
equations are solved iteratively, and the iteration
is deemed to have converged when the residuals
reduce to a value below 10-4. In addition, mesh
convergence tests are performed to ensure mesh
independent solutions.
2.2 Flow analysis
Impeller power number Np has been commonly
used to characterize performance of mixers in
stirred tanks, and it can be expressed as (e.g.,
Delgon and Meyer, 2006):
P
(1)
Np
N 3 D6

where P is the power draw/consumption (W),


the density of liquid (kg/m3), N the impeller
rotational speed (1/s), and D the impeller diameter
(m). P can be evaluated from the rotational
velocity and the torque/moment on the impeller:
P = 2NM

(2)

where M is the moment on the impeller (Nm). A


nomenclature is given in Appendix A.
504

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

Another widely used number is the flow number,


which is a measure of pumping capacity of an
impeller. The flow number, Nq, reads as (Paul et
al., 2004)
Q
(3)
Nq
ND 3
where Q is flow rate produced by the impeller
(m3/s), and its computation is over a surface and
related to types of impellers (Rutherford et al.,
1996; Skocilas et al., 2013).
Turbulence kinetic energy k (m2/s2) and
turbulence dissipation rate (m2/s3) reflect two
main features of a turbulence flow. The total
kinetic energy K (m5/s2) and total turbulence
dissipation rate E (m5/s3) within a flow domain
can be evaluated by
(4)
K kd

to ensure that the CFD model performs


appropriately for flows at different scales. The
grids used for the three cases are listed in Table 1.
For the small-size tank case, three meshes with
different grid spacing are used to test the mesh
convergence of the computed solutions.
(a)

E d

(5)

In a turbulent flow, fluid particles aggressively


mix with each other and their paths may be
chaotic. Generally speaking, the more chaotic the
paths are, the higher the degree of mixing will be.
A commonly used method to quantify the degree
of chaos is the Lyapunov exponent (Lu et al.,
2005):
(6)
S (t ) e t S0
where is the Lyapunov exponent (1/s), t the
time (s), S0 the initial distance between a pair of
selected fluid particles (m), and S(t) the distance
between the two particles when t>0 (m). A
positive and a negative Lyapunov exponent
directly reflect the speed at which a pair of
particles initially adjacent to each other are either
moving away from or towards each other,
respectively.

H/T
1

w/T
1/10

D/T
1/3

d/D
3/4

a/D
1/4

h/D
1/5

Baffle
width,
W (m)

Diameter,
D (m)

Off bottom
clearance,
C/D

Blade
width,
w/D

0.122

0.609

0.5

0.20

b/D
1/5

(b)

Dimension

3. MODEL VALIDATION
3.1 Mixing
tank
configurations
computational meshes

C/T
1/3

T,H=
1.22 m

Hub
diameter,
dh/D
0.13

and
Fig. 1 Tank configurations. (a) Rushton impeller. (b)
45 PBT impeller (thickness of blade is
simplified as zero).

The mixing tanks adopted in calibration and


validation of the CFD model consists of two types
of cylindrical containers with flat bottom, one
with four equally spaced baffles agitated by a
standard six-blade Rushton impeller, which is
referred to as a small- or a medium-size tank, and
the other with a 45pitched blade turbine (PBT)
impeller, which is termed as a large-size tank
(Fig. 1). For the small-, medium-, and large-size
tanks, the diameters of the tanks are respectively
0.15 m, 0.45 m, and 1.22 m, and the rotational
speeds are respectively 200 rpm, 150 rpm, and
200 rpm. Different sizes of tanks are considered

Table1 Meshes for mixing tanks at different sizes


Tank size

Mesh

Number of
mesh elements

Grid 1

960,000

Grid 2

2,625,277

Grid 3

8,200,000

Medium, D=0.45 m

Grid 4

2,625,277

Large, D=1.22 m

Grid 5

7,422,222

Small, D=0.15 m

505

Hub
height,
h/D
0.11

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

Table 2 Power number and flow of small-size tanks.

3.2 Validation
3.2.1 Small-size tank

Approaches

Figure 2 compares the mean values of radial and


tangential velocity obtained on the three grids in
this study, the computed values by Delgon &
Meyer (2006), and experimental data by Wu et al.
(1989) for the flow in the small-size tank with the
Rushton impeller. The differences in terms of root
mean square of the computed results of this study
and Delgon & Meyer (2006) against the
experimental data by Wu et al. (1989) are also
presented in this figure. The mean values of
velocity are obtained by averaging its value at 24
locations (every 15 degrees) along the
circumferential direction and at a distance of
2.775 cm, or r/T=0.185, away from the axis of the
tank. It is seen from the figure that the modeling
results compare reasonably well with the
measurements by Wu et al. (1989). The solutions
and their differences with the experimental data in
the figure show that mesh refinement slightly
improves the comparison of the computed mean
velocities with the measured data, and the
solution on Grid 3 can be considered as a mesh
independent solution for the averaged radial and
tangential velocities. In addition, with regard to
the measured data, the simulation from this study
provides more accurate solutions than that given
by Delgon and Meyer (2006), who used 1.9x106
elements and a first-order upwind scheme. Table
2 indicates that the power number and the flow
number computed by this study are consistent
with the values measured and computed by other
authors.

Np

Nq

This study with Grid 3

5.02

0.71

Delgon and Meyer (2006)

5.07

Falk and Villermaux (1997)

4.5-5

Lane and Koh (1997)


Costes and Couderc (1988)

4.5
0.73

3.2.2 Medium-size tank


Escudi and Lin (2003) conducted laboratory
experiments for flows in a medium-size tank,
while Delafosse et al. (2008) carried out a large
eddy simulation (LES). The solution of the flow
field and the computed velocity and power
number using Grid 4, together with those obtained
in other investigations, are presented in Fig. 3 and
Table 3. The figure and the table indicate that, in
comparison with the results from other
approaches, the computational results from this
study present a reasonable approximation. Fig. 3
also shows that the LES modeling by Delafosse et
al. (2008) provides the best comparison with the
measurement data. In particular, the LES
modeling captures the flow with more accuracy
not only near the blades, which is located at
z/T=3.4, but also in regions away from them.
Nevertheless, given that LES is very expensive
and possibly formidable for flows within actual
mixing tanks in a WWTP, it is contended to adopt
the modeling approach of this study, which works
reasonably well at an affordable computational
cost.

Fig. 2 Comparison of velocities obtained from this


study (Re=1.6x106, defined in terms of tip
velocity and diameter of the impeller) and
others in case of the small-size tank. Utip is the
velocity at the impellers tip.

Fig. 3 Comparison of velocities of flows within


medium-size tanks obtained from this study
(Re=1.1x107, in terms of tip velocity and
diameter of the impeller) and others.

506

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

Table 3 Power number of medium-size tanks.


Approaches

Np

This study

5.49

Delafosse et al. (2008)

5.50

accuracy. This computational setup will thus be


the basis for modeling of flows around the fullscale INVENT HyperClassic mixers in the
following section.
Table 4 Power number and flow number of large-size
tanks.

3.2.3 Large-size tank


There is lack of both experimental and
computational data for flows in large-scale stirred
tanks because of the high cost associated with
field tests and computations. In this study,
computation is made for flows in a large-size tank
with the details shown in Table 1. The solution
and a measurement for flows in tanks that are
geometrically similar but smaller in size are
compared with the results obtained with LES
modeling in Fig. 4. Similar to results obtained by
other authors, the peak of the computed axial
velocity occurs around 2r/D = 0.8 but is
marginally under predicted. The computed power
number and flow number, together with those
calculated in Beshay et al. (2001) and measured
by Wu et al. (2001), are presented in Table 4.
Beshay et al. (2001) used a baffled tank with
0.8m in diameter and a C/D ratio of 0.5, which
are identical to those in this study, and Wu et al.
(2001) carried out a series of experiments with
different types of PBTs using a Laser Doppler
Anemometer.

Approaches

Np

Nq

This study

1.21

0.76

Beshay et al. (2001)

1.20

Wu et al. (2001)

0.76

4. SIMULATIONS OF FLOWS IN STIRED


TANKS WITH INVENT MIXERS
4.1 Background
In an attempt to identify a more cost effective
strategy, the New York City Environmental
Protection (NYCEP) has initiated a pilot study by
installing four INVENT HyperClassic mixers
within a section of an influent channel at the
Wards Island WWTP (CCNY Institute for
Municipal Waste Research, 2011). The
dimensions of the channel are 27.74 m in length,
3.66 m in width, and 3.54 m in depth, and the
configuration and sizes of INVENT mixer are
shown in Fig. 5a. The mixers operate at a speed
up to 26 rpm, and they are powered at 2.24 kW
per mixer. The rotating mixers are located 0.254
m above the bottom of the channel, and it is
expected that they produce a radially outwards
flow in conjunction with micro-vortices at the
bottom (Fig. 5b). The INVENT HyperClassic is
a new type of mixer, and there exists a gap
between available knowledge and the detailed
flow patterns generated by it. The intent of this
paper is to confirm the results of this pilot study
and extend the effort to other wastewater
treatment plants in NYC. Hence, it is necessary to
make a reasonable prediction of flow behaviors
associated with the INVENT mixers and better
understand their impact on performance and
efficiency of the mixers.

Fig. 4 Computed profile and LES modeling by


Roussinova et al. (2003) for mean axial
velocity within large-size tank (Re=2.3x108,
which is defined by tip velocity and diameter
of impeller), measurement of Kresta (1991) in
a smaller size tank, and their comparisons.

4.2 INVENT mixer in a closed compartment


In summary, the simulations of this study for the
flows in small-, medium-, and large-scale tanks
compare favorably with available experimental
and computational data. Therefore, it is
anticipated that the setup of the FLUENT model,
including the meshes, algorithms, and relevant
parameters, are appropriate and they can capture
large-scale flow structures with reasonable

Before predicting flow patterns in the influent


channel at the Wards Island WWTP, a
preliminary simulation was carried out for a
single INVENT mixer rotating at 26 rpm in a 3.66
m x 3.66m x 3.54 m box, which has the same
width as the actual channel at the plant. A mesh
with 1,308,719 elements was used. Fig. 6 shows
507

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

the flow pattern obtained by the simulation. It is


seen that the computed flow pattern is similar to
what the manufacturer of INVENT has
described as shown in Fig. 5b; the flow has largescale, i.e., from bottom- to top-wall circulation
zones above the mixer, and indeed there are micro
vorticies near the blade tip and the bottom. The
power number computed for this flow is 0.31.
This value is approximately 10% less than 0.35
that is observed by Fot et al. (2013), and the
under prediction may be explained by the
discussion of Cavadas and Pinho (2004).

4.3 INVENT mixers in influent channel


4.3.1 Geometry, boundary condition & meshes
In this study, CFD modeling of flows associated
with a single INVENT mixer and subsequently
two mixers in a full-scale influent channel was
carried out in an attempt to obtain an
understanding of the large-scale flow patterns and
mixing efficiency in the NYCs WWTP. The
configurations and the dimensions of a section of
the channel and mixers being studied are shown
in Fig. 7. In this figure, there are two mixers, one
at x=0 with a clockwise rotation (top views) and
the other at x=-6.93 m with an anti-clockwise
rotation at 26 rpm. The flow enters the channel
from the left end and exits at the right end, as well
as through the four outlets on one of its sidewall.

(a)

(b)

Fig. 7 Configuration of INEVNT mixers and a


section of influent channel.

In the computation, an inlet velocity is imposed at


the entrance, and outlet conditions are selected for
the downstream exit and the outlets on the side
wall. At the entrance, the turbulence kinetic
energy k0 and dissipation rate 0 are specified as
k0=0.005U02 ; 0=Ck0(3/2)/(0.03R)
(7)
where U0 is the inlet velocity, and R is the
hydraulic radius. C is a constant in the standard
k- model and it equals 0.09. In reference with the
flow conditions in the influent channel at the
Wards Island WWTP, U0=0.19 m/s is used. Eq.
(7) is consistent with what is proposed by Pun and
Spalding (1976) and Lai and Yang (1997) for
developing turbulence flows at inlets.
Fully unstructured, non-uniform grids with
tetrahedral body-fitted control volumes are used
(Fig. 8). Since the region near the rotating
impellers is expected to have a large gradient for
the flow variables, the mesh elements used for
this region are built with fine grid spacing to
resolve the flow structures and ensure that the
associated patterns of mixing are captured
appropriately.

Fig. 5 INVENT mixer and designed flow patterns. (a)


Configuration and dimensions. (b) Flow
patterns associated with a INVENT mixer
(INVENT, 2004).

Fig. 6 Simulation of velocity magnitude and


streamline generated by an INVENT mixer in
a closed container. Re=5.4x106, in terms of tip
velocity and diameter of the mixer.

508

Engineering
Applications
of Computational
Mechanics
8, 4,
No.
(2014) (2014)
Engineering
Applications
of Computational
FluidFluid
Mechanics
Vol. Vol.
8, No.
pp.4 503517

(a)

(b)

Fig. 8 A mesh for influent channel and INVENT mixers. (a) Mesh at center, vertical plane. (b) Mesh at horizontal
plane at 0.254 m above channel bottom.

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)

(i)

(j)

Fig. 9 Simulation of velocity magnitude and streamline in the presence of a single mixer by coarse and fine mesh,
which respectively have 1,297,760 and 2,648,684 elements. Re=7.0x105, in terms of entrance velocity and
width of the channel. (a), (c), (e), (g), (i) Coarse mesh. (b), (d), (f), (h), (j) Fine mesh. (a), (b) x=0. (b), (c) x=-13
m. (e), (f) z=0. (g), (h) y = 0.254 m. (i), (j) y=3.5 m.
509

Engineering
Applications
of Computational
Mechanics
8, No.
4 (2014)
Engineering
Applications
of Computational
FluidFluid
Mechanics
Vol.Vol.
8, No.
4, pp.
503517 (2014)

4.3.2 Results and analysis

(a)

To begin with, an initial simulation of the flow


associated with a single INVENT mixer, which is
designated as No. 1 mixer as shown in Fig. 7.
This simulation uses both a coarse and a fine
mesh, and the result is presented in Fig. 9. The
figure shows the velocity field on the x cross
sections at the mixer and the exit, the central
vertical plane, a horizontal plane near the bottom
wall, and a plane near the top wall. As seen from
the predicted streamlines in Figs. 9a and 9b, on
the cross section at the mixer, the water moves
upwards near the right wall, crosses the shaft, and
travels to the left, and then moves downwards to
the bottom, forming a large-scale circulation. This
circulation is indeed large; it is from the right to
the left wall and from the bottom to the top wall,
and at the size of the cross section. This
circulation shows no symmetry, and its velocity
magnitude is relatively small in comparison to the
rest of the region next to the lateral wall. Clearly,
these flow patterns differ considerably from that
in the ideally designed case as shown in Fig. 5b,
which indicates two large-scale circulation zones
symmetrical about the axis, and that in the close
box as shown in Fig. 6, which shows larger
velocity next to the lateral wall. Such large-scale
cross-section circulation is further shown in Figs.
9c and 9d, and it is expected to occur in the whole
channel. It is believed that these large-scale
circulations will promote large-scale (e.g., from
the bottom- to the top-wall) mixing within the
flow and suspension of solid particles. On the
central vertical plane, both of the two solutions
have a clockwise vortex in front of the rotator and
a low velocity zone at upper corner at the exit
(Figs. 9e and 9f). Additionally, Figs. 9g and 9h
reveal that, next to the surface of the impeller,
there is a region where high velocity occurs,
indicating that the water near the surface is
rotating together with the impeller at a much
higher speed in comparison with water away from
it.
It is seen from all of the plots in Fig. 9 that the
results of the coarse and the fine mesh have very
similar solutions, indicating that the solution of
the fine mesh can be considered as a mesh
independent solution. The mesh convergence is
further confirmed by a quantitative comparison of
solutions obtained on the coarse and the fine mesh
as shown in Fig. 10.
Subsequently, simulation for the flow with two
mixers is carried out, as shown in Fig. 11. Again,
as in the flow with a single mixer, there is a largescale circulation at spatial size of the cross

(b)

(c)

Fig. 10 Distribution of velocity magnitude along


normal directions of walls obtained on coarse
and the fine mesh. (a) y=1.77 m, z=0 m. (b)
x=-3.47 m, z= 0 m. (c) x=-3.47 m, y=1.77 m.

section, from the bottom/top to the top/bottom,


and from the left/right to the right/left wall, at the
x cross section at No. 1 mixer (Fig. 11a).
Similarly, at the x cross section across No. 2
mixer and the exit, there are large-scale
circulations (Figs. 11b and 11c). It is expected
that these circulations in x cross sections will be
the main mechanism for the liquid to move up
and down, promoting large-scale mixing (e.g.,
from the bottom- to the top-wall, from the left- to
the right-wall) and keeping particles in suspension
in the actual influent channel. In addition, there is
a clockwise vortex on the vertical plane in front
of each mixer (Fig. 11d). Near the bottom of the
channel and in front of No.2 mixer, there is a
backwards flow due to its rotation (Fig. 11e).
In general, the upstream mixer has an effect on
the flow field at the downstream mixer, but not
vice versa. This is illustrated by comparing Figs.
9a, and 11a; the flow field at the upstream mixer

510

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

similarity between the flow field of single mixer


and that of two mixers. In the case of two mixers,
the flow field at the downstream mixer is
apparently different from that at the upstream,
implying the former is affected by the latter (Figs.
11a, 11b).
The flow associated with two mixers is similar to
that with a single mixer at a few regions in space.
For instance, in comparison to the single mixer, as
indicated above, the flow field near No. 1 mixer
remains about the same after No. 2 mixer is
added. In addition, it is hard to differentiate flows
with one mixer and two mixers by the flow
patterns of the flow fields at the top of the
channels (Figs. 9h, 11f). Yet, the two flows have
differences. These differences can be seen at x
cross sections at the exits of the channels as
shown in Figs. 9d and 11c. At the exits, although
both of them have similar patterns with a
clockwise vortex at the lower left corner and an
anti-clockwise vortex at the upper right corner,
the pathlines of the flow associated with two
mixers have more numbers of rotations, and it is
expected thus it has a stronger mixing.
A 3D view of the vortices in case of two mixers is
shown in Fig. 12, with the blue and red color
representing vorticity in opposite directions. It is
observed that strong vorticity occurs at the rims of
the blades as well as at the outer edge of the
rotators. In addition, high vorticity also occurs
near the outlets on the lateral wall. Since
magnitude of vorticity reflects rotation speed of a
fluid element about its own axis, a larger
magnitude of vorticity indicates a lager rotation
speed and thus a stronger mixing at scales of fluid
particles. Therefore, Fig. 12 shows that there is
strong mixing of the flow near the impellers and
outlets on the wall. In addition, this figure also
presents streamlines extending in the downstream
direction in the form of spirals, which confirms
the large-scale circulations or secondary flows on
x cross sections discussed previously.

(a)

(b)

(c)

(d)

(e)

(f)

Fig. 11 Velocity magnitude and streamline at vertical


central and a horizontal plane in case of two
mixers. Re=7.0x105, in terms of entrance
velocity and width of channel. (a) x=0. (b)
x=-6.9 m. (c) x=-13 m. (d) z=0. (e) y = 0.254
m. (f) y=3.5 m.

is almost the same as that of the single mixer


flow, and almost completely unaffected by the
presence of the downstream mixer. As seen in
Figs. 9d, 9f and Figs. 11c, 11d, in the
neighborhood of No. 1 mixer, there is also a good

Fig.12 3D representation of vorticity in vertical


direction and streamlines in case of two
mixers.
511

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

Turbulence kinetic energy and dissipation rate


represent important characteristics of flows. Fig.
13 shows that both of these variables present
higher values near the impellers. In order to
illustrate how they change with flow conditions,
values of K and E in the flow fields are shown in
Table 5 for different flow conditions. The table
indicates that the values of both K and E double
approximately when the number of impellers
increases from one to two. When the inlet
velocity is doubled in the case of two mixers,
together with much higher values for turbulence
kinetic energy and dissipation rate at the entrance,
the values of K and E of the flow field increase,
too, as shown in the table.
Fig. 14 Initial locations of the released particles on a
cross section, which is shown in Fig. 7.

(a)

S(t) between the two particles of each pair is


measured. Using Eq. (6), the Lyapunov exponent
of this pair is then evaluated, and the results are
shown in the Appendix. A summary of the
Lyapunov exponent of these pairs of particles is
given in Table 6. Table 6 shows that the average
values of the Lyapunov exponent for all of the 50
pairs of fluid particles are positive in case of one
impeller, two impellers, and two impellers with
the double of the inlet velocity. As explained in
Section 2, a positive component implies that a
pair of fluid particles move away from each other,
and the larger the component, the faster they
separate from each other. Since the average
values are positive, it is concluded that a pair of
fluids particles that are close to each other in all
of the three situations tend to move away from
each other. It should be noted that since the values
of the Lyapunov component are evaluated using
the solutions for time-averaged velocities in the
RANS equations, this discussion on the chaotic
behavior of fluid particles is essentially in the
sense of time average also.

(b)

Fig.13

Distribution of turbulence kinetic energy and


dissipation rate in case of two mixers.

Table 5 Computed total kinetic energy and total


dissipation at different entrance flow
conditions and different number of mixers.
U0=0.19 m/s.
One mixer
& U0
Two mixers
& U0
Two mixers
& 2U0

k0(m2/s2)
1.87e-4

0(m2/s3)
6.49e-6

K (m5/s2)
1.15

E (m5/s3)
0.92

1.87e-4

6.49e-6

2.04

1.80

7.46e-4

5.19e-5

5.11

2.99

The chaotic behavior of the flows can be


described using the Lyapunov exponent. To
understand such behavior, the simulation results
are used to compute the Lyapunov exponent. The
computation is carried out on 10 x cross sections,
evenly distributed along the channel. The first
section is located at the entrance, and the second
section is depicted in Fig. 7. At each of such cross
sections, five pairs of fluid particles with S0=0.05
m are released, with one at the center and the
other four at a certain distance from it in the east,
west, south, and north direction. Fig. 14 shows the
locations of the particles on the second cross
section. After a lapse of one second, the distance

Table 6 Summary of computation for Lyapunov


exponent.
One mixer& U0

Two mixers& U0 Two mixer & 2U0

i (s )

0.004

0.045

0.038

i (s 1 )

0.032

0.077

0.074

In general, larger values for turbulence kinetic


energy and dissipation rate are associated with a
stronger turbulence mixing, which is due to
turbulence fluctuations at small temporal and
spatial scales. Table 5 shows that in comparison
with those of one impeller, the flow with two
512

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

impellers has higher values for turbulence kinetic


energy and dissipation rate, and thus the latter
involves a better degree of turbulence mixing. In
this sense, the table also shows that, in the
presence of two impellers, an increase in the inlet
velocity, together with an increase of turbulence
kinetic energy and dissipation rate at the entrance,
does lead to a stronger turbulence mixing. By
definition, the absolute value of Lyapunov
exponent directly reflects the speed at which the
particles are either moving towards or away from
each other, and thus the degree of fluid particle
mixing, in the sense of time average. Table 6
shows that, in comparison to the flow with a
single mixer, the flow with two impellers has a
larger average of the absolute values of the
exponent, and thus is more chaotic and involves a
larger degree of mixing, again, in the logic of
time average. However, an increase of inlet
velocity produces an average and an absolute
value of the exponent similar to those in the
previous case, and hence it may not necessarily
lead to a more efficient mixing in the sense of
time average, although small-scale turbulence
mixing, which is characterized by kinetic energy
and dissipation, is more pronounced.

INVENT HyperClassic mixers, and it will shed


light in flows within the influent channel installed
with INVENT mixers in the NYC WWTP and
those in other similar WWTPs.
The current study is an ongoing work, and will be
improved in the next-step by investigation and
overcoming the limitations of the current study. In
this paper, flows are treated as steady state using
MRF. For a more realistic simulation, unsteady
effects should be taken into account, and this can
be done using a sliding mesh technique in
FLUENT (ANSYS Inc., 2011). It is expected that
simulations on the basis of the RANS equations in
this study have captured features of averaged
flows at large scales generated by INVENT
impellers at the WWTP. However, it remains a
challenge for this approach to accurately account
for turbulence and mixing (Joshi et al., 2011a;
2011b). Although a preliminary study has been
made on the involved mixing process with regard
to turbulence kinetic energy and dissipation rate
and the Lyapunov exponent, it will be further
investigated from other important aspects such as
characteristic mixing time (Paul et al., 2004). This
paper simulates single phase flows, but effects of
bio-particles should be included as their
concentration becomes high (e.g., Yu et al.,
2011). Field measurement within the influent
channel at the WWTP in NYC has been planned
to capture both flow fields and suspended solids
concentrations, and it is expected that its results
may be used to further validate and calibrate our
CFD studies. Given the promising performance of
the CFD approach demonstrated in this paper, all
of these issues will be kept as topics for future
work.

5. CONCLUDING REMARKS
On the basis of validation against both
computational data and laboratory measurements
for a set of flows in stirred tanks, FLUENT is
employed to simulate flows associated with the
newly developed INVENT HyperClassic mixers
installed in an influent channel at a NYC WWTP.
One of the main features of the flows within the
influent channel installed with INVENT mixers
revealed by the simulations is the presence of
large-scale circulations or secondary flows on x
cross sections, from the bottom- to the top-wall,
and from the left- to the right-wall, which is
believed to favorably promote mixing and be a
key mechanism to keep solid particles in
suspension. It is demonstrated that, in comparison
with a single impeller, two impellers introduce
not only higher total turbulence kinetic energy
and total dissipation rate but also a larger value
for the Lyapunov exponent, and thus stronger
mixing of the flows. An increase of inlet velocity
also increases total turbulence kinetic energy as
well as dissipation rate, and thus small-scale
turbulence mixing. However, it does not
necessarily lead to higher values for the
Lyapunov exponent, or, mixing in the sense of
time averages. The effort in this paper provides
an insight on mixing of flows associated with the

ACKNOWLEDGEMENT
This research is sponsored by the NYCEP.
Special thanks go to Mr. Yuklong Ma, the Project
Manager, BWT and Mr. Steven Moltz, the
Process Engineer, at the Wards Island WWTP,
NYC for their support in executing this project.
We are grateful to Dr. Y. Andreopoulos for his
valuable suggestions. Assistance from the
FLUENT technical team is also acknowledged.
NOMENCLATURE
abC-

513

impeller blade width (m) (Rushton


impeller)
impeller hub diameter (m) (Rushton
impeller)
off bottom impeller clearance (m)

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

C - constant in the standard k- turbulence


model
d - impeller disc diameter (m) (Rushton
impeller)
dh - impeller hub diameter (m) (PBT impeller)
E - total rate of dissipation of turbulence
(m2/s3)
D - impeller diameter (m)
S0 - initial distance between a pair of selected
fluid particles (m)
S(t) - the distance between the two particles when
t>0 (m)
h - impeller blade height (m) (Rushton
impeller)
h - hub height (m) (PBT impeller)
H - height of fluid in the tank (m)
K - total kinetic energy (m2/s2)
kturbulent kinetic energy (m2/s2)
k0 - initial turbulent kinetic energy (m2/s2)
M - moment on the impeller (Nm)
N - impeller speed (s-1)
Np - impeller power number (dimensionless)
Nq - impeller power number (dimensionless)
P - power drawn by impeller (W)
Q - flow rate generated by an impeller (m3/s)
R - hydraulic radius (dimensionless)
r - radial distance from impeller centre (m)
Re - Reynolds number
ttime (s)
T - tank diameter (m)
Utip - impeller tip velocity (m/s)
Uo - channel entrance velocity (m/s)
W - baffle width (m)
w - blade width (m) (PBT impeller)
Pair
Cross No.
section
i
1
2
3
1
4
5
6
7
8
2
9
10
11
12
13
3
14
15
16
17

Coordinate of point
1 (m)
x1
6.85
6.85
6.85
6.85
6.85
4.85
4.85
4.85
4.85
4.85
2.85
2.85
2.85
2.85
2.85
0.85
0.85

y1
1.75
1.75
1.75
0.85
2.55
1.75
1.75
1.75
0.85
2.55
1.75
1.75
1.75
0.85
2.55
1.75
1.75

z1
-0.85
0
0.85
0
0
-0.85
0
0.85
0
0
-0.85
0
0.85
0
0
-0.85
0

z-

Greek symbols
0 -

rate of dissipation of turbulence energy


(m2/s3)
initial rate of dissipation of turbulence
energy (m2/s3)
density (kg/m3)
tangential distance from impeller center
plane (rad)
dynamic viscosity (kg/ms)
Lyapunov exponent (1/s)
flow domain

Abbreviations
CFD - computational fluid dynamics
LES large eddy simulation
MRF - multiple reference frames
PBT - pitched blade turbine
RANS - Reynolds-averaged Navier-Stokes
WWTP - wastewater treatment plant
LYAPUNOV EXPONENT AT SELECTED
LOCATIONS
Ten cross sections are evenly distributed in the
main stream direction. Lyapunov exponent is
evaluated at five locations on each cross section,
totally 50 locations at the ten cross sections. At
every location, a pair of particles are released, one
at (x1,y1,z1) and the other at (x2,y2,z2), and then
they move with the flow.

coordinate of point
2 (m)
x2
6.85
6.85
6.85
6.85
6.85
4.85
4.85
4.85
4.85
4.85
2.85
2.85
2.85
2.85
2.85
0.85
0.85

vertical distance from tank bottom (m)

y2
1.75
1.75
1.75
0.9
2.6
1.75
1.75
1.75
0.9
2.6
1.75
1.75
1.75
0.9
2.6
1.75
1.75

z2
-0.8
0.05
0.9
0
0
-0.8
0.05
0.9
0
0
-0.8
0.05
0.9
0
0
-0.8
0.05
514

2 mixers & U0

2 mixers & U0

i (1/s)
-0.000115
-0.000116
-0.000071
0.000275
-0.000365
-0.000476
-0.003266
-0.002129
0.009972
-0.010306
-0.047167
-0.035288
-0.060826
-0.049891
-0.021387
0.037203
-0.049979

i (1/s)
-0.000081
0.000079
-0.000153
0.000271
-0.000298
-0.002476
-0.004016
-0.001434
0.00972
-0.011825
-0.066008
-0.035577
-0.069541
-0.043269
-0.019351
0.088795
-0.040063

2 mixers &
2U0
i (1/s)
-0.000093
0.000012
-0.000198
0.000304
-0.000413
0.000009
0.000244
-0.00094
0.009722
-0.010058
0.00422
0.007651
-0.002717
0.029116
-0.043351
0.23286
-0.115684

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

10

18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50

0.85
0.85
0.85
-1.15
-1.15
-1.15
-1.15
-1.15
-3.15
-3.15
-3.15
-3.15
-3.15
-5.15
-5.15
-5.15
-5.15
-5.15
-7.15
-7.15
-7.15
-7.15
-7.15
-9.15
-9.15
-9.15
-9.15
-9.15
-11.15
-11.15
-11.15
-11.15
-11.15

1.75
0.85
2.55
1.75
1.75
1.75
0.85
2.55
1.75
1.75
1.75
0.85
2.55
1.75
1.75
1.75
0.85
2.55
1.75
1.75
1.75
0.85
2.55
1.75
1.75
1.75
0.85
2.55
1.75
1.75
1.75
0.85
2.55

0.85
0
0
-0.85
0
0.85
0
0
-0.85
0
0.85
0
0
-0.85
0
0.85
0
0
-0.85
0
0.85
0
0
-0.85
0
0.85
0
0
-0.85
0
0.85
0
0

0.85
0.85
0.85
-1.15
-1.15
-1.15
-1.15
-1.15
-3.15
-3.15
-3.15
-3.15
-3.15
-5.15
-5.15
-5.15
-5.15
-5.15
-7.15
-7.15
-7.15
-7.15
-7.15
-9.15
-9.15
-9.15
-9.15
-9.15
-11.15
-11.15
-11.15
-11.15
-11.15

1.75
0.9
2.6
1.75
1.75
1.75
0.9
2.6
1.75
1.75
1.75
0.9
2.6
1.75
1.75
1.75
0.9
2.6
1.75
1.75
1.75
0.9
2.6
1.75
1.75
1.75
0.9
2.6
1.75
1.75
1.75
0.9
2.6

0.9
0
0
-0.8
0.05
0.9
0
0
-0.8
0.05
0.9
0
0
-0.8
0.05
0.9
0
0
-0.8
0.05
0.9
0
0
-0.8
0.05
0.9
0
0
-0.8
0.05
0.9
0
0

-0.013442
0.120852
0.021372
0.130088
-0.078761
-0.061999
0.09892
0.052887
0.01944
0.08216
-0.033909
-0.075293
0.101699
0.007107
0.056696
0.005136
-0.053938
0.046151
-0.0308
0.021847
0.032525
-0.028278
0.025451
-0.011908
-0.002029
0.015314
-0.005598
0.010958
-0.010435
-0.00661
0.010203
-0.002965
0.008563

-0.016352
0.123763
0.023285
0.092918
-0.088919
-0.033082
0.088144
0.043086
0.057668
0.021481
0.063197
-0.084934
0.114645
-0.036908
0.078446
0.047307
0.321767
0.035662
-0.044695
0.32479
0.098081
1.286008
0.058615
-0.006054
0.009737
-0.075772
-0.034815
0.023109
-0.016649
-0.025482
-0.014778
-0.025561
0.025636

-0.140907
0.214428
-0.047573
-0.052288
-0.010937
-0.074886
0.231069
0.037947
0.041643
-0.082245
-0.009249
-0.089281
0.125253
-0.021913
0.070416
0.050995
0.240004
0.128825
-0.016711
0.085874
0.075596
0.922968
0.184939
0.009747
0.020785
-0.051638
-0.087413
0.058603
-0.015415
0.000058
-0.026089
-0.022216
0.016316

4. Beshay KR, Kratna J, Fot I, Brha O


(2001). Power input of high-speed rotary
impellers. Acta Polytechnica 41(6):18-23.
5. Cavadas AS, Pinho FT (2004). Some
characteristics of stirred vessel flows of
dilute polymer solutions powered by a
hyperboloid impeller. Canadian Journal
of Chemical Engineering 82(2): 289302.
6. CCNY Institute for Municipal Waste
Research (2011). Odor and Hydraulic Study
of Mixers in the Wards Island Wastewater
Treatment Plant Preliminary Settling Tank
Influent Channel. Report, New York City
Environmental Protection Bureau of
Wastewater Treatment, under contract #
WI-278, Dec. 9, 2011.
7. Costes J, Couderc JP (1988). Study by laser
Doppler anemometry of the turbulent flow
induced by a Rushton turbine in a stirred

REFERENCES

1. Alexopoulous
AH,
Maggiorios
D,
Kiparissides C (2002). CFD analysis of
turbulence non-homogeneity in mixing
vessels: A two-compartment model.
Chemical Engineering Science 57:17351752.
2. ANSYS, Inc. (2011). ANSYS FLUENT
User's Guide, Release 14.0, November,
2011.http://cdlab2.fluid.tuwien.ac.at/LEHR
E/TURB/Fluent.Inc/v140/flu_ug.pdf.
3. Aubin J, Fletcher DF, Xuereb C (2004).
Modeling turbulent flow in stirred tanks
with CFD: the influence of the modeling
approach,
turbulence
model
and
numerical
scheme.
Experimental,
Thermal and Fluid Science 28:431-445.

515

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

8.

9.

10.

11.

12.

13.

14.

15.

16.

17.

18.

tank: influence of the size of the units - I.


mean flow and turbulence. Chemical
engineering and Science 43 (10): 2751
2764.
Delafosse A, Line A, Morchain J, Guiraud
P (2008). LES and URANS simulations of
hydrodynamics in mixing tank: Comparison
to PIV experiments. Chemical Engineering
Research and Design 86:1322-1330.
Delgon DA, Meyer CJ (2006). CFD
modelling of stirred tanks: Numerical
considerations.
Minerals
Engineering
19:1059-1068.
Elnekave M, Tufekci N, Kimchie S, Shelef
G (2006). Tracing the mixing efficiency of
a primary mesophilic anaerobic digester in
a municipal wastewater treatment plant.
Fresenius Environmental Bulletin 15(9B),
Special Issue: SI. 1098-1105.
Escudi R, Lin A (2003). Experimental
analysis of hydrodynamics in a radially
agitated tank. AIChE Journal 49:585603.
Falk L, Villermaux J (1997). Numerical
scale-up and design of high efficiency
mixers for control and optimization of the
yield and selectivity in chemical reactors.
Applied Thermal Engineering 17(8-10):
845-859.
Fot I, Seichter P, Pel L, Rieger F, Jirout T
(2013). Blending characteristics of high
speed rotary impellers. Chemical and
Process Engineering 34 (4):427-434.
INVENT (2004). Manual for the
HYPERCLASSIC Stirring Systems of
NY Bowery bar. Erlangen, Germany.
INVENT (2010). HYPERCLASSIC Mixer/
Aerator.
http://www.invent-et.com/
hyperclassic-mixer-aerator-3/.
Joshi JB, Nere NK, Rane CV, Murthy BN,
Mathpati CS, Patwardhan AW, and Ranade
VV (2011a). CFD simulation of stirred
tanks: comparison of turbulence models.
Part i: radial flow impellers. Canadian
Journal of Chemical Engineering 89:26-83.
Joshi JB, Nere NK, Rane CV, Murthy BN,
Mathpati CS, Patwardhan AW, Ranade VV
(2011b). CFD simulation of stirred tanks:
comparison of turbulence models, part ii:
axial flow impellers, multiple impellers and
multiphase dispersions. Canadian Journal
of Chemical Engineerin. 89:754-816.
Koh PTL, Schwarz MP, Zhu Y, Bourke P,
Peaker
R,
Franzidis
JP
(2003).

19.

20.

21.

22.

23.

24.

25.

26.

27.

28.

516

Development of CFD models of mineral


flotation cells. Proc. Third International
Conference on CFD in the Minerals and
Process Industries, Melbourne, 171-175.
Kresta SM (1991). Characterization and
Prediction of the Turbulent Flow in Stirred
Tanks.
Ph.D.
Thesis,
MacMaster
University, Hamilton, Canada.
Kuscu OS, Sponza DT (2006). Treatment
efficiencies of a sequential anaerobic
baffled reactor (ABR)/completely stirred
tank reactor (CSTR) system at increasing pnitrophenol and COD loading rates. Process
Biochemistry. 41(7):1484-1492.
Lai JCS, Yang C Y (1997). Numerical
simulation of turbulence suppression:
Comparisons of the performance of four k-
turbulence models. Int. J. Heat and Fluid
Flow 18:575-584.
Lane GL, Koh PTL (1997). CFD simulation
of a Rushton turbine in a baffled tank. Proc.
Inter. Conf. on CFD in Mineral & Metal
Processing and Power Generation, CSIRO,
377-385
Lu J, Yang GL, Oh H, and Luo ACJ (2005).
Computing Lyapunov exponents of
continuous dynamical systems: method of
Lyapunov vectors. Chaos, Solitons and
Fractals 23:18791892.
Oshinowo L, Jaworski Z, Dyster KN,
Marshall E, Nienow AW(2000). Predicting
the tangential velocity field in stirred tanks
using multiple reference frames (MRF)
model
with
validation
by
LDA
measurements, Proc. 10th Euro. Conf. Mix.
281289.
Paul EL, Atiemo-Obeng VA, Kresta SM
(2004). Handbook of Industrial Mixing:
Science and Practice (Vol. 1). WileyInterscience.
Pun WM, Spalding DB (1976). A General
Computer Program for Two-Dimensional
Elliptic Flows. HTS/76/2, Imperial College.
Roussinova V, Kresta SM, Weetman R
(2003). Low frequency macro instabilities
in a stirred tank: scale-up and prediction
based on large eddy simulations. Chemical
Engineering Science 58:2297-2311.
Rutherford K, Mahmoudi SMS, Lee KC,
Yianneskis M (1996). The influence of
Rushton impeller blade and disk thickness
on the mixing characteristics of stirred

Engineering Applications of Computational Fluid Mechanics Vol. 8, No. 4 (2014)

29.

30.

31.

32.

33.

34.

vessels. Institution of Chemical Engineers


74(3):369-378.
Sharma AK, Guildal T, Thomsen HR,
Jacobsen BN (2011). Energy savings by
reduced mixing in aeration tanks: results
from a full scale investigation and long
term
implementation
at
Avedoere
wastewater treatment plant. Water Science
and Technology 64(5):1089-1095.
Skocilas J, Fort I, Jirout T (2013). A study
of cfd simulations of the flow pattern in
anyagitated system with a pitched blade
worn turbine. Chemical and Process
Engineering 34 (1):39-49.
Vaiopoulou E, Gikas P (2012). Effects of
chromium on activated sludge and on the
performance of wastewater treatment
plants: a review Water Research.
46(3):549-570.
Wu H, Patterson GK (1989). Laser-Doppler
measurements of turbulent flow parameters
in a stirred mixer., Chemical Engineering
Science 44(10):2207-2221.
Wu J, Zhu Y, Pullum L (2001). The effect
of impeller pumping and fluid rheology on
solid suspension in a stirred vessel.
Canadian Journal of Chemical Engineering
79:177-186.
Yu L, Ma J, Chen S (2011) Numerical
simulation of mechanical mixing in high
solid anaerobic digester. Bioresource
Technology 102(2):1012-1018.

517

Das könnte Ihnen auch gefallen